ebook img

The Anderson-Mott transition PDF

120 Pages·2011·5.18 MB·English
by  
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview The Anderson-Mott transition

The Anderson-Mott transition D. Belitz Department ofPhysics, and Materials Science Institute, University ofOregon, Eugene, Oregon 97409 T.R. Kirkpatrick lnsti tute forPhysical Science and Technology, and Department ofPhysics, University ofMaryland, College Park, Maryland 20742 The interacting disordered electron problem isreviewed with emphasis on the quantum phase transitions that occur in a model system and on the field-theoretic methods used to describe them. An elementary discussion ofconservation laws and diffusive dynamics isfollowed by adetai1ed derivation ofthe extended nonlinear sigma model, which serves asan effective field theory for the problem. Ageneral scaling theory ofmetal-insulator and related transitions isdeveloped, and explicit renormalization-group calculations for the various universality classes are reviewed and compared with experimental results. Adiscussion ofper- tinent physical ideas and phenomenological approaches to the metal-insulator transition not contained in the sigma-model approach is given, and phase-transition aspects ofrelated problems, like disordered su- perconductors and the quantum Hall effect, are discussed. The review concludes with alist ofopen prob- lems. CONTENTS 4. Symmetry considerations 298 a. Spontaneously broken symmetry 298 b. Externally broken symmetry 299 I. Introduction 262 c. Remaining soft modes and universality II. Diffusive Electrons 266 classes 300 A. Diffusion poles 266 C. Summary ofresults for noninteracting electrons 301 1~ The quasiclassical approximation for electron 1. Renormalization ofthe nonlinear sigma model 301 transport 267 2. Composite operators, ultrasonic attenuation, a. A phenomenological argument for diffusion 267 and the HalI conductivity 302 b. Linear response and the Boltzmann equation 268 3. Conductance fluctuations 303 c. Diagrammatic derivation ofthe diffusive den- IV. Scaling Scenarios for the Disordered Electron Problem 303 sity response 270 A. Metal-insulator phase-transition scenarios 304 d. Beyond the quasiclassical approximation 272 1. The noninteracting problem 304 2. Disorder renormalization ofelectron-electron 2. The interacting problem 306 and electron-phonon interactions 273 B. Possible magnetic phase transitions 308 a. Dynamical screening ofdiffusive electrons 273 V. Metal-Insulator Transitions in Systems with Spin-Flip b. Coupling ofphonons todiffusive electrons 274 Mechanisms 310 B. Perturbation theory for interacting electrons; early A. Theory 310 scaling ideas 276 1. Systems with magnetic impurities 310 III. Field-Theoretic Description 276 2. Systems in external magnetic fields 313 A. Field theories for fermions 276 3. Systems with spin-orbit scattering 315 1. Grassmann variables 277 a. Asymptotic critical behavior 315 2. Fermion coherent states 278 b. Logarithmic corrections to scaling 317 3. The partition function for many-fermion systems 278 B. Experiments 319 B. Sigma-model approach to disordered electronic sys- 1. Transport properties 320 tems 279 2. Thermodynamic properties and the tunneling 1. The model in matrix formulation 279 density ofstates 323 a. The model in terms ofGrassmann fields 279 3. Nuclear-spin relaxation 324 b. Spinor notation and separation in phase VI. Phase Transitions in the Absence ofSpin-Flip Mecha- space 281 nisms 325 c. Decoupling ofthe four-fermion terms 284 A. Theory 326 d. The model in terms ofclassical matrix fields 286 1. The loop expansion 326 2. Saddle-point solutions 289 a. One-loop results 326 a. The saddle-point solution in the noninteract- b. Absence ofdisorder renormalizations 328 ing case and Hartree-Fock theory 289 c. Two-loop results 329 b. BCS-Gorkov theory as asaddle-point solu- d. Resummation ofleading singularities 330 tion ofthe field theory 291 2. The pseudomagnetic phase transition 332 3 The generalized nonlinear sigma model 292 a. Numerical solution 332 ~ a. An effective theory for diffusion modes 292 b. Analytic scaling solution 332 b. Cxaussian theory 293 c. Renormalization-group analysis 333 c. Loop expansion and the renormalization d. Summary 335 group 295 3. The metal-insulator transition 337 d. Fermi-liquid corrections and identification of B. Experiments 338 observables 296 1. Transport properties 338 e. Long-range interaction 298 2. Thermodynamic properties 340 Reviews ofModern Physics, Vol.66,No.2,April 1994 0034-6861/94/66(2) /261(120)/$17.00 1994 The American Physical Society 261 262 D. Belitz and T.R.Kirkpatrick: The Anderson-Mott transition VII. Destruction ofConventional Superconductivity by Dis- at zero temperature, so that the Auctuations determining order 341 the critical behavior are quantum mechanical rather than A. Disorder and superconductivity: Abrief overview 341 thermal in nature. B. Theories for homogeneous disordered superconduc- Metal-insulator transitions can be divided into two tors 343 1. Generalizations ofBCSand Eliashberg theory 343 categories (see, for example, Mott, 1990). In the first a. Enhancement and degradation ofthe mean- category, some change in the ionic lattice, such as a field T, 343 structural phase transition, leads to a splitting ofthe elec- b. Breakdown ofmean-field theory 345 tronic conduction band and hence to a metal-insulator 2. Field-theoretic treatments 345 transition. In the second category the transition is purely a. The Landau-Ginzburg-Wilson functional 345 electronic in origin and can be described by models in b. The nonlinear-sigma-model approach 349 which the lattice is either fixed or altogether absent as in c. Other approaches 350 models of the "jellium" type. It is this second category C. Experiments 350 1. Bulk systems 350 which forms the subject of the present article. Histori- 2. Thin films 351 cally, the second category has again been divided into VIII. Disorder-Induced Unconventional Superconductivity 352 two classes, one in which the transition is triggered by . A. Amechanism for even-parity spin-triplet supercon- electronic correlations and one in which it is triggered by ductivi.ty 352 disorder. The first case is known as a Mott or Mott- B. Mean-field theory 354 Hubbard transition, the second as an Anderson transi- 1. The gap equation 3S4 tion. 2. The charge-fluctuation mechanism 355 3. The spin-fluctuation mechanism 356 Mott's original idea (Mott, 1949, 1990) of the C. Isthis mechanism realized in nature? 