<<

View metadata, citation and similar papers at core.ac.uk brought to you by CORE

provided by University of Birmingham Research Portal

University of Birmingham

Non-Fermi liquids Schofield, Andrew

DOI: 10.1080/001075199181602

Document Version Peer reviewed version Citation for published version (Harvard): Schofield, A 1999, 'Non-Fermi liquids', Contemporary Physics, vol. 40, no. 2, pp. 95-115. https://doi.org/10.1080/001075199181602

Link to publication on Research at Birmingham portal

General rights Unless a licence is specified above, all rights (including copyright and moral rights) in this document are retained by the authors and/or the copyright holders. The express permission of the copyright holder must be obtained for any use of this material other than for purposes permitted by law.

•Users may freely distribute the URL that is used to identify this publication. •Users may download and/or print one copy of the publication from the University of Birmingham research portal for the purpose of private study or non-commercial research. •User may use extracts from the document in line with the concept of ‘fair dealing’ under the Copyright, Designs and Patents Act 1988 (?) •Users may not further distribute the material nor use it for the purposes of commercial gain.

Where a licence is displayed above, please note the terms and conditions of the licence govern your use of this document.

When citing, please reference the published version. Take down policy While the University of Birmingham exercises care and attention in making items available there are rare occasions when an item has been uploaded in error or has been deemed to be commercially or otherwise sensitive.

If you believe that this is the case for this document, please contact [email protected] providing details and we will remove access to the work immediately and investigate.

Download date: 01. Mar. 2020 Non-Fermi liquids

A. J. Schofield

The University of Cambridge, Department of Physics, The Theory of Condensed Matter Group, The Cavendish Laboratory, Madingley Road, Cambridge. CB3 0HE November 11, 1998

Our present understanding of how the interactions between affect the metallic state has, for forty years, rested on the foundations of Landau’s Fermi-liquid theory. It provides the basis for understanding in terms of weakly interacting (-like) particles. Recent years have seen the discovery of metals which appear to fall outside this framework—perhaps most notably in the normal state of the high temperature cuprate superconductors. While the theory for understanding the cuprate metals remains controversial, there are a number of clear examples where we do believe we understand the new underlying theoretical concepts. In this article I illustrate four such routes towards forming a non-Fermi liquid and illustrate, where possible, how these have been realized in a number of materials. The proximity to a quantum phase transition and reduced effective dimensionality can both play important roles.

1. Introduction over the past decade there is a growing field of con- densed matter physics devoted to understanding metals Condensed matter physics is a subject continually in- where the electron seems to be precisely the wrong place spired by the fabrication of new materials. With each to start. Of course we are well aware that the basic in- new generation of materials synthesized comes a new gredients of solids are atoms with their valence and core set of challenges for the condensed matter physics com- electrons and nuclei but, as often happens in condensed munity to understand and exploit. To the observer this matter, in bringing such atoms together what emerges may seem surprising since the basic interactions gov- can be very different from the constituent parts, with erning the motion of the electrons and atomic nuclei in the possibility of completely new types of ‘particles’. a solid have long been known. While this is true, with (Two more familiar examples of new particles appearing each new compound we see these basic forces at work in condensed matter systems are phonons–quantized in a different local environment and the result is rarely lattice vibrations–and magnons–waves of spin in a mag- a trivial extrapolation of the physics we knew before. net.) Yet for the understanding of the metallic state the Instead, with each level of complexity we see new types electron has remained the unrivaled basis since Drude’s of phenomena arising - every bit as fundamental as the initial work at the beginning of this century (Drude bare interactions with which we began (see Anderson 1900). The success of the single electron picture of met- 1972). Examples of such radically new behaviour in- als rests on Landau’s seminal work in the 1950’s devel- clude the appearance of fractionally charged objects in oping Fermi-liquid theory (Landau 1956, 1957, 1958). the fractional quantum Hall effect, the observation of Yet as this century closes we are seeing the discovery of super-massive electrons in so-called heavy ma- materials, including the cuprate superconductors and terials and the possibility of the electron decaying into other oxides, which seem to lie outside this framework. new types of particle in certain one-dimensional mate- In this review I will attempt to give a flavour of our rials. One of the current areas of excitement in the field attempts to understand such ‘non-Fermi liquid’ metals. has been motivated by the discovery of certain metallic Many of the theoretical ideas have been developed using compounds which seem to fall outside of the framework rather complex mathematical machinery which I make of our current theory of metals. no apology for omitting in favour of a more descriptive It is hard to imagine describing the physics of metals approach. There are however a number of results which without beginning with the electron yet, remarkably,

1 Non-Fermi liquids A. J. Schofield 2 can be obtained relatively simply using Fermi’s golden rule (together with Maxwell’s equations) and I have in- cluded these for readers who would like to see where some of the properties are coming from. The outline of this review is as follows. I begin with a description of Fermi-liquid theory itself. This the- ory tells us why one gets a very good description of a metal by treating it as a gas of Fermi particles (i.e. that obey Pauli’s exclusion principle) where the interactions are weak and relatively unimportant. The reason is that the particles one is really describing are not the original electrons but electron-like that emerge from the interacting gas of electrons. Despite its Figure 1: The ground state of the free Fermi gas in mo- recent failures which motivate the subject of non-Fermi mentum space. All the states below the liquids, it is a remarkably successful theory at describ- are filled with both a spin-up and a spin-down elec- tron. A particle-hole excitation is made by promoting ing many metals including some, like UPt3,wherethe interactions between the original electrons are very im- an electron from a state below the Fermi surface to an portant. However, it is seen to fail in other materials empty one above it. and these are not just exceptions to a general rule but are some of the most interesting materials known. As an example I discuss its failure in the metallic state of 2. Fermi-Liquid Theory: the electron quasi- the high temperature superconductors. particle I then present four examples which, from a theo- The need for a Fermi-liquid theory dates from the retical perspective, generate non-Fermi liquid metals. first applications of quantum mechanics to the metallic These all show physical properties which can not be state. There were two key problems. Classically each understood in terms of weakly interacting electron-like electron should contribute 3kB/2 to the specific heat objects: capacity of a metal—far more than is actually seen ex- perimentally. In addition, as soon as it was realized • Metals close to a quantum critical point. When a that the electron had a magnetic moment, there was phase transition happens at temperatures close to the puzzle of the magnetic susceptibility which did not absolute zero, the quasiparticles scatter so strongly show the expected Curie temperature dependence for that they cease to behave in the way that Fermi- free moments: χ ∼ 1/T . liquid theory would predict. These puzzles were unraveled at a stroke when • Metals in one dimension–the Luttinger liquid. In Pauli (Pauli 1927, Sommerfeld 1928) (apparently one dimensional metals, electrons are unstable and reluctantly—see Hermann et al. 1979) adopted Fermi decay into two separate particles (spinons and statistics for the electron and in particular enforced the holons) that carry the electron’s spin and charge exclusion principle which now carries his name: No two respectively. electrons can occupy the same quantum state. In the absence of interactions one finds the lowest energy state • Two-channel Kondo models. When two indepen- of a gas of free electrons by minimizing the kinetic en- dent electrons can scatter from a magnetic impu- ergy subject to Pauli’s constraint. The resulting ground rity it leaves behind “half an electron”. state consists of a filled Fermi sea of occupied states in momentum space with a sharp demarcation at the • Disordered Kondo models. Here the scattering F and momentum pF =¯hkF (the Fermi from disordered magnetic impurities is too strong surface) between these states and the higher energy un- to allow the Fermi quasiparticles to form. occupied states above. The low energy excited states While some of these ideas have been used to try and un- are obtained simply by promoting electrons from just derstand the high temperature superconductors, I will below the Fermi surface to just above it (see Fig. 1). show that in many cases one can see the physics illus- They are uniquely labelled by the momentum and spin trated by these examples in other materials. I believe quantum numbers of the now empty state below the that we are just seeing the tip of an iceberg of new types Fermi energy (a hole) and the newly filled state above of metal which will require a rather different starting it. These are known as particle-hole excitations. point from the simple electron picture to understand This resolves these early puzzles since only a small their physical properties. fraction of the total number of electrons can take part