3S7 correlation-induced transition was intended to explain Other Theoretical Approaches and Related Topics 357 why certain materials with one electron per unit cell, e.g., A. Results in one dimension 357 NiO, are insulators. Mott imagined a crystalline array of B. Local magnetic-moment effects 360 atomic potentials with one electron per atom and a 1. Local moments in the insulator phase 360 Coulomb interaction between the electrons. For 2. The disordered Kondo problem 361 3. Formation oflocal moments 363 sufficiently small 1attice spacing, or high electron density, 4. Two-fluid model 365 the ion cores will be screened, and the system will be me- C. Delocalization transition in the quantum Hall effect 366 tallic. Mott argued that for lattice spacing larger than a D. Superconductor-insulator transition in two- critical value the screening will break down, and the sys- dimensional films 368 tem will undergo aerst order transiti-on to an insulator. X. Conclusions 369 This argument depended on the long-range nature ofthe A. Summary ofresults 369 Coulomb interaction. A related, albeit continuous, B. Open problems 370 metal-insulator transition is believed to occur in a tight- 1. Problems concerning local moments 371 2. Description ofknown or expected phase transi- binding model with a short-ranged electron-electron in- tions 371 teraction known as the Hubbard model (Anderson, a. Phase diagram for the generic universality 1959a; Gutzwiller, 1963; Hubbard, 1963). The model class 371 Hamiltonian is b. Critical-exponent problems 372 c. The superconductor-metal transition 372 (l.la) d. The quantum Hall effect 372 3. Problems with the nonlinear sigma model 372 a. Nonperturbative aspects and completeness 373 where a, and 8; are creation and annihilation opera- b. Symmetry considerations and renormalizabil- tors, respectively, for electrons with spin cr at site i, the ity 373 summation in the tight-binding term is over nearest 4. Alternatives to the nonlinear sigma model 374 neighbors, and 5. The 2-dground-state problem 374 Acknowledgments 374 (1.1b) References 374 Here t is the hopping matrix element, and Uis an on-site repulsion energy (U)0). Despite its simplicity, remark- I~ INTRODUCTION ably little is—known about the Hubbard model. In one di- mension (1 d) it has been solved exactly by Lieb and In the field of continuous phase transitions, metal- Wu (1968), who showed that at half filling the ground insulator transitions play a special role. In the first place, state is an antiferromagnetic insulator for any U)0. In they are not nearly so well understood, either experimen- higher dimensions, various approximations have suggest- tally or theoretically, as the classic examples ofliquid-gas ed that the model with one electron per site shows a con- critical point, Curie point, A, point in He, etc. In the tinuous metal-insulator transition at zero temperature as second place, one important subclass of metal-insulator a function of U (see Mott, 1990). This is generally be- transition consists ofwhat are now called quantum phase lieved to be true, but has not been firmly established. Re- transitions, i.e., continuous phase transitions that occur cently the Hubbard model has received renewed atten- Rev. Mod. Phys.,Vol.66,No.2,April 1994 D. Belitz and T.R. Kirkpatrick: The Anderson-Mott transition 263 tion, mostly because models closely related to it have 1.0 been suggested to contain the relevant physics for an un- 00 derstanding of high- T, superconductivity (Anderson, 0.8— o 1987). This renewed interest has spawned new approxi- 6— 0 0 mation schemes as well as new rigorous results. For in- A stance, it has been proven (among other things) that the 0.4— Oo ground state ofthe Hamiltonian, Eqs. (1.1),at half filling 0 has spin zero in all dimensions (Lieb, 1989). The Hub- 0.2— ~~ bard model has also been studied in the limit of infinite 0 0 dimension (Metzner and Vollhardt, 1989). It turns out 0 0.I2 0.I4 0.I6 0I8 o that in this limit the model can be reduced to a one- n/n dimensional problem that has been solved numerically. A number of nontrivial results have been =obtained. In FIG. 1. Numerical simulation data for the diffusion coefficient particular, it has been established that in d ao with in- D vs the scatterer density n ofa 2-d Lorentz model: , Bruin creasing U there is a Mott transition from a metallic to {1972,1974, 1978);O, Alder and Alley (1978),and Alley (1979), an insulating state (Georges and Krauth, 1992; Jarrel, as quoted by Gotze et al. {1982). D is normalized by its 1992;Rozenberg et al., 1992). Boltzmann value D' ', and n by its critical value n,. After Much more is known about the Anderson transition Gotze et al. (1982). (Anderson, 1958;for a recent review see Lee and Ramak- rishnan, 1985),which can be described by the Anderson model, whole band, then energies E, separate localized states in P=t y a,+e,+y E,a,+a, . (1.2) the band tails from extended ones in the band center, and (ij) i the metal-insulator transition can be triggered by sweep- ing the Fermi energy across the "mobility edge" E, This is a model for noninteracting electrons, so spin pro- (Mott, 1966, 1990);see Fig. 2. vides only trivial factors oftwo and can be omitted. The A further important development occurred when c; are randomly distributed site energies, governed by some distribution function characterized by a width 8'. Wegner (1976a) used real-space renormalization-group