Non-Fermi liquids A. J. Schofield 3



1

2

2

x jxj < ;

d 

in the processes contributing to the specific heat and 1

2

+ V x = E V x =

2 dx

magnetic susceptibility. The majority lie so far below 2

1 jxj  : the Fermi surface that they are energetically unable to Energy find the unoccupied quantum state required to mag- N=4 netize them or carry excess heat. Only the electrons within kB T of the Fermi surface can contribute kB to the specific heat so the specific heat grows linearly with N=3 temperature and is small. Only electrons within µBB of N=4 the Fermi surface can magnetize with a moment ∼ µB leading to a temperature independent (Pauli) suscepti- N=2 bility. Both quantities are proportional to the density N=3 of electron states at the Fermi surface. N=1 These new temperature dependencies exactly N=2 matchedtheexperimentsbothonmetalsandthen N=1 N=0 3 later on the fermionic isotope of Helium - He (see, for N=0 example, Wheatley 1970). But this in turn raised ques- tions. Why should a theory based on a non-interacting 0 λ 1 picture work so well in these systems where interactions are undoubtably important? Once interactions are Figure 2: Adiabatic continuity is illustrated in a non- present the problem of finding the low energy states interacting problem by turning on a quadratic potential of the electrons becomes much harder. In addition to to a particle confined in box. While the energy levels the kinetic term which favours a low momentum, the and the details of the eigenstate wavefunctions evolve energy now contains a potential term which depends on subtly , the good quantum numbers of the initial prob- the relative position of all of the electrons. The energy lem (the number of nodes, N, in the wavefunction) are scales of the kinetic energy and Coulomb interaction still the appropriate description when the perturbation are comparable at metallic electron densities and, has been applied. if that were not enough, Heisenberg’s uncertainty principle prevents the simultaneously definition of the states of the interacting electrons with the those of the momentum and the position. How can one proceed and non-interacting Fermi gas. He supposed that the good still hope to retain the physics of the non-interacting quantum numbers associated with the excitations of the electron gas which experiment demands? non-interacting system would remain good even after The answer provided by Landau rests on the concept the interactions were fully applied. Just as Pauli’s ex- of “adiabatic continuity” (Anderson 1981): labels as- clusion principle determined the allowed labels with- sociated with eigenstates are more robust against per- out the interactions through the presence of a Fermi turbations than the eigenstates themselves. Consider surface, this feature would remain even with the in- as an example the problem of a particle in a box with teractions. We therefore retain the picture of Fermi impenetrable walls illustrated in Fig. 2. In elementary particles and holes excitations carrying the same quan- quantum mechanics one learns that the eigenstates of tum numbers as their electron counter-parts in the free this problem consist of standing sine waves with nodes Fermi gas. These labels are not to be associated with at the well walls. The eigenstates of the system can be electrons but to ‘quasiparticles’ to remind us that the labelled by the number of additional nodes in the wave- wavefunctions and energies are different from the cor- function with the energy increasing with the number of responding electron in the non-interacting problem. It nodes. Now switch on an additional weak quadratic is the concept of the fermion that lies at potential. The new eigenstates of the problem are no the heart of Fermi-liquid theory. It accounts of the mea- longer simple sine waves but involve a mixing of all the sured temperature dependences of the specific heat and eigenstates of the original unperturbed problem. How- Pauli susceptibility since these properties only require ever the number of nodes still remains a good way of the presence of a well defined Fermi surface, and are labelling the eigenstates of the more complicated prob- not sensitive to whether it is electrons or quasiparticles lem. This is the essence of adiabatic continuity. that form it. Landau applied this idea to the interacting gas of Retaining the labels of the non-interacting state electrons. He imagined turning on the interactions be- means that the configurational entropy is unchanged tween electrons slowly, and observing how the eigen- in the interacting metal. [This also means that quasi- states of the system evolved. He postulated that there particle distribution function is unchanged from the would be a one-to-one mapping of the low energy eigen- free particle result (see Fig. 3a).] Each quasiparticle Non-Fermi liquids A. J. Schofield 4

Probability Probability loses ω k,ε ' k+q,ε '' ε gains ω 1 1 quasiparticle q,ω z p,ε p-q,ε−ω filled Fermi sea ε' 0 0 εF Energy εF Energy (a) (b) Figure 4: The scattering process for a quasiparticle with energy  above the Fermi surface involves the creation Figure 3: The probability that a state of a given en- of a particle-hole excitation. ergy is occupied at T = 0: (a) For electrons in a non- interacting system, or Landau quasiparticles in a Fermi example of these modes are the ‘zero sound’ oscillations liquid; (b) For electrons in an interacting Fermi liquid. of the Fermi surface whose restoring force provided by Note the discontinuity at the Fermi energy F remains, the f function. though reduced in size. The ‘jump’, z, is often consid- Before proceeding further we should check, as Lan- ered as the order parameter of the Fermi liquid. dau did, that this procedure is internally consistent. Quasiparticles and holes are only approximate eigen- contributes additively to the total entropy of the sys- states of the system. In writing Eq. 1 we have ne- tem. This is not true for the energy. In an interact- glected the possibility that measuring the energy with ing system we must take into account that, unlike the the Hamiltonian could change the quasiparticle distri- free Fermi gas, the energy of individual excitations will bution (δn~k,σ) itself. (That is to say that there remain not generally add to yield the total system energy. In matrix elements in the Hamiltonian which, when acting Landau’s theory, he accounted for the modified energy on a quasiparticle state, ‘scatter’ it into another state.) through two terms. First, when a quasiparticle moves Recall that acting the Hamiltonian on a true eigenstate there will now be a back-flow in the filled Fermi sea as leaves the wave function unchanged up to a multiplica- the quasiparticle ‘pushes’ the ground state out of the tive constant (the eigenvalue). We can estimate the way. This modifies the inertial mass of the quasiparticle lifetime of these approximate eigenstates by consider- ∗ m → m . (Note that this is in addition to the effect of ing the decay rate of a quasiparticle with energy  above the crystal lattice—which produces a band mass which the Fermi surface at absolute zero. We can use Fermi’s can be included in the free electron picture—and also golden rule the change induced by interactions with phonons.) Sec- X 1 2π 2 ond, a quasiparticle’s energy also depends on the dis- = |Vif | δ( − f ) , (4)  tribution of other quasiparticles which Landau included τ ¯h f via his ‘f function’. The total energy of the interacting system is now expanded as a functional of the quasi- where the sum is over the possible final states f.We particle distribution δn~ : will assume, for the time being, that the scattering ma- kσ | | X X trix elements Vif are constant and we will just enforce pF 1 energy conservation and, crucially, the Pauli principle E = (¯hk−pF )δn~ + f~ ~0 0 δn~ δn~0 0 , m∗ kσ 2 kσ,k σ kσ k σ ~k,σ ~k~k0,σσ0 for quasiparticles. At absolute zero the only scattering (1) allowed by the Pauli principle lowers the energy of the for an isotropic system. Using this one can then com- original quasiparticle by an amount ω by making an pute the equilibrium properties such as the specific electron-hole pair in the filled Fermi sea. This process heat and Pauli susceptibility we considered in the non- is illustrated in Fig. 4. The condition that the quasi- interacting problem above. One finds particle must scatter into an unoccupied state requires ∗ ω<. In addition only occupied states within ω of 1 m pF 2 the Fermi surface can absorb this energy by making a cv = 3 kBT, (2) 3 ¯h particle state above the Fermi surface. Thus our sum ∗ m pF 1 2 over final states becomes χ = 2 a µB . (3) Z Z Z π ¯h 1+F0  ω ∞ 1 2π 2 0 0 00 00 ∼ |V | gF dω gF d δ(−ω− + )gF d , These are similar to the free Fermi gas results except τ ¯h 0 0 −∞ a for the modified mass and the F0 term in χ which is (5) related to the Landau f function and is known as a π 2 3 2 ∼ |V | gF  , (6) Landau parameter. Landau’s theory also predicts new ¯h behaviour as the interaction between quasiparticles al- where gF is the density of states at the Fermi surface. lows for collective modes of the system to develop. An Thus, close to the Fermi surface where  is small, the Non-Fermi liquids A. J. Schofield 5

Box 1. The Kondo Model

The Kondo model gives us a paradigm for under- peaks near TK but then falls linearly to zero as the en- standing how a Fermi liquid arises in a number of the tropy associated with the impurity spin [kB ln(2S +1)] heavy-fermion metals. At its simplest, it describes the becomes quenched in forming the S = 0 singlet. The behaviour of a single spin-one-half (S =1/2) magnetic scattering at T = 0 saturates at the unitarity limit ion in an otherwise non-interacting sea of electrons. and falls as T 2 for very small temperatures. The strik- The model is essentially zero dimensional since all the ing emergence of the Fermi liquid forms for these low action occurs around the location of the ion. Typically temperature properties is due to the low energy ex- this magnetic ion prefers to align its magnetic moment cited states having a one-to-one correspondence with anti-parallel to that of any nearby electron (i.e. anti- a weakly interacting Fermi gas (Nozi`eres 1974). The ferromagnetically). Passing electrons scatter from the Kondo temperature sets the effective Fermi energy of impurity and both can exchange their spin directions this local Fermi liquid. in the process. Kondo (Kondo 1964) showed that, in contrast to most forms of scattering in a metal which normally reduce as the temperature is lowered, this spin-flip scattering grows logarithmically with decreas- ing temperature. Higher order perturbation treatments (Abrikosov 1965) predicted that the scattering would diverge at a finite ‘Kondo’ temperature (TK)—anim- possibility since the most scattering a single impurity Figure B1: At high temperatures the Kondo impurity can do is the unitarity limit when it behaves as an im- scatters conduction electrons but as the temperature penetrable sphere. is lowered the effective interaction between impurity Understanding what really happens for T ∼< TK re- and conduction electrons grows. Eventually a singlet quired ideas of scaling (Anderson and Yuval 1971) and bound-state is formed which acts as an inert potential the renormalization group (Wilson 1975). It revealed scatterer. that the proper way to think of Kondo’s logarithm was to view the strength of the antiferromagnetic interac- The physics of the Kondo model exactly parallels tion between the ion and the electrons as effectively asymptotic freedom and quark confinement in QCD. growing with decreasing temperature. At first this just At high energies we see free spins (analogous to asymp- enhances the scattering but, as the temperature is low- totically free quarks with colour) but, as the energy is ered further, the coupling becomes so strong the that lowered, the spins become bound into singlets (analo- magnetic ion prefers to bind tightly to a single elec- gous to the baryon colour singlets more familiar to us tron and form an inert singlet state. The susceptibility at terrestrial energies). The observation of Kondo type associated with the impurity shows free-spin Curie be- behaviour in UPt3 and other heavy-fermion systems has haviour at high temperatures but, as the singlet forms, been coined ‘asymptotic freedom in a cryostat’ (Cole- the susceptibility saturates. The impurity specific heat man 1993).

1-aT2

B const

χ

C/k -1 -1 -4 T ρ/ρ(0) log T T log T

T T T (a)K (b)K (c) K Figure B2: Impurity contributions to the (a) specific heat, (b) susceptibility and (c) resistivity in the Kondo model as a function of temperature. For T  TK these quantities recover the Fermi liquid forms as the impurity binds to the conduction electrons. Non-Fermi liquids A. J. Schofield 6

A(k,ω ) A(k,ω ) system single electrons are eigenstates of the system so − the spectral function is a delta function δ(ω ~k). In ω2 the interacting system a given electron may take part in many eigenstates of the system and so the spectral function is spread out in energy. Nevertheless for mo- ω menta near kF there is probability z that the electron hp (k-k )/m ω hp (k-k )/m* (a) F F (b) F F may be found in the quasiparticle eigenstate with mo- mentum ~k.SoatT = 0 the electron spectral function Figure 5: The spectral function: the probability that in a Fermi liquid has a sharp peak at the new quasi- − 2 an electron with momentum k may be found with a particle energy with width proportional to (k kF ) , given energy. (a) In a non-interacting system, electrons reflecting the finite lifetime, and an integrated weight are eigenstates and so the probability is a delta func- under the peak of z (see Fig. 5). tion centred on the electron energy, (k). (b) In the Let us pause now to summarize the main features of Fermi liquid this probability is now spread out but re- Fermi-liquid theory. The success of the non-interacting tains a peak at the new quasiparticle energy. This peak picture of electrons is understood in terms of the exis- tence of Fermi quasiparticles as approximate low energy sharpens as k → kF . eigenstates of the interacting system. We find that:

quasiparticle is always well defined in the sense that • Equilibrium properties have the free electron form its decay rate (2) is much smaller than its excitation but with modified parameters. energy (). Far from the Fermi surface (i.e. large ) • The low energy eigenstates are fermion quasiparti- adiabatic continuity will break down since the quasi- cles with a scattering rate 1/τ ∼ max(2,T2). particle will decay before the interaction can be com- pletely turned on. At temperatures above absolute zero • z is the ‘order parameter’ of the Fermi liquid: the ambient temperature sets a minimum energy scale the overlap of an electron and quasiparticle at the so for quasiparticles near the Fermi surface the scat- Fermi surface. tering rate goes like T 2 (Abrahams 1954). Landau’s picture of the interacting electron gas was therefore • New collective modes can also appear. thought always to be valid provided one is concerned with small energy excitations and at low enough tem- How well does this theory perform when tested against peratures. This decay rate of quasiparticles is impor- experiment? tant in determining the transport properties of a Fermi Thus far it looks as though a new free parameter liquid and results, for example, in a T 2 low temperature has been introduced for each experiment. To test the resistivity. theory we should demonstrate that experiments over- determine these free parameters. This is most straight- So far we have just discussed the properties of the 3 Fermi liquid in terms of the quasiparticle. What of the forwardly done in He which is isotropic. There it turns electrons themselves? Adiabatic continuity tells us that out that four experimental quantities (specific heat, in the quasiparticle wavefunction, there must remain compressibility, susceptibility and zero sound velocity) a fraction of the original non-interacting excited state specify three of the Landau parameters and there is wavefunction good internal agreement (see Wheatley 1970). In the √ metallic state the presence of a crystal lattice makes this |ψqp(~kσ)i = z|φel(~kσ)i+particle − hole excitations etc . test harder since the reduced symmetry allows many (7) more Landau parameters. However, remarkably, recent That fraction, z, is known as the quasiparticle weight experiments have confirmed the picture in one of the and, in a sense, plays the role of the order parameter of most strongly interacting metals known: UPt3.The the zero temperature Fermi liquid state. A simple con- key feature of this material is the presence of Uranium sequence of the step in the quasiparticle distribution at f electrons which are tightly bound to the atomic nu- T = 0 is that, if one could analyze the electron distribu- cleus and are surrounded by a sea of conduction elec- tion function, it too would show a discontinuous jump of trons. At high temperatures the f electrons behave size z at the Fermi momentum (see Fig. 3). A theoreti- as free magnetic moments with the classical Curie sus- cal tool for following the fate of the original electrons in ceptibility (Frings et al. 1983). As the temperature is the interacting Fermi liquid is called the spectral func- lowered the free spins start to become bound to con- tion A(ω,~k) (see, for example, Mahan 1990). It mea- duction electrons to make up extremely heavy Landau sures the probability that an electron with momentum quasiparticles (see Fig. 6). [The Kondo model (see ~k can be found with energy ω. In the non-interacting Box. 1) provides our theoretical picture for how this Non-Fermi liquids A. J. Schofield 7

1. The volume of the Fermi surface includes the f electrons. UPt 3 1000 a-axis 2. The measured quasiparticle mass accounts for the enhanced specific heat.

Both these observations confirm the success of Fermi- liquid theory. µeff =2.9

/(mol/emu) 3. The Mystery of the Cuprate Superconduc- -1 500 tors

χ If Fermi-liquid theory gives such a good account of the metallic state, why look for alternatives? The reason is the discovery of metals where the fermion quasiparticle does not seem to reflect the character of the measured low energy eigenstates. In this sense, the electron (or electron-like quasiparticle) may no longer be the appro- 0 500 1000 priate way to think of the low-lying excitations. The T/K prime example is the metallic state of the copper oxide superconductors and so it seems fitting to motivate the Figure 6: The inverse spin susceptibility of UPt3 (after search for alternative theories of metals by summarizing Frings et al. 1983). At high temperatures we see the the puzzles presented by these materials. 1/T behaviour associated with free magnetic moments. The superconducting cuprates encompass almost This large moment becomes bound to the conduction thirty distinct crystalline structures (Shaked et al. electrons at low temperatures to form a heavy electron 1994) and contain upwards of three different elements. Fermi liquid. They are united by the common feature of a layer struc- ture of CuO2 planes. In their pristine state these com- pounds are typically not metals—they are insulators with antiferromagnetic order. (The magnetic moment occurs—although in UPt3 we of course have a dense periodic arrangement of magnetic ions rather than a on the copper site alternates in direction as you move single impurity.] The resulting Fermi liquid indeed re- from one site to its neighbour.) This in itself is a signa- covers the free electron forms for the equilibrium prop- ture that the strength of the interaction between elec- erties. The coefficient of the T -linear term in the spe- trons is important since simple electron counting in the cific heat (the intercept of the graph in Fig. 7 after absence of interactions would suggest that these mate- Stewart et al. 1984), though, is 450 mJ/mol K2—two rials should have a half filled band and hence metallic orders of magnitude larger than that of a free electron properties (see Box 2). In order to make these ma- gas. The effective mass of the quasiparticles has been terials metallic (and superconducting at low tempera- vastly enhanced. This is adiabatic continuity pushed tures) one removes some electrons from each copper- to the limits. To ‘close’ the theory one would like an oxide layer by doping with another element which typ- independent check on the quasiparticle mass. This can ically resides between the copper oxide planes. The be done from high magnetic field and low temperature antiferromagnetism then disappears and the material measurements of the “de Haas van Alphen effect”. At becomes a metallic conductor. sufficiently high magnetic fields, quasiparticles can be The metallic behaviour that arises is characterized by driven around their Fermi surface by the Lorentz force. significant anisotropy. The electrical resistance perpen- The quantization of these orbits leads to oscillations of dicular to the CuO2 plane direction can be up to one the magnetization as a function of applied field: a kind thousand times greater than that for currents carried in of spectroscopy of the Fermi surface. Using this (Taille- the planes (Ito et al. 1991, Hussey 1998). Many peo- fer and Lonzarich 1988, Julian and McMullan 1998) one ple have stressed the importance of the effectively two can map out the shape of the quasiparticle Fermi sur- dimensional nature of the metallic cuprates and this is face and, from the temperature dependence, deduce the in part what has prompted the search for unusual met- quasiparticle effective mass. These masses and Fermi als in low dimensions. As I have already hinted, the surfaces are shown in Fig. 7.Twokeyobservations metallic state itself is unusual. Space precludes a de- emerge even in this most interacting of environments: tailed analysis of all of the anomalous properties so I will concentrate on just a few observations. Non-Fermi liquids A. J. Schofield 8

4

3 m cm) m_* 2 m = 16 17 ab ρ /( Ω

1 38% 24%

18 0 200 400 600 800 1000 T/K

+ Figure 8: The resistivity of La1.85Sr0.15CuO4 is linear in the temperature (after Takagi et al. 1992) and indicates 16% a scattering rate for electrical currents proportional to temperature.