Anderson argued that for 8'/t large but finite the system methods to argue that the dynamical conductivity could d= be written in the scaling form is an insulator. In 1 it actually is insulating for all W)0 (Mott and Twose, 1961;Borland, 1963). For large — — 8', or near the band tails, the states were proven to be lo- ~) b (d 2)f(tb 1/v Qbd) (1.3) calized (Frohlich et a/., 1985). In d =3, Anderson pre- vcdaaiclllutyeed(oSafchmWoen/thra,al-mianmspuerlreadtoicratniodntraBnwrsehintiiicoghn, w1ta9os7o3cc)o.cnufTirrhmiasetdaasnnwounmezlelerrioa-s ptHhoeeirnetfretq(eui.segn.sc,oymt,=ebfdWiism—eannWsioa,nrfbl/eitsWrsar,yodirsstcatan=lecefpEafrraomEme,tfet/rhE,e, fc),riitIsiIcaainls- all other early (pre-1979) work, both numerical and unknown scaling function, and v an unknown correlation analytical, flaulcstouatfioounns.d a metal-insulator transition in length exponent. Equation (1.3) predicts that the static d =2, which was later shown to be incorrect (see below). conductivity vanishes at the metal-insulator transition S' — The reason for the insulating behavior at large is with an exponent s=v(d 2), and that the dynamical that the electrons become "trapped" or "localized" in the potential This is not an intrinsically quantum-mechanical phenomenon. Anderson's work had been motivated in part by work on classical percola- tion (Broadbent and Hammersley, 1957),and the classical Lorentz model (see, for example, Hauge, 1974) shows a transition between a diffusive and a localized phase much like the Anderson model (see Fig. 1). In the Lorentz model, a classical pointlike particle moves in a random array of fixed scatterers, often taken to be hard disks (d =2) or hard spheres (d=3). The localization does not require attractive potentials, but rather comes about by trapping of the particle in cages at suf5ciently high scatterer density. Ifthe scattering potentials are soft, the E (1) E (2) delocalization can be achieved not only by decreasing the C C density ofscatterers, but also by increasing the energy of FIG.2. Schematic picture ofthe"density ofstates Nvs the ener- the scattered particl8e.' An analogous effect exists in the gy Ein the Anderson model. E, ' are mobility edges, and the quantum case: if is not so large as to localize the states in the shaded regions are localized. Rev.Mod. Phys.,Vol.66,No.2,April 1994 264 D. Belitz and T.R.Kirkpatrick: The Anderson-Mott transition conductivity at the critical point, t=O, goes' like effect. Rather, one must try to separate the two effects by O'" ' ". The importance of this result lay in the means of the above-mentioned thermodynamic and demonstration that the metal-insulator transition could density-of-states anomalies, which accompany the in- be discussed in the canonical terms of critical teraction effect but not weak localization, or by means of phenomenon theory (see, for example, Ma, 1976;Fisher, the magnetoresistance. The magnetoresistance is nega- 1983). This line ofapproach to the model was taken one tive for the weak-localization model, since a magnetic step further with Wegner's (1979)mapping ofthe Ander- field destroys the phase coherence that is essential for son localization problem onto an efFective field theory. producing the interference effect (Altshuler, Khmel- The methods employed by Wegner and his followers are nitskii, Larkin, and Lee, 1980),while it is zero or positive the central theme ofthis review and will be explained in for various interaction models (Fukuyama, 1980; detail in Sec.III. Altshuler et a/., 1981). Abrahams et a/. (1979) conjectured that in d=2 all These perturbative considerations in the weak-disorder states are localized by arbitrarily weak disorder, as they regime raised the question of whether and how interac- = are in d 1. In contrast to Anderson localization per se, tion anomalies affect the metal-insulator transition. this is a pure quantum effect, which is not present in the Since the Coulomb interaction between the electrons is, Lorentz model and which can be understood as an in- of course, always present, a pure Anderson transition terference phenomenon (Bergmann, 1983, 1984). In con- cannot be expected to be realized in nature unless the in- trast to the 1-d case, the effect for weak disorder in d =2 teraction turns out to be irrelevant for the nature of the is only logarithmic for temperatures that are not too low. transition. As we shall see later, this in general is not the A key prediction was that in thin metallic films the resis- case. Unfortunately, this means that the comparatively tance at moderately low temperatures should 1ogarith- simple models developed for disordered noninteracting mically increase with decreasing temperature, a predic- electrons are insufhcient for understanding experiments. tion that has become known as the "weak-localization" On the other hand, most materials that display a metal- eff'ect. The observation of such logarithmic rises (Dolan insulator transition are highly disordered, and the pure and Osheroff, 1979;see also Bergmann, 1984)was widely Mott transition picture is equally inadequate. Consider, hailed as a confirmation of the theory. Later it became for instance, the case ofphosphorus-doped silicon, a par- clear that at least in some cases the agreement had been ticularly well studied example of a system showing a fortuitous, as the importance ofinteraction effects, which metal-insulator transition. Suppose one starts with pure are neglected in the weak-localization model, was not ap- silicon, which is an insulator at T=O. Upon doping with preciated early on. For many subsequent years the field the donor phosphorus, extra electrons are brought into enjoyed considerable activity, which has been reviewed the system. However, for low dopant concentrations the by Bergmann (1984) and by Lee and Ramakrishnan overlap between donor states is exponentially small, and (1985). one expects an insulator with hydrogenlike impurity Shortly after