600

2 The first noticed peculiarity is the extraordinary lin- 500 24 ear temperature dependence of the resistivity (Gurvitch 400 and Fiory 1987) at dopings which maximize the su- perconducting transition temperature (illustrated in 300 Fig. 8). Optical measurements confirm that this is due

C/T /(mJ/mol K200 ) to a scattering rate which is proportional to tempera- ture (Forro et al. 1990). While this is clearly different 22% 100 from the T 2 scattering that Landau’s theory might pre- 0 dict, we should remember that the usual quasiparticle 0 100 200 300 T/K2 2 scattering only becomes apparent at very low temper- atures which, in this case, is obscured from view by Figure 7: The consistency of the Fermi liquid descrip- the transition into the superconducting state. Electron- tion has been demonstrated in UPt3. The five Fermi phonon scattering usually gives a linear resistivity from surface sheets (from Julian and McMullan 1998) and moderate temperatures upwards (which is why we often effective mass of the quasiparticles have been mapped use the resistance of a platinum wire as thermometer). out by de Haas van Alphen measurements (Taillefer and However in the case of the cuprates it seems that this Lonzarich 1988). They confirm that the 5f 3 electrons scattering is purely electronic in origin since microwave are absorbed into the Fermi liquid and that the quasi- measurements show the scattering rate plummeting on particle masses are consistent with the mass enhance- entering the superconducting state (Bonn et al. 1993). ments measured in specific heat (after Stewart et al. So the measured resistivity presents us with two puz- 1984). The percentages reflect the contribution from zles: what causes the linear T scattering and how to quasiparticles on each sheet to the total specific heat. explain the absence of the phonon scattering? We have discussed the scattering of electric currents but one can also measure the scattering of currents gen- erated in a Hall effect experiment. Typically both elec- tric currents and Hall currents should measure the same scattering rate. In Fig. 9 we see that the scattering rate from Hall currents rises quadratically with the temper- ature in absolute contradiction to the resistivity exper- Non-Fermi liquids A. J. Schofield 9

Box 2. The tJ model

In the tJ model we imagine a highly simplified view doping. We now remove some electrons so of a CuO2 plane. It starts from a picture of the parent that there are empty sites in the lattice: insulating compounds with a square lattice of single atomic orbitals each with exactly one electron. The systems is said to be half-filled since the Pauli principle would allow a maximum of two electrons in an orbital. Normally at half-filling we would expect a metallic state since any electron can move through the system to carry current by hopping onto a neighbouring site. Now we imagine turning on the inter-electron repulsion so that it becomes energetically unfavourable for more than a single electron to occupy a given atomic site. Figure B3: The tJ model describes the competition be- Now the half-filled case is an insulator because any tween hole motion and antiferromagnetism in a doped electron moving to a neighbouring site already finds Mott insulator. an electron there and pays the price of the repulsive The system becomes a metal since electrons near empty interaction. There is an energy gap to make a current sites can move without restraint (and lower their ki- carrying state and we have a “Mott insulator”. This netic energy by an amount t. The neighbouring elec- explains the insulating nature of the parent cuprates. trons would still like to remain antiferromagnetically In fact no electron really likes to be fixed on a single aligned with strength J and at no time must any site site—it’s like being held in a small box and its kinetic contain more than one electron. The presence of this energy is high. This can be lowered if the electron constraint means that there is no small parameter in made ‘virtual’ tunnelling hops onto the neighbouring the theory with which one can perform perturbation occupied sites and back. This the electron can only do theory. This is the basic physics of the tJ model and, if the neighbouring site has the opposite spin (the Pauli because the moving electrons disorder the magnetism, principle remains absolute). So we see that the inter- we see that the t and the J represent competing inter- actions also favour the antiferromagnetic arrangement actions. While there are many fascinating proposals of spins seen in the parent compounds. for understanding the physics of this model, there are The tJ model describes what happens on few definitive results.

iments (Chien, Wang and Ong 1991). How can a single have is the so called tJ model (see Box 2) for which no quasiparticle have two relaxation rates? solution exists outside one dimension (see later). Furthermore there is the puzzle of how many charge One might be surprised that I have underplayed the carriers there are in these metals. Some experiments, relatively large values of the superconducting transition like penetration depth in the superconducting state temperature for which the high temperature supercon- (Uemura et al. 1991), suggest a number proportional ductors received that epithet. This is not because the to the (small) number of holes made by doping the in- superconductivity is not important or unusual. [These sulating state. Yet other experiments such as angle- are the first materials where the superconducting order resolved photo-emission see a Fermi surface containing parameter has been established to have a symmetry the large total number of conduction electrons in the different from the usual s wave form (see Annett et al. system (Campuzano et al. 1990). 1996).] It is because the understanding of superconduc- The proposed solutions to these questions remain tivity usually requires an understanding of the metallic highly controversial and have led to some very exciting state from which it forms. We should note that our un- and far reaching ideas which, even if they don’t ulti- derstanding of conventional superconductivity relies on mately find fulfillment in the physics of the cuprates, a Fermi-liquid starting point. The pairing instability is will certainly resurface in the physics of other com- a consequence of the sharp discontinuity in occupation pounds. What makes the subject exciting and at the at the Fermi surface which the fermion quasiparticle same time difficult is the absence so far of any solv- picture provides (Bardeen, Cooper and Schrieffer 1957). able model describing how interactions give rise to the The cuprates are not alone in exhibiting superconduc- metallic state of the cuprates. The simplest model we tivity from an unusual metallic state. There is another Non-Fermi liquids A. J. Schofield 10

6000 we have the exciting possibility of the material scien- tist developing materials to demonstrate the theorist’s 5000 model. Since we are here concerned with the failure of Fermi- liquid theory, it is as well to mention one clear example 4000 where adiabatic continuity breaks down. Once level crossings occur in the spectrum then it is no longer H

θ possible to follow the labels through from the non- 3000 interacting case. Typically this happens when there is cot (at 1 Tesla) a phase transition in the system such as the formation of a superconducting state when electron bound states 2000 are favoured. In this review we will be primarily inter- ested in how the Fermi liquid can fail within the normal metallic state so we will not be considering such phase 1000 transitions. However, our first example of a non-Fermi liquid will be one where the approach to a phase tran- sition can disrupt the Fermi-liquid state by destroying 0 123 4 567 8 9 10 11 12 13 the Landau quasiparticle. T2 /(104 K2 ) 4. Metals near a quantum critical point: de- stroying the Landau quasiparticle Figure 9: The inverse Hall angle in La1.85Sr0.15CuO4 in (after Harris et al. 1995). This measures the scattering In our discussion of Landau’s approach we showed that rate for Hall currents and shows it to be proportional a quasiparticle close to the Fermi surface was a long to T 2. This is in marked contrast to the scattering of lived eigenstates by determining the decay rate. In do- electric currents which is proportional to T . The puz- ing this we assumed that the matrix elements for scat- zle is how can a single quasiparticle have two distinct tering were independent of the momentum and energy scattering rates? transfered. The Pauli principle confines all scattering particles to the vicinity of the Fermi surface so there is little scope for large transfers of energy in scattering. In uranium alloy, UBe13, which shows superconductivity the limit of small energy transfer our assumption is of- (Ott et al. 1983, Bucher et al. 1973) but with little ten valid. However, when a system approaches a second evidence of a well formed metallic Fermi-liquid state. order phase transition we know that fluctuations of the It is clear that the properties of new superconductors order parameter slow down and occur over increasingly and their unusual metallic states are intimately linked. long wavelengths. A moving quasiparticle can then eas- If the task is to find new descriptions of the metallic ily generate a large disturbance in the medium which state, where should we begin? There have been many can, in turn, affect other quasiparticles in the vicin- speculative suggestions, but in this article I want to ity dramatically enhancing the scattering cross-section. focus on examples of non-Fermi liquids which we be- This effect is limited by the ordering temperature which lieve we do understand at least from a theoretical view locks the fluctuations into a long range ordered state. point. In fact all four of the examples I will discuss have Below this temperature again our initial assumptions been applied (loosely in some cases) to account for the remain valid and the Landau quasiparticle is saved. Re- physics of the cuprates or uranium alloys. These exam- cently much work (both theoretical and, as we will see, ples show a great richness in behaviour although they experimental) has explored phase transitions occurring may appear at first sight to be rather artificial ‘toy’ at T =0K (Hertz 1976, Moriya 1985, Millis 1993). models. This is a consequence of our lack of the math- While the types of phase transition will be familiar (for ematical tools with which to treat problems where the example from a paramagnetic to a magnetic state) they interaction between electrons is strong and the tools are unusual in that Nernst’s theorem tells us that the that we do have are often only applicable to systems in entropy should always be zero at zero temperature. A reduced dimensions. However, the ingenuity of mate- zero temperature phase transition must therefore be a rial scientists means that these toy systems can often transition between two ordered states. For our pur- be realized in nature by the clever tailoring of mate- poses the most important feature will be that the quasi- rials. Whereas in the past the theoretical physicist’s particle scattering-cross section can now grow without job has often involved finding the simple model that limit ultimately destroying the consistency of Landau’s best describes the physics of a complex material, now Fermi liquid picture. Non-Fermi liquids A. J. Schofield 11

A zero temperature phase transitions occurs at a of metals that do arise in d = 1 later. In higher dimen- quantum critical point—so called because quantum me- sions we need to make D(q, ω) more singular at low q chanics determines the fluctuations of the order param- in order to destabilize the Fermi liquid by this route. eter. It turns out that some of the physical properties Singular interactions are a consequence of long near a zero temperature phase transition can be de- range force laws (large distances correspond to small termined simply by using Fermi’s golden rule, together wavenumbers in reciprocal space). The usual Coulomb with the appropriate matrix elements. The matrix el- force itself adds to the Hamiltonian a term: ement encapsulates the scattering mechanism (in this 0 − ρe(~r)ρe(~r ) −→ ρe(q)ρe( q) case the long range fluctuations of the order parameter 0 Dc(q)= 2 , (9) 4π0|~r − ~r | 4π0q which is trying to develop at the phase transition). In fact I use the example of the magnetic interaction be- where ρe is the density of electronic charge. This has tween moving charges because the matrix elements can exactly the sort of singularity at small q one might ex- be computed from Maxwell’s equations. (It turns out pect to destabilize the Fermi liquid. In fact the Fermi that the matrix elements are the same as those near a liquid remains stable because the collective behaviour of zero temperature transition to a ferromagnetic state.) the other electrons screens the long range Coulomb re- The following section is meant for those readers who pulsion. Any local build up of charge causes the nearby would like to see where some of the non-Fermi liquid electrons move away revealing more of the background properties come from. However, it may be omitted if lattice of positive ions. This neutralizes the charge you just want to know the final results. buildp up beyond the Debye-H¨uckel screening length 2 ξ ∼ 0/2πe gF (Debye and H¨uckel 1923). Thus our Properties near a quantum critical point once long-ranged interaction is actually a short dis- tance Yukawa-type potential and is perfectly innocuous To see in some detail how Landau’s original argument as far as the quasiparticle is concerned. The screened is spoiled near a quantum critical point we must revisit Coulomb interaction no longer diverges as q → 0: our Fermi golden rule expression for the decay rate. I 0 will now do a change of variable and express the same e e 0 e e − ρ (~r)ρ (~r ) −|~r − ~r |/ξ −→ ρ (q)ρ ( q) 0 e Dsc(q)= 2 −2 . quantity as Eq. 6 in terms of the momentum and energy 4π0|~r − ~r | 4π0(q + ξ ) transferred in any scattering process. The result (valid (10) in dimension d)is [Strictly one needs to consider the frequency depen- dence of the electron’s response (i.e. dynamical screen- Z  Z 2k d−1 2 1 2π F q dq |D(q, ω)| ing) to determine the influence of the Coulomb interac- = gF ωdω d 2 . (8) τ ¯h 0 ω/hv¯ F (2π/L) (¯hvF q) tion on the Fermi liquid (Silin 1957, Pines and Nozi`eres 1966).] The integral over ω is simply the number of possible It turns out, however, that the Amperean interac- hole excitations that can be created. The appearance tion between moving charges (more familiar to us as of ω in the lower limit of the momentum integral, q, the force between two current carrying wires) is much appears because a minimum momentum must be trans- more dangerous to the Fermi liquid (Holstein, Norton fered to give a change in energy of ω. The integration and Pincus 1973, Reizer 1989 and 1991). These forces over the direction of the momentum has already been turn out to be much weaker than the Coulomb law and 2 performed and gives the factor of (¯hvF q) in the denom- their danger is only apparent at extremely low energies inator reflecting the increased time available for small and hence at inaccessibly low temperatures. However deflections. Finally D(q, ω) is the matrix element for an almost identical form of matrix element arises near a the scattering process. If this is independent of q and ferromagnetic quantum critical point and this is exper- ω then the q integral is not sensitive to the value of the imentally realizable. Since deriving the matrix element lower limit and is independent of ω. The subsequent at a quantum critical point is beyond the scope of this −1 2 integration over ω recovers the usual τ ∼ gF  result article, I will use the Amperean interaction as my exam- we had before. ple of a singular interaction. In this case seeing why the Already at this point we can see that something inter- force law is singular is a straight-forward application of esting happens in one dimension (d = 1). The singular Maxwell’s equations. nature of the q integral, even when the matrix element The interaction between moving charges is due to the D(q, ω) is constant, leads to a decay proportional to local magnetization which they set up. In the case of  to this order. This is a signal of the breakdown in the Coulomb interaction the potential energy term in adiabatic continuity since there is no limit when the the Hamiltonian is just φ(~r)ρe(~r): the charge density quasiparticle energy is well defined: There are no one times the electrostatic potential. The term we need dimensional Fermi liquids! We will discuss the nature for the Amperean force law is the product of the local Non-Fermi liquids A. J. Schofield 12 p current density and the vector potential: A~(~r)·~j(~r). For behaves like ω/q and this is the case we will con- the Coulomb case we know that a local charge produces sider. Doing the integral of Eq. 8 maylooktrickybut a1/r potential. To find the vector potential from a the singular nature makes everything simple. We are local current density, ~j, we consider the fields generated only trying to extract the energy dependence of the de- by its magnetization m~ in the presence of the currents cay rate so we will neglect the prefactors. We note that 2 from other electrons in the metal J~. Using Maxwell’s as q decreases then DAmp grows as 1/q until q be- equations comes smaller than the inverse skin depth when DAmp saturates. This happens when q2 ∼ ω/q (i.e. when ∂B~ ∂E~ q ∼ ω1/3). Since ω1/3 is always greater than the lower ∇×~ E~ = − , ∇×~ B~ = µ0J~+µ00 +µ0∇×~ m,~ ∂t ∂t limit of the integral (ω)asω → 0 we can use it as the (11) lower limit. We can then find the energy dependence and the definition of the conductivity in the metal of the scattering rate by considering ~ ~ ~ J = σE, we can find the magnetic field B and hence Z Z   ~ ∇×~ ~  d−1 2 the vector potential B = A. It is usual to Fourier 1 ∼ q dq 1 ωdω 2 2 , (15) transform the result to obtain the appropriate interac- τ 0 ω1/3 q q tion term ∼  ∼ T for d = 3 at temperature T.(16) ~ · ~ ~ ~ · ~ µ0j~q,ω j−~q,−ω There is now no regime where the Landau quasipar- DAmp(~q, ω)=A(j) J = 2 2 2 . (12) q + iωσµ0 − ω /c ticle is sufficiently long-lived to count as an approxi-

2 2 mate eigenstate. In the language of adiabatic conti- (The ω /c term does not play an important role here nuity, switching on the interaction adiabatically takes and so we will drop it from now on.) By comparing the longer than the lifetime of the eigenstate itself and so equation for this interaction with that of the screened one can never continue from the non-interacting state Coulomb interaction (Eq. 10) we see that the appropri- to the interacting one. When the scattering rate goes ate “screening length” for the current-current interac- linearly to zero with the energy as in this case, we have tion is the skin depth a ‘marginal Fermi liquid’ (Varma et al. 1989). p Surprisingly, despite the absence of quasiparticles, we ξ ∼ 1/ iωσµ0 . (13) can still use this calculation to determine some prop- Unlike the Coulomb case considered previously, this erties of this non-Fermi liquid metal: the temperature screening length diverges at low ω (i.e. low energies) dependence of the resistivities and heat capacity. The and so fails to usefully suppress the growing scatter- decay rate (which is related to the resistivity) and the ing matrix element at low energies. For the Amperean effective mass (which gives the heat capacity) are in- interaction, the quasiparticle scattering cross section extricably linked in this example. This is because the grows without the limit and this destroys the Fermi liq- scattering of quasiparticles can not help but produce uid. Since this type of interaction is always present in some back-flow which contributes to the effective mass a metal, why is Fermi-liquid theory a good description and the quasiparticle energy. The decay rate may be of any metal? The answer lies in the overall scale of the viewed as an imaginary component of the energy in ∼ − interaction. Comparing the ratio of the Coulomb and the time evolution of the wavefunction exp( it/¯h). → Amperean interactions we see that (for currents car- When the decay rate is non-analytic as  0then ried by quasiparticles near the Fermi surface moving one can obtain the real part from the imaginary part with velocity vF ) through the Kramers-Kronig relation. The essence of this is that if the quasiparticle scattering rate is T α   2 2 (with α some fractional power), the heat capacity also DAmp ∼ je ∼ vF 4π0µ0 2 4π . (14) α Dsc ρe c has a T dependence. In the case we have just consid- ered α = 1. This is non-analytic at the origin since, Thus the current-current interaction is 106 times weaker as a decay rate is always positive, it should really be −1 than the Coulomb interaction which is why its effects τ ∼||. In that case the heat capacity become T ln T . are only likely to be visible at micro Kelvin tempera- So we have a logarithmic correction to the usual linear tures. in T specific heat capacity. We can see what behaviour might be produced, The other quantity which we can compute is the though, by computing the quasiparticle lifetime from resistivity. One might be tempted to conclude that, Eq. 8 with Dj−j as found above. In a clean metal with since the scattering rate is linear in temperature, then no impurities the conductivity is limited, not by the the resistivity should also proportional to temperature. mean-free-path, but by the wavevector: σ ∼ 1/q.The However this fails to account for the effectiveness of skin depth enters the so-called anomalous regime and the scattering at destroying electrical current. Small q Non-Fermi liquids A. J. Schofield 13

0.3 scattering may destroy the quasiparticle but it is not 120 15 effective at removing momentum from the net flow of 300 cm] 2 10 current. For that to happen large angle scattering must 200 D 2 c