the prediction of the weak-loca1ization states. This is indeed what is observed (see Fig. 3). With effect it was realized that very similar effects can be increasing phosphorus concentration one finds a caused by a completely different physical mechanism. broadening ofthe lines due to impurity pairs, and finally Altshuler and Aronov (1979a, 1979b) showed that the a broad continuum due to a distribution ofimpurity clus- electron-electron interaction in 3-d weakly disordered ter sizes, with the system still being an insulator (Fig. 3). systems leads to a square-root cusp in the tunneling den- With a further increase in phosphorus concentration, one sity ofstates and to corresponding square-root anomalies in the temperature and frequency dependence of the E(meV) specific-heat coeKcient and the conductivity. The latter 0 10 20 30 40 50 60 anomaly has the same functional form as the 3-d weak- - localization contribution (Cxorkov et a/., 1979). In 2-d nD= 1.9xI10"cmI -' 1 —30 20 the corresponding effects are 1ogarithmic and, in particu- lar, the conductivity was predicted to have a logarithmic —10 temperature dependence just like the weak-localization 6 30 -.-„ effect, even ifthe interference effects that cause the latter 20— are neglected (Altshuler, Aronov, and Lee, 1980; 10— A Fukuyama, 1980). It thus became clear that an observed 30 anomaly in the conductivity by itself could not be taken 20 as evidence for the presence of the weak-localization 4.7x10I5 io 0 100 200 300 400 500 E(cm ) =t, The former behavior can be seen by choosing b the FIG.3. Far-infrared absorption coefBcient a for three different latter by choosing b=Q ' ". Scaling properties at the metal- donor concentrations nz in Si:P. The critical concentration in insulator transition will be covered in detail in Sec.IV. this system is n,-=3.7X10' cm . From Thomas et al. {1981). Rev.Mod. Phys.,Vol.66,No.2,April 1994 D. Belitz and T.R.Kirkpatrick: The Anderson-Mott transition 265 expects, according to Mott's argument, a transition to a (Castellani, Di Castro, Lee, and Ma, 1984) and by inter- metal at some critical concentration n,. Again, this is pretations in terms ofFermi-liquid theory (Altshuler and what is observed (see Fig. 4). The observed transition is Aronov, 1983;Castellani and Di Castro, 1986). continuous and cannot be understood purely in terms of In contrast to these successes, an understanding ofthe a Mott transition. The reason lies in the fact that the metal-insulator transition in the absence of either mag- doping process not only introduces excess electrons, but netic impurities or magnetic fields has proven much at the same time creates disorder, since the dopant atoms harder. The difFiculties are twofold: the interaction am- are randomly distributed in the host lattice. One there- plitude in the particle-hole spin-triplet channel scales to fore would expect the transition, in part, to follow the infinity if it is not cut off by magnetic effects, and the Anderson model. particle-particle or Cooper interaction channel has a The conclusion that has been reached over the years is structure that is not easily amenable to standard that neither Anderson's nor Mott's picture by itself is renormalization-group techniques. The first problem has sufficient to understand the observed metal-insulator received much attention (Finkel'stein, 1983a, 1984b, transition. Rather, one has to deal simultaneously with 1984c; Castellani, Di Castro, Lee, Ma, Sorella, and Ta- disorder and interactions between the electrons, neither bet, 1984, 1986; Castellani, Kotliar, and Lee, 1987)and one of which is a small effect near the transition. This was originally interpreted as being related to local mo- has proven to be a very hard problem, which is far from ment formation, or as signaling an exotic metal-insulator having been solved completely. Somewhat ironically, the transition in which the scaled disorder Aows to zero at most precise experiment, viz., the one on Si:P shown in the transition. More recent work has suggested that it Fig. 4, has proven the hardest to understand for reasons actually signals the presence ofa phase transition that is that will be discussed in detail in Sec.VI. Nevertheless, magnetic in nature and distinct from the metal-insulator substantial progress has been achieved "in our understand- transition (Kirkpatrick and Belitz, 1990b, 1992b; Belitz ing ofthis "Anderson-Mott transition. and Kirkpatrick, 1991). The second problem has been An important development in this respect was the considered (Castellani, Di Castro, Forgacs, and Sorella, work ofFinkel'stein (1983a, 1984a, 1984b),who extended 1984; Finkel stein, 1984b; Kirkpatrick and Belitz, 1993), the field-theoretic description ofthe Anderson transition but the proposed solutions so far are not mutually con- (Wegner, 1979;Efetov et al., 1980) to allow for interac- sistent and cannot even tentatively be considered final. tions. This model not only allowed for the use of They will be discussed in Secs.V and VI. Another prob- renormalization-group methods to deal with strong dis- lem is that experiments on doped semiconductors order, but also was able to consider interactions of arbi- (Paalanen et al., 1986, 1988) show thermodynamic trary strength. It thus achieved two important improve- anomalies that cannot be consistently explained within ments over previous perturbative work and quickly led to the field-theoretic model and have prompted rather a description of the Anderson-Mott transition in the different theoretical approaches (Bhatt and Fisher, 1992). presence of magnetic impurities or a magnetic field These will be considered in Sec.IX. (Finkel stein, 1984a), which will be reviewed in Sec. V. For all these reasons the metal-insulator-transition These results were soon supplemented by a derivation in problem cannot be considered solved. However, the pro- terms of resummed many-body perturbation theory gress made within the last decade has not been reviewed, and it is the purpose ofthe present article to describe the 10 current state ofaffairs. In doing so, one difficulty is that ~ I the problem has been tackled by a large variety of ap- INSULATOR METAL proaches that are very different with respect to both the underlying physical ideas and the technical methods 0 used. On the physical side, one can distinguish between 102 6' phenomenological approaches, on the one hand, which ~ try to get clues from experiments about what physical effects are important near the transition and must be in- cluded in the theory, and what one might call the slow- 10— mode- philosophy on the other hand. The latter starts from the assumption that the physics near the metal- insulator transition will be dominated by the low-lying excitations ofthe system, which can be extracted from a 1 simplified. microscopic model. On the technical side, in- 0 tuitive phenomenology, many-body perturbation theory, n (10~8cm 3) the renormalization group, and effective field-theoretic FICz. 4. Divergence ofthe dielectric susceptibility y (~), and techniques have all played an important role. Since the vanishing ofthe static conductivity o (0),both extrapolated to problem remains unsolved, one cannot really afford the zero temperature, at the metal-insulator transition in Si:P n is luxury of taking one of these points of view exclusively. the Pconcentration. After Rosenbaum et a7.(1983). We shall focus, however, on the low-lying-mode philoso- Rev.Mod. Phys.,Vol.66,No.2,April 1994 D. Belitz and T.R. Kirkpatrick: The Anderson-Mott transition phy, implemented by field-theoretic techniques, for two The localization problem has been reviewed previously reasons. First, this line ofapproach is relatively new for a number of times, most notably by Altshuler and Aro- interacting systems and has recently led to developments nov (1984), Bergmann (1984), Lee and Ramakrishnan that have not been covered in previous reviews. Second- (1985),Finkel'stein (1990),and MacKinnon and Kramer ly, we believe that these techniques have the best chance (1993). We have tried to avoid duplication ofmaterial as ofeventually providing us with a complete, microscopic far as possible and often refer to these reviews rather theory ofthe metal-insulator transition. One should keep than trying to be complete. With some exceptions, we in mind, however, that even ifone accepts the slow-mode also concentrate on the metal-insulator transition proper approach, it is in general still an open problem how to and its immediate vicinity, excluding effects at weak dis- determine all ofthe relevant slow modes. We shall come order or deep in the insulator. This holds in particular back to this problem in Secs.II,III,and X. for our selection of experiments to be discussed in detail The plan ofthis review is as follows. We start in See. in Secs. V and VI. Throughout the paper we use units II with an elementary discussion of the slow modes, i.e., such that Planck's constant A. Boltzmann's constant kz, the diffusive modes that result from the conservation and minus the electron charge e, are equal to unity unless laws for particle number, spin, and energy. All of the otherwise mentioned. material presented in that section can be found in various books and review articles. We feel, however, that our II. DIFFUSIVE ELECTRONS discussion is necessary both for pedagogical reasons and to put the results ofthe field theory presented later in the proper context. Section III is devoted to an explanation A. Diffusion poles ofthe technical apparatus that will be used in most ofthe rest ofthe paper. That section is rather technical and ex- The dynamics of conserved quantities show peculiari- tensive for two reasons: the field-theoretic methods un- ties that arise from the fact that, due to the conservation derlying much of the work to be reviewed are not as law, their values cannot change arbitrarily in space and widely known among condensed-matter theorists as, say, time. Let us consider the density n(x,t) of a conserved Green's-function techniques, and the details of the quantity Xin some many-particle system. In equilibri- derivation ofthe fundamental model describing interact- um, n is constant in space and time: n(x,t)=no. Sup- ing disordered electrons have never been published. The pose a fluctuation 6n is created, n(x,t)=no+An(x, t), section is written for readers who wish to work actively and we ask how the system will go back to equilibrium. with the field theory. Anybody who is mainly interested Since n is conserved, it can do so only by transporting in learning about the results presented in the later sec- some Xout ofor into the region where 5nAO. Ifthis re- tions can skip over most of the technical details in Sec. gion is large, this will take a long time. Therefore long- III. Section IV is devoted to a general discussion ofpos- wavelength fluctuations of conserved quantities will de- sible scaling scenarios for a metal-insulator transition of cay very slowly. Slowly decaying fluctuations determine interacting electrons, i.e., the question ofhow to general- the low-lying modes and are of central importance for a ize Eq. (1.3) to the interacting case. Sections V and VI description of the system. An example is classical Quid review explicit calculations that show how these scaling dynamics, where the conserved quantities are particle scenarios are realized in various universality classes. Sec- number, momentum, and energy, and the slow modes are tion VII is devoted to a related subject, namely, the de- first sound, heat diffusion, and transverse momentum struction of (conventional bulk) superconductivity near diff'usion (see, for example, Forster, 1975;Boon and Yip, the metal-insulator transition. This problem is actually 1980). part of a more general one, namely, the question ofhow We shall be concerned with the dynamics of electrons collective phenomena like superconductivity, magnetism, moving in