90 T/K 0.2 occur. The transport lifetime takes this into account 100 ρ / [µΩ 5 4/3 and is obtained from the same expression as Eq. 8 with 0

8 P /kbar 16 22 0

2 2 4/3 an additional factor of (1 − cos θ) ∼ q /k where θ c 60 0 50 100 F T/K A B CD 1.6 1.6

0 T/K is the scattering angle. Doing this above gives a re- 0.1 5/3 C sistivity of T - a stronger temperature dependence 30 B

2 ρρ µΩ than the usual T in a Fermi liquid. The other quan- magnetic ( (T)- )/T /A [ cm/K ] tity we have discussed, the Pauli susceptibility, can not 0 0 be obtained from our Fermi golden rule expression and 0 5 10 15 20 0102030 Pressure /kbar requires more analysis to obtain. (a) (b) Temperature /K

Examples of quantum critical points Figure 10: MnSi: a low-temperature long-wavelength Summarizing these results, we have shown how the in- helical magnet. (a) The phase diagram showing that teractions between currents will ultimately destroy the quantum critical point occurs at 14.8 kbar. (b) As the Fermi-liquid state in any metal at very low tempera- quantum critical point as approached, the resistivity 5/3 tures. The new behaviour that we expect to see in- takes on a T temperature dependence showing that cludes the quasiparticles are more strongly scattered than in −1 a normal Fermi liquid (after Pfleiderer et al. 1997). • a marginal quasiparticle scattering rate: τ ∼  ∼ T , diverges at low temperatures when compared to the ex- • CV ∼ T ln T , pected Fermi-liquid form until the phase transition to • χ ∼ ln T (not proved here), the ferromagnet occurs. When this phase transition is pushed to absolute zero we have a true quantum crit- • ∼ 5/3 ρ(T ) T . ical point and this divergence proceeds without limit. 5/3 It is a curious paradox that, while Landau’s arguments The resistivity takes on the new T form we proved would suggest that the Fermi-liquid description is valid earlier. Similar behaviour has also been seen in high in the low temperature limit, interactions like this actu- pressure experiments on ZrZn2—another ferromagnet ally provide a low temperature bound on its stability. with a low Curie temperature (Grosche et al. 1995). The Kohn-Luttinger instability (Kohn and Luttinger There are a number of other known non-Fermi liquids 1965) to a superconducting state similarly acts in a gen- which arise from singular interactions. A very simi- eral way to prevent one ever obtaining a T =0Fermi lar form of interaction occurs when a two dimensional liquid. In most metals these effects are unobservable, electron gas is subjected to high magnetic fields, but it but there are a growing number of cases where new comes from quite a different source. This perhaps is the types of singular interactions can lead to a non-Fermi most unusual Fermi liquid we know, although again the liquid state which is observed. I’ve already indicated singular interactions imply that it is not truly a Fermi that metals near a quantum critical point provide us liquid at the lowest temperatures. with such examples and so too do electrons in a half- At high magnetic fields we see the fractional quantum filled Landau level and I will now discuss what can be Hall effect (Tsui, St¨ormer and Gossard 1982). Electrons seen in experiment. in a magnetic field and confined to two dimensions de- When a metal is on the verge of a ferromagnetic insta- velop a discrete quantized energy spectrum where each bility then one finds that the effective interactions be- level can hold a macroscopic number of particles— a tween quasiparticles have exactly the same form as they number which depends on the strength of the magnetic do in the Amperean case we’ve just considered. One ex- field. When the lowest energy level is partially filled by ample of this is in the compound MnSi (Pfleiderer et al. a fraction with an an odd denominator (like 1/3 full), 1997). It becomes an almost ferromagnetic helical mag- the ground state shows unusual stability. The Hall ef- net at 30K but under pressure this Curie temperature fect develops a a plateau and becomes independent of is lowered, until at 14800 atmospheres the magnetism a field for a small range of nearby fields as the elec- has been completely ‘squeezed’ out of the system. It tron fluid is reluctant to move away from this stable has not yet been possible to measure the heat capacity point. The excitations of this insulating state carry under such extreme conditions but resistivity measure- fractions of the electronic charge (Laughlin 1983) and ments can be done. In Fig. 10 we see how the resistivity are a fascinating area of active research which I won’t Non-Fermi liquids A. J. Schofield 14

10 60 paramagnet µΩ 40 15 cm]

TN 20

5 ρ / [µΩ 0 081.2 1.2 Resistivity / [ cm] T/K 0 020406080 Temperature /K antiferromagnet T/K1.2 1.2 T S Figure 12: The resistivity of CePd2Si2 at the critical 0102030pressure (28 kbar). The observed temperature depen- dence, T 1.2, is seen over two decades of temperature. It Pressure /kbar has not yet been explained. (Data after Grosche et al. 1996.) Figure 11: CePd2Si2: a low temperature antiferromag- net. Under pressure the antiferromagnetism can be sup- pressed to zero temperature giving a quantum critical cal points. There exist a growing family of metals with point. Not only do we see non-Fermi liquid behaviour very low Ne´el temperatures below which antiferromag- here but also there is a superconducting transition (af- netic order develops. An example is CePd2Si2 where ter Julian et al. 1996, Mathur et al. 1998) the Ne´el temperature can be squeezed to zero in pres- sures of 28000 atmospheres (see Fig. 11). Similar argu- ments to those presented above can be used to compute detail here as I am concentrating on metallic states. the expected temperature dependence of the resistiv- The ground state can be thought of as being formed ity: it turns out to be T 3/2. However, unlike the fer- from bound states of electrons with quanta of flux (Jain romagnetic case, the power law that is observed is T 1.2 1989 and 1992) making ‘composite ’. What- (Grosche et al. 1996) (see Fig. 12). Why this should be ever magnetic field is left unbound after the compos- is presently not understood but is potentially a question ite fermions form, is the effective field experienced by of fundamental importance. The cuprate superconduc- the composite fermions. They then undergo a conven- tors are also systems close to antiferromagnetism as we tional integer quantum Hall effect. Between the quan- have seen, and it has been argued that these two puz- tum Hall plateaux the electron gas passes through a zling phenomenon are linked. Tantalizingly, this system metallic phase. When the lowest energy level is ex- and others close to antiferromagnetism (Mathur et al. actly half filled the composite fermions try to form a 1998) also show superconductivity at the quantum crit- metallic state where there is no effective field remain- ical point lending weight to a connection between this ing. We then have a Fermi liquid of composite fermions type of quantum critical point and the cuprates. How- (Halperin, Lee and Read 1993). The residual interac- ever the superconducting transition is at 0.4 K which tions in this metal can also have a singular form and so leaves this scenario needing to explain why Tc is so high affect its properties. These come from an interaction in the cuprates. between currents and charge density which is left over In our discussion thus far we have used only the resis- from approximations of binding magnetic flux to the tivity as the signature of non-Fermi liquid physics. Pri- underlying electrons. This type of interaction (coupling marily, this is because all of the above materials require a vector to a scalar) is not usually allowed in a metal pressures so high to reach the critical point that ther- but occurs here in the presence of a magnetic field. It modynamic measurements (e.g. specific heat or Pauli is the absence of a long range interaction (which would susceptibility) are rather difficult. This can be over- otherwise suppress the fluctuations of density) that can come by using doping rather than pressure to tune the causes the effective coupling to be singular. Neverthe- metal to a quantum critical point. One such exam- less one sees strong evidence for a well formed Fermi ple is CeCu6−xAux (see von L¨ohneysen 1996) which liquid in the experiments (Willett 1997). is paramagnetic with x = 0 but on adding a small Perhaps the most puzzling of the systems with a sin- mount (x =0.1) of gold, the metal develops antifer- gular interaction are antiferromagnetic quantum criti- romagnetism (see Fig. 13). Here one can look at all the Non-Fermi liquids A. J. Schofield 15

2.5

CeCu6-x Au x 2.0 paramagnet 1.5 N antiferro- T/K 2

magnetic 2 1.0

0.5 single crystals polycrystals 0 0.0 0.5 1.0 1.5

x C/T /(J/mol K ) 1

Figure 13: The CeCu6−xAux system shows an antifer- romagnetic quantum critical point driven by gold dop- ing. Using doping to tune to the critical point opens up the possibility of doing more measurements on the new metallic state. It comes at the price, though, of in- 0.04 0.1 14 creasing the sample disorder which can complicate the theoretical understanding of the non-Fermi liquid state T/K (after Pietrus et al. 1995).