a random array of static scatterers. Since the etc., are affected by strong disorder in the vicinity of a scatterers can absorb momentum, the only conserved metal-insulator transition. Since the answer obviously re- quantities are particle number (or charge), energy, and quires a solution ofthe problem in the absence ofthe col- possibly spin. In general, all of these have diffusive dy- lective phenomenon, these issues have only recently start- namics. The central assumption of the theory we shall ed to be addressed. In Sec.VIII we discuss a recent sug- review is that the slow decay of charge, spin, and gestion ofdisorder-induced spin-triplet superconductivity energy-density Auctuations leads to, and dominates, the in 2-d systems. In its existing form the slow-mode field physics near the metal-insulator transition. The basic theory is unlikely to accomplish the ultimate goal ofpro- strategy for a description at zero temperature is to start viding a complete microscopic theory of all phenomena with perturbation theory in the diffusive phase and to observed close to the metal-insulator transition. Rather, study the instability of that phase. Of course, this it will have to be supplemented by physical ideas presumes the existence ofa diffusive phase somewhere in developed through other approaches. Some of these are the phase diagram. Ifthis is not true, one can use pertur- discussed in Sec.IX. Section Xprovides a summary and bation theory only at finite temperature, where transport a discussion ofwhat we consider to be the most pressing is diffusive due to inelastic processes. As the temperature open problems in the field. approaches zero, perturbation theory will then break Rev. Mod. Phys.,Vol.66,No.2,April 1994 D. Belitz and T.R. Kirkpatrick: The Anderson-Mott transition 267 down everywhere in the parameter space. An example I Id;, over the randomly situated scattering centers may be seen in the 2-d systems, for which all theoretical (Edwards, 1958). For simplicity we assume pointlike approaches now agree that at T=O electrons are in gen- scatterers (i.e., s-wave scattering only). p(q) is the densi- eral never diffusive (Abrahams et al., 1979). More re- ty operator, cently it has been suggested that th)e spin dynamics may = + never be diffusive, not even for d 2 (Bhatt and Fisher, P(q) X&k—q/2~k+q/2 (2.1c) k 1992). The possible consequences of this are currently not quite clear. However, in the simplest possible Since we are dealing with noninteracting electrons, spin scenario the absence of spin diffusion would merely results only in trivial factors of two and can be change the universality class (see Sec. III.B.4.c) of the suppressed. We shall add the Coulomb interaction later. metal-insulator transition. We shall discuss this proposi- tion further in Sec.)IX. Here we proceed under the as- a. A phenomenological argument fordiffusion sumption that for d 2 at T=O there is a small-disorder phase where charge, spin, and energy are diffusive. Let us start with a very simple phenomenological argu- In this subsection we consider noninteracting electrons ment for diffusive density dynamics (see, for example, in an environment ofelastic, spin-independent scatterers. Forster, 1975). Of course this approach does not depend In this case the conservation laws for spin and energy do . on microscopic details and also holds for interacting elec- not add anything to particle number conservation, and trons as well as for systems outside the quasiclassical re- the spin and heat diffusion coefficients are the same as the gime. Consider a macroscopic number-d. ensity Auctua- charge or number diffusion coefficient. For the spin tion 5n(x,t). Particle number conservation implies the diffusion coefficient this is obvious, and for the heat continuity equation diffusion it has been shown by Chester and Thellung (1961),Castellani, DiCastro, and Strinati (1987),and Stri- 8 5n(x, t)+V j(x,t)=0, (2.2) nati and Castellani (1987). We can therefore restrict our- at selves to a discussion of particle number diffusion. The with j the (macroscopic) number current density. It is situation changes, of course, as soon as the electron- plausible to assume that for a slowly varying density the electron interaction is taken into account; see Sec. current is proportional to the negative gradient of the III.B.3.d. density, — j( tx)= DV5n(x, t) . (2.3) 1. The quasiclassical approximation for electron transport The positive coefficient D is called the diffusion constant. The basic building block of the theory is the diffusive More precisely, jshould be expressed in terms ofa chem- density response ofthe electrons in the quasiclassical ap- ical potential gradient and an Onsager coefficient, which proximation. We first discuss three derivations of the in turn can be expressed in terms of a density gradient density response, in order ofincreasing technical sophis- and the diffusion coefficient (see, for example, DeGroot tication. For the time being, we consider noninteracting and Mazur, 1962). Combination of Eqs. (2.2) and (2.3) electrons with a Hamiltonian yields Fick's law, — 8'=g [k /2m —p]8k+&k+ 1 g u(q)p+(q) . (2.1a) Db, 5n(x,t)=0 . — (2.4) at Here ak and ak are creation and annihilation operators We Fourier transform and find the solution, for electrons in state k,p is the chemical potential, we as- — sume free electrons with mass m, and V is the system 5n(q, t)=5n(q, 0)e D~2t', t &0, (2.5a) volume. u is a random potential whose strength is given which display:s the slow decay of long-wavelength Auc- by tuations mentioned above. We de6ne a Laplace trans- (2.1b) form in time by 5n(q,z)—+i fdt 6(+r)e'"5n(q, t), with NF the density ofstates (DOS) per spin at the Fermi + level, and ~ the elastic mean free time in the Boltzmann for Imz m~0, (2.5b) approximation. n,. is the scatterer density, whose appear- with complex frequency z. This yields ance results from performing the ensemble average =0)— = 5n(q, t 5n(q,z) (2.5c) z+iDq The same istrue for number and heat difFusion in the classical Here 5n(q,z) as a function of z has a branch cut at Lorentz model mentioned in Sec.I. Quantum mechanics does Imz=0 and two Riemann sheets. The physical sheet is not change this. the one with no singularities. The analytic continuation Rev. Mod. Phys.,Vol.66,No.2,April 1994 268 D. Belitz and T.R. Kirkpatrick: The Anderson-Mott transition =—1 taoptohlee oatthzer=s—heieDtqfroamndazbo=vieDaqn,d rbeeslpoewctivtheely.realThaxisisphoales x"(q,n)= 2l [x (q,n+io) —x (q,n —iO)] is called a diffusion pole. Its relation to diffusive dynam- ==1—f ics is obvious from Eqs. (2.5). dt e' 'X (q,t) . (2.10b) The retarded and advanced susceptibilities are given by b. Linear response and the Boltzmann equation x '"(q,n)=x (q,n+io)=x' (q,n)+ix" (q,n), We now turn to a microscopic derivation ofEqs. (2.5). (2.11a) Again, the formal part ofthis subsection is valid for gen- eral systems. Suppose the deviation from equilib„rium,, where 5n, is created by an external chemical potential p,, (x,t). =1— — Then the Hamiltonian contains a term x' (q,Q)= [x (q,Q+iO)+x (q,n io)] . (2.11b) „, „, H, (t)=—fdxp(x)p, , (x,t), (2.6) is positive semidefinite and determines the energy dis- happ where p is the density operator, Eq. (2.1c). Linear- sipation in the system. It is therefore also called the "dis- response theory (Fetter and Walecka, 1971; Forster, sipative part" of the susceptibility. It is related to the 1975) then tells us that the change in the expectation "reactive part" happ by means of a Kramers-Kronig rela- value ofp, to linear order in p, is given by tion, 5n(x,t)—=&p(x,t)&—&p(x,t)&„„ „=0 dn' x',"n,(q nn') - (2.12a) „, =i f dt' fdx'X ( x'xt,t')p, ( 'xt'), (27a) dn' xI q» — (2.12b) n n with the density susceptibility (x,x',t,t')=&[p +(x,t),p(x', t')]& . (2.7b) Fouinsaldlye,nsitthye dAiusscitpuaattiivoens painrtthhaepp siyssrteemlatedbyttohetheAuscptounattaionne-- dissipation theorem (Callen and Welton, 1951). Corre- Here [a,b]=itb ba for an—y two operators a, b. The sponding relations hold for the general A —8susceptifbil- averaging is performed with the unperturbed Hamiltoni- ity, Eq. (2.9b), and generally for any causal function (z) an. Ifwe include the ensemble average in the definition instead ofx (q,z). ofthe brackets in Eqs. (2.7), the system is translationally The susceptibility determines the response of the sys- invariant in space and time, and we have tem to external perturbations. The exact form of the „, bn (q,t)=i f dt'X (q,t t')p,, (q,—t'), (2.8a) r(eKsupobnos,e 1s9ti5ll7)d.epSeunpdpsosoen„tthhee,npaeturtruerbaotifot„nhe,piesrtusrubdadteionlny switched on at t=O: p, (q,t)=6(t)p, (q). Then one with finds from Eq. (2.8a) (q,t)=&[p+(q,t),p(q, t=0)]& . (2.8b) z)=——1 „, 5n(q, X q(q,z)p, (q) . (2.13) A Laplace transformation according to Eq. (2.5b) yields z the causal density susceptibility, which is equal to minus Now suppose that the perturbation is turned on adiabati- Zubarev's (1960)commutator correlation function, cal„ly , at t=—oo „an,d switched off at t=0: Xqq(q, z)=+i fdt 0(+t)e'"X (q,t) p, (q,t)=6( t)e"p,, (q„) (e,~—0). Then one finds =—«p+(q);p(q) », , (2.9a) 5n(q,z)=C& (q,z)p, (q) . (2.14a) Here N is Kubo's relaxation function, with the notation « a;B»,=+ifdt e(+t)e"'&[A(t)B] (2.9b) +',,(q z)==1—Ix„(qz)—x',,(q)] (2.14b) & for any operators A, B. X (q,z) has the usual properties with the static susceptibility of causal functions (see, for example, Forster, 1975), which we list here without derivations. Causality allows (q)=X (q,z=iO)= dQ X"(q,n)/n . (2.14c) for a spectral representation, Kubo has noted that the static:susceptibility g in general dn x",,«») is different from the isothermal susceptibility, z)= (q, (2.10a) = xqq(q) &p(q) &— (q), (2.15) with a spectral function ~sexi q Bp Rev.Mod. Pbys.,Vol.66,No.2,April 1994 D. Belitz and T.R. Kirkpatrick: The Anderson-Mott transition 269 which enters the Kubo function (q,z)= 2NL (q,z), (2.19a) (q,z)=—z1 y (q,z)—tr))pn (q) (2.16a) with the longitudinZal current Kubo function i At q=0, the isothermal den—sity susceptibility is related C (q, )=——(1/q') «q j+(q) q.j(q)»—I to the compressibility rr= (t)V/i)p)T/V by (Forster, Z 1975) (2.19b) Bn/Bp=n ir, (2.16b) Let us consider Eqs. (2.19) in the limit q~O. Kubo so (Bn/Bp)(q) can be interpreted as the wave-vector- (1957) has shown that Nl (q=O,z) determines the dependent compressibility. Kubo (1957)has shown that dynamical conductivity, fsoharllergeondciocunvtearriabnloesnonotnreivihaals ynon=egrgo.dicIn tvhaisriarbelveise,w anwde o.(z)= i@—1(q=O,z) . we shall not distinguish between y and y . The distinc- For small frequencies, cr is further related to the diffusion tion is crucial, however, for a description of, e.g., the in- constant by an Einstein relation (Kubo, 1957), sulating side of the metal-insulator transition (Gotze, 1981),and it may be important for a solution ofsome of lim cr(Q+iO)=+ Bn D . (2.21) the open problems discussed in Sec. X. Let us also Q 0 Bp mention that for free electrons one has y (q) It follows from Eq. (2.19a)that in the hydrodynamic lim- =(Bn/t)p)(q) =g(q), with g the static Lindhard func~tion it y vanishes like q as a result ofparticle number con- (Lindhard, 1954; Pines and Nozieres, 1989). In the servation, homogeneous limit, g(q =0)=K+. That is, for free elec- trons the compressibility and the single-particle DOS are lim y (q,z)= q i.D Bn sgn(Imz)+o(q 2) . the same. This is not so for interacting electrons, as can q~0 z Bp be seen already at the level ofFermi-liquid theory, wheIre He~re0o.