Figure 14: The specific heat of CeCu5.9Au0.1 at the familiar indicators of Fermi-liquid behaviour and show quantum critical point. The heat capacity shows a how they deviate near the critical point. The specific T ln T form indicating that the thermodynamics of the heat shows a T ln T (see Fig. 14), the resistivity is linear non-Fermi liquid are totally changed by the proximity in temperature and the Pauli susceptibility diverges as to the critical point. (Data after von L¨ohneysen et al. − ln T at low temperatures. The danger with doping 1996.) as a tuning mechanism is that the metals are now off stoichiometry and, as such, one must wonder about the role of disorder near the critical point. This is theoreti- the Luttinger liquid. Discussing one dimension may cally an open problem. Nevertheless, the experimental seem rather esoteric when we live in a three dimen- situation is clear: we have a well tried route to the sional world. In fact many systems, from the “blue non-Fermi liquid at a quantum critical point. bronze” molybdenum alloys to some organic Bechgaard salts have properties which are highly anisotropic. Elec- If all that was known about non-Fermi liquids was tron motion is essentially confined to one dimension by how the quasiparticle could be destroyed by singular the very low probability of the electron hopping in the interactions, we would seem to have found only the two remaining directions. It is in this type of system exception to prove the general rule. However in one that we have the possibility of seeing a Luttinger liquid dimension we see a radical new type of metallic be- state develop. haviour where completely new types of particle emerge to replace Landau’s Fermi quasiparticle. If the essence of the Fermi liquid was the Landau quasiparticle, then the essence of the Luttinger liquid is spin-charge separation and the appearance of spinon 5. Luttinger Liquid: the Bose Quasiparticle and holon quasiparticles. Their existence relies on a We have already seen that our general scattering rate very special property of one dimensional systems: near argument would predict an absence of Fermi liquids in the Fermi surface all particle-hole excitations at fixed one dimension even with a constant matrix element. momentum have the same kinetic energy. This is illus- Considering higher order terms only makes matters trated in Fig. 15 where we see that in two (and higher worse. All is not lost however, for a new type of adi- dimensions) the energy depends both on the magni- abatic continuity has been proposed by Haldane (Hal- tude of q and on its direction relative to the local Fermi dane 1981) which gives us the possibility of quantify- surface. In one dimension there is only one direction ing the new metallic state that emerges in its place: and so fixing q determines the energy completely. This Non-Fermi liquids A. J. Schofield 16

Energy

(i) δq δE k 0 δE δq (ii)

(a) (b)

Figure 15: One dimension has the special feature that

k v k all particle-hole excitations with a given momentum ~q v have the same energy. (a) In high dimensions one can make (i) a high energy excitation or (ii) a low energy Figure 16: The spectral function of a one dimensional one depending on whether ~q is normal or transverse to Luttinger liquid. Notice how, in contrast to the Fermi the local Fermi surface. (b) In d = 1 there is only one liquid (Fig. 5), there are now two singular features cor- direction in the problem and so fixing δ~q, determines responding to the spinon and holon and they generally the energy change δE. This leads to density waves be- disperse with different velocities, vσ and vρ (after Voit ing the proper description of physics in one dimension. 1993 and Dias 1996). is important because adding together the all possible electron moving with differing velocities (see Fig. 16). particle-hole excitations with a specific excitation mo- One can make a very simple picture of how this hap- mentum ~q gives the wavefunction for a density wave: a pens by considering a single electron in a Mott insu- compression and rarefraction of electron density with a lating state. This is illustrated in Fig. 17 where we wavelength 2π/|~q|. The special property of one dimen- consider the physics of the tJ model (Box 2) but now sion means that here a density wave has a well-defined in one dimension. Starting with the insulating state we kinetic energy. Now the potential energy is also usu- have an antiferromagnetic arrangement of spins. Now ally determined by the density of particles with a given we remove an electron which of course removes both a wavelength (see for example the Coulomb interaction of spin and leaves behind a charged state. As the hole now Eq. 10) and so the density wave also has a well defined moves we note that the place where the disruption in potential energy. This is enough to tell us that density the spin arrangement and the position of the hole have waves form the new eigenstates of the one dimensional now moved apart. The spin and charge of the original metal. More generally the potential energy can depend electron have separated and formed independent enti- on both the density of spin and the density of charge, ties. so spin density waves may have a different energy from This type of picture has, for many years, seemed charge density waves. Thus the good quantum num- no more than a na¨ıve picture of a phenomenon whose bers of the system are those of spin and charge density. proper description requires the powerful mathematical So we can completely by-pass the problem of how the machinery of bosonization. (This is the technique which electrons behave by working only with the densities. formally expresses the problem in terms of spin and This leads to the remarkable phenomenon of spin- charge densities.) However, very recently exactly the charge separation. The electron carries with it both experiment described above has been done on the an- spin (its magnetic moment) and its electrical charge. tiferromagnetic chain compound SrCuO2 (Kim et al. In one dimension these can, and generally do, become 1996). A single electron is removed from the chain by two separate entities which move independently as they a photo-emission process whereby an incident photon form the spin and charge density eigenstates. The elec- kicks out an electron. The probability for doing so tron dissolves into its spin part (a spinon) and its charge depends on the underlying dynamics of the spin and part (a holon). Its clearly not a Fermi liquid any more charge degrees of freedom we excite. Since this starts because the good quantum numbers look nothing like as an insulating system the dynamics of the hole are the old fermion quasiparticle labels. If we ask where the the dynamics of an almost empty band. The allowed original electron has gone by determining the spectral momentum of the spinon is restricted by the magnetic function, we no longer see the single sharp quasiparti- order. The spectrum that is seen is shown in Fig. 18d. cle peak of the Fermi liquid. Instead we see two sharp For some momenta of the photo-electron, only the holon features characterizing the spin and charge parts of the is allowed to carry the momentum away and one sees Non-Fermi liquids A. J. Schofield 17

2t Holon dispersion πJ/2 Spinon dispersion

kh k s 0 0 1.0 2.0 1.0 2.0 Holon energy Spinon energy

-2t −πJ/2 (a) (b) (a) Photoemission band 1.0 2.0 0.0 ??? 0 k// πJ 2 -0.5

-1.0 (b) 2t Energy to the top of the band

Energy from highest-1.5 peak /(eV)

0 0.5π π (c) (d) wave vector

holon spinon Figure 18: Photo-emission experiments can actually re- produce the physics of the simple picture of Fig. 17— (c) here in SrCuO2. (a) The removal of a single electron creates a holon in an otherwise empty band. (b) The magnetic order restricts the allowed momentum of the Figure 17: A simple picture of spin-charge separation spinon: the thick line shows the forbidden regions. (c) in one dimension. Consider the 1d tJ model when an For certain momenta of emitted photo-electron there electron is removed from the antiferromagnetic Mott is just a single way in which the momentum can be insulating state by a photon in a photo emission exper- distributed between the spinon and holon but in other iment (a). This leaves behind a disruption in both the parts of the zone there is no such restriction. (d) In the spin and charge order. (b) As electrons move into the measurements we see rather broad features where the vacant site, the locations of the spin and charge disor- momentum and energy is distributed between a num- der separate. They have become distinct particles—a ber of possible spinon and holons states. In other parts spinon and a holon. of the zone a single dispersing peak is seen (after Kim 1996 and Shen 1997). a well defined dispersing peak following the dispersion of the holon. For other momenta, both the spinon and the holon can be excited and so the momentum blesome processes (Luttinger 1963). Haldane’s Lut- is distributed between them. Instead of a sharp(ish) tinger liquid hypothesis parallels adiabatic continuity quasiparticle peak in the spectrum one sees for these in the interacting Fermi liquid but now using the one momenta a broader spectrum distributed between the branch Luttinger model as the starting point. Much band energies of the spinon and the holon (see Fig. 18). as adding interactions to the non-interacting Fermi gas The observant reader will perhaps have noticed that leads to renormalization of (Landau) parameters to the special property upon which all of this relies is a form the Landau Fermi-liquid state, in one dimension consequence of the linearity of the energy spectrum near the additional processes which spoil the special proper- a single Fermi point. In reality scattering from near one ties of the Luttinger model just lead to a renormaliza- Fermi surface point to the other side (Fig. 19a) or dis- tion of the parameters in the model. In fact Haldane persion curvature as one moves away from the Fermi showed there were just four free parameters which char- points (Fig. 19b) means that the momentum of the acterize the low energy properties of one dimensional particle-hole excitation no longer uniquely determines metallic states. These properties not only set the values the energy. The Luttinger model is a simplified version of the spinon and holon velocities but also the so-called of the metallic states that does not contain these trou- anomalous exponents which control, among other prop- Non-Fermi liquids A. J. Schofield 18