(q ) denotes terms that vanish faster than q as the compressibility contains the Landau parameter 0, q If we compare Eq. (2.22a) with the zero- while XF does not [see, for example, Pines and Nozieres, frequency result 1989, and Eqs. (3.126) and (3.127) below]. In the pres- ence ofdisorder, the distinction between the two quanti- lim y (q,Q+iO) = Bn (q), (2.22b) Q~O Bp ties is crucial (Lee, 1982). In order to determine the general form of the density we see that the limits q—+0 and Q~O do not commute, response, we use the equations ofmotion for the suscepti- so g (q,z) must be nonanalytic at q =0, z=0. Indeed, bility (Zubarev, 1960), we can use Eq. (2.22a) to write the Kubo function in the z« 3;B»,=&[A,k ]&—« [8,A ];B», hydrodynamic limit as — =&[J,S]&+«J;[H,k]», . (2.17) (q~0,z)= t)n /t)p (2.23) z+iDq sgn(Imz) For the density operator, Eq. (2.1c),we have We thus recover the diffusion pole of Eq. (2.5c). Note [~p(q) l=q j(q» (2.18a) that the Kubo function is the appropriate response func- with the current-density operator tion to compare with Eq. (2.5c),since in our phenomeno- logical argument we had assumed an adiabatically Pl &1 ~k+—/2~k+ /2 (2.18b) plarxepaacrecdordinnogneqtouiltihberiuumnpertsutrabteed, whsyicshtemw'sasdaylnloawmeicds.toFroer- later reference we note that Eqs. (2.21)—(2.23) are gen- Equations (2.18) are the microscopic analog of Eq. (2.2). erally valid, not just for noninteracting electrons. Notice that they are a consequence of particle number The question remains how to calculate o. or D. In the conservation and remain valid for interacting elec- quasiclassical approximation, we can use the Boltzmann trons. Applying Eqs. (2.17) twice, and using equation with the result (see, for example, Ziman, 1964) [j+(q),p(q)]&=qn/m, we find & n/I (2.24a) Bn/Bp A nonergodic variable in this sense is one whose correlations where r is the elastic mean free time. In three dimen- do not decay in the limit of long times, so its Kubo functtiohn at—sions, diverges for small z like 4{q,z~O)= f(q)/z. Notice — this is always the case for conserved quantities at q =0,but is a =n,2rrvz m.dosing(1 cos8)cr(8—), (2.24b) nontrivial property at q&0. For a mathematical discussion of 7 ergodicity, see Khinchin (1949). where o(t)) denotes the differential scattering cross sec- Rev.Mod. Phys.,Vol.66,No.2,April 1994 270 D. Belitz and T.R. Kirkpatrick: The Anderson-Mott transition tion. For isotropic scattering the cosB term does not and its Fourier transform cBoonrtnribauptper,oxiamndatioifnwewtereraetcotvheer scEaqtt.er(i2n.glb).proWceesscaninatlhsoe vr (q,iQ„)= Pdre iQ"„7~. (q,r) . (2.25b) 0 solve the Boltzmann equation directly for the density Here ~ denotes imaginary time, T, is the imaginary-time response. In general one obtains Eq. (2.23) for the densi- ty response function. For the special case of isotropic ordering operator, Q„=2~Tn,with n an integer, is a bo- scattering the explicit result for the diffusion coefficient is sonic Matsubara frequency, and @=1/Tis the inverse again given by Eqs. (2.24) (see, for example, Hauge, temperature. ~ (q,iQ„),which is often called the polar- 1974). ization function, is identical to minus the causal density =iQ„. susceptibility, Eq. (2.10a),taken at z The retarded and advanced susceptibilities, Eq. (2.1la), can be ob- c. Oiagrammatic deri vation ofthe diffusive density response tained by analytical continuation to real frequencies, We shall now calculate the density and current correla- ~ '"(q,Q)=m.(q,iQ„~Q+i0) . (2.26) tion functions explicitly by means of many-body pertur- bation theory. In order to do so, we have to rewrite the commutator correlation function, Eq. (2.8b), in terms ofa An analogous polarization function can be formed with the current operator. The Kubo formula for the conduc- time-ordered correlation function. This can be done us- tivity, Eq. (2.20),then takes the form ing standard techniques (Fetter and Walecka, 1971; Mahan, 1981)and allows for a convenient handling ofthe correlation function formalism at finite temperatures. o.(Q)=i .1 ~L(q=0,i.Q„)+n iQ„—+0+i0 We define an imaginary-time correlation function in the Matsubara formalism, (2.27a) vr (q,r)=—(T,p+(q, r)p(q, r=O)), (2.25a) where ml(q, iQ„)=—f dre " (T,(q/q) j+(q,r)(q/q) j(q,&=0)) . (2.27b) 0 I The Wick theorem can now be used to evaluate the which is shown graphically in Fig. 5(b). With the explicit time-ordered representations ofthe correlation functions expression for G' ', in Fpeorrturobuartiopnresetnhteorpyu.rposes, the small parameter for a GI '(p,icy„)=[ice„p—/2m+@] (2.28') perturbative treatment is the density of scatterers n, the integrals are easily done, and one obtains the familiar The averaging over the random positions of the scatter- Lindhard function (Lindhard, 1954). The same result can ers can be performed using the technique developed by be obtained by evaluating the commutator in Eq. (2.7b) Edwards (1958;see also Abrikosov et al., 1975; Mahan, and performing the Fourier-Laplace transform. For 1981). The building blocks of the theory are, first, the finite impurity concentrations, we know from the previ- bare-electron Green's function, (a) G' '(q,ice„)=—f dr e " (T,a (r)a+ (r=0))H G&p& Up (b) and, second, the impurity factor uo, Eq. (2.1b). co„=2rrT(n+I/2), with n an integer, is a fermionic Matsubara frequency, and the index 00 indicates that the average is to be taken with the free-electron part of G(p) 6' the Hamiltonian only. Diagrammatically, we denote ' by a directed light line, and uo by two broken lines [one for each factor ofu(q)] and a cross (for the factor ofn,;); see Fig. 5(a). To zeroth order in the impurity density, the density polarization function is given by FIG.5. Diagrammatic elements ofperturbation theory: (a) di- agrammatic representation ofthe bare Careen's function and the m' '(q,iQ„)=gTg G' '(p,ice„) impurity factor; (b) diagrammatic representation of the bare P ECO density polarization function; (c) the Cxreen's function in the XG' '(p+q,iso„+iQ„), (2.29) Bsitoyrnpoalpaprirzoaxtiimonatiofnu;ncti(odn).conserving approximation for the den- Rev. Mod. Phys.,Vol.66,No.2,April 1994

Description:
metal-insulator transition is believed to occur in a tight- by evaluating the commutator in Eq. (2.7b) and performing the Fourier-Laplace transform.
See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.