Energy Energy decay of holons and spinons respectively. While this proposal remains controversial, the idea of spin-charge δq' separation in more than one dimension is a very active ε research area at present. εF F 6. Two-channel Kondo model: novel quasipar- δq ticles k k Thus far we have had Fermi quasiparticles (in the Fermi -kF kF -kF kF (a) (b) liquid), Bose quasiparticles (in the Luttinger liquid) and no quasiparticles at all near a quantum critical point. That might be expected to exhaust the possi- Figure 19: (a) Scattering from one Fermi point to the bilities! In fact in low dimensional systems there can other or (b) large energy excitations (and also Umklapp exist excitations which fall outside these classes. I have scattering), spoil the special property of 1d illustrated already mentioned the possibility of particles carrying in Fig. 15. The Luttinger liquid hypothesis argues that fractional charge in the fractional quantum Hall effect. the density wave eigenstates are still adiabatically con- The excitations of these systems in two dimensions have tinuous with the true low energy eigenstates even in the quantum statistics that can lie in between fermions and presence of these processes. bosons. Under particle exchange these particles acquire a more general phase factor (eiθ) in contrast to the usual 1 for bosons/fermions and are known as “anyons”— erties, the nature of the singularities seen in the electron for ‘any statistics’ (Leinaas and Myrheim 1977)! spectral function. Interactions then can lead to completely new types Identifying unambiguously a Luttinger liquid is made of state appearing at low energies. Rather than use harder by the fact that the thermodynamic probes of anyons as an example, I will discuss the appearance of specific heat and Pauli susceptibility retain their old a new type of excitation in a metal in the “two channel Fermi liquid forms even in a Luttinger liquid. Using Kondo” problem. The strange particle in this problem the resistivity to identify a Luttinger liquid is compli- is essentially “half of a spin-half” degree of freedom. cated by the issue of how impurities control the scat- Kondo models have for many years been a favourite tering (in one dimension a single impurity limits the of condensed matter theorists and the single channel current by blocking the current path). In addition one Kondo model is described in Box 1. dimensional systems are typically rather unstable to If the physics of the single-channel Kondo model is long range ordered spin- and charge-density wave forma- that of a local Fermi liquid, then the two-channel case tion which can completely destroy the metallic state. It is the physics of the local non-Fermi liquid. In the two- seems that the most unambiguous probe should be the channel case, one imagines a single spin one-half im- photo-emission experiment of the type performed above purity which interacts antiferromagnetically with two which potentially could measure the spectral function conduction seas of electrons (hence the two channels) of Fig. 16 (see Gweon et al. 1996 and Voit 1998). Nev- which do not otherwise interact (Nozi`eres and Blandin ertheless, from a theoretical perspective this non-Fermi 1980). The conduction electrons are totally oblivious of liquid state in 1D is convincingly established and has the other sea of electrons and do not even experience a been found relevant for experiments ranging from the Pauli exclusion principle from them. Only the impurity TMTSF Bechgaard salts TMTSF (Bourbonnais et al. sees that there are two channels. Experimentally this is 1984, Wzietek et al. 1993, Dardel et al. 1993 and Zwick hard to realize as described above, but there are claims et al. 1997) to quantum wires (Tarucha et al. 1995 that tunneling experiments through certain two level and Yacoby et al. 1996) to edge state tunnelling in the systems can be modeled in a very similar way (Ralph fractional quantum Hall effect (Milliken et al. 1996 and et al. 1994). Chang et al. 1996). The extra complication of two conduction channels Phil Anderson has suggested that the Luttinger liq- does little to affect the physics at high temperatures. uid may not only be confined to the realm of one di- For weak coupling one has a free spin one-half object mension but may be the appropriate starting point for which scatters both channels of electrons and results understanding the two dimensional metallic state of the in the same logarithmically growing scattering as the cuprates (see Anderson 1997). Part of the supporting temperature is lowered. However, as the coupling con- experimental evidence which he appeals to is the two stant grows, the impurity spin has a problem. It would separate scattering rates measured in the decay of elec- like to form a singlet but the symmetry of the problem trical and Hall currents. These he attributes to the forbids it from favouring any one of the channels for Non-Fermi liquids A. J. Schofield 19

rity spin’s entropy. In the single-channel Kondo model this falls smoothly from ln 2 reflecting the two degrees of freedom of the free spin, to 0 at very low temper- atures. In the two-channel case the entropy also falls 1 from ln 2. However√ it saturates at 2 ln 2 at low temper- atures as if a 2 degree of freedom is left. It transpires that this object can be represented rather simply as the real part of the normally complex electron—a so-called Majorana fermion named after Majorana’s purely real representation of the Dirac equation (Majorana 1937). It remains free at low temperatures and disrupts the local Fermi liquid one had in the single-channel case. New power laws emerge for the impurity spin contri- bution to the low temperature properties: heat capacity ∼ T ln T ,√ Pauli susceptibility ∼ ln T and the resistiv- ity has a T correction. While it is hard to make a straightforward application of this model to a physical system, there have been a number of proposals suggest- ing that this physics can be realized in certain uranium alloys (Cox 1987) as well as in the various tunnelling problems mentioned before. In fact strange power laws are seen in a number of uranium alloys (see , for exam- Figure 20: In the two-channel Kondo problem, a single ple, Maple at al. 1996), but the two-channel Kondo magnetic impurity interacts with two orthogonal elec- interpretation remains controversial. Nevertheless it tron wavefunctions. The magnetic ion can no longer provides us theoretically with the intriguing possibil- make the usual non-magnetic singlet state at low tem- ity that the physics of non-Fermi liquids may lead to peratures because the symmetry of the problem makes completely new kinds of low energy particle controlling favouring one electron over the other impossible. A the behaviour of exotic metals. magnetic state can still undergo further Kondo scatter- ing. The resulting ground state bears no resemblance 7. The disordered Kondo scenario to a non-interacting gas of electrons and is a local non- Fermi liquid. In our survey of non-Fermi liquids I have deliberately tried to choose examples which are stoichiometric - that is where a pure crystal exhibits an unusual metallic making that singlet. The other possibility is to make a state. The reason for doing this has been to isolate the linear superposition of a singlet with each channel, but role of disorder which further complicates our picture this leaves the unbound spin of the spectator channel of non-Fermi liquid alloyed materials. Although there carrying a two-fold spin degeneracy. It turns out that exist many powerful techniques for dealing with dis- this then would behave like a new spin-half impurity order in condensed matter physics, they unfortunately which in turn wants to undergo another Kondo effect are not easily extended to include the effect of interac- (see Fig. 20). There is no simple solution to the impu- tions. Having said that, there are in fact many more rity spin’s dilemma. Solving the problem requires the examples of non-Fermi liquid metals which one might application of conformal field theory techniques (Af- claim to be disordered. Almost all of the cuprates are fleck and Ludwig 1991) and the ‘Bethe Ansatz’(Andrei made by partially substituting one atom for another in and Destri 1984, Tsvelik and Wiegmann 1985)—a class the process of doping the Mott insulating state. (The of wave functions which solve a number of interacting reason that we believe that these are in fact clean ma- low dimensional problems. The mathematical complex- terials is because the copper oxide planes—where the ity of these solutions forbids detailed discussion of them action is taking place—are left unscathed by the dop- here. There do exist a number of ‘simplified’ treatments ing except for a change in carrier density.) There exist (Emery and Kivelson 1992, Sengupata and Georges also a growing number of diluted alloys of uranium and 1994, Coleman Ioffe and Tsvelik 1995) which reformu- cerium which also exhibit non-Fermi liquid behaviour. late the two-channel Kondo model (see Schofield 1997) Some have been interpreted in terms of the two-channel and make the physical properties obtainable from per- Kondo scenario mentioned above. There is, however, at turbation theory. What emerges can be seen from least one other possibility, namely that the non-Fermi the calculated temperature dependence of the impu- liquid behaviour is coming from single channel Kondo Non-Fermi liquids A. J. Schofield 20 impurities but the disorder creates a distribution of We see at once that the specific heat and the suscepti- Kondo temperatures (Bernal et al. 1995, Miranda, Do- bility immediately adopt non-Fermi liquid forms at low brosavljevi´c and Kotliar 1997). One such example is temperatures with C/T ∼ χ ∼−ln T instead of T in- UCu5−xPdx (Bernal et al. 1995) where one imagines dependent. The rising resistivity as the temperature is the magnetic uranium ions sitting in a random envi- lowered suggests disorder in the system as well as a low ronment of a Cu/Pd alloy. energy scattering mechanism. Of course these results I have argued that one such impurity favours a Fermi depend to some extent on the distribution of Kondo liquid (see Box 1) so why is it that a distribution of temperatures but, provided this distribution tends to a such impurities can lead to anything different? The finite number (not zero) at low temperatures, the low answer lies in the fact that even relatively weak dis- temperature forms are robust. In fact this type of be- order yields a fraction of the impurity spins with ex- haviour has been seen in many alloyed materials (see tremely low Kondo temperatures and these magnetic Miranda, Dobrosavljevi´c and Kotliar 1996 for a table). moments remain unquenched and strongly scatter the To be sure that disorder is driving the non-Fermi conduction electrons even at low temperatures. This liquid physics we would like some independent signa- is a consequence of the exponential dependence of the ture of it. This can come from nuclear magnetic res- Kondo temperature on the local properties of the im- onance (NMR) and muon spin rotation (µSR) studies purity spin: (MacLauglin, Bernal and Lukefahr 1996). These mea- TK = D exp(−λ) , (17) surements probe very precisely the local environment where D is the bandwidth of the conduction electrons at particular atomic sites. What is found is that in and λ is a measure of the local density of conduction UCu5−xPdx the copper atoms appear to sit in a variety electronic states at the magnetic site and of the coupling of local environments strongly suggesting the presence between the moment and the conduction electrons. of disorder. These authors are even able to extract the We can perform a crude calculation of the non-Fermi distribution of Kondo temperatures and show that it liquid properties by assuming that the Kondo temper- does satisfy the requirement of being finite at low tem- atures in the alloy are uniformly distributed between 0 peratures, and is consistent with the measured heat ca- pacity and susceptibilities. While this interpretation is and an arbitrary scale, To. By using approximate forms for the specific heat, the resistivity and the suscepti- not universally accepted in this particular material (see, bility of a single Kondo impurity, we can then simply for example, Aronson et al. 1996) it does serve as the average over the distribution of Kondo temperatures to simplest example for a new route to non-Fermi liquid obtain the bulk response. Our approximate forms will physics when interactions and disorder combine. be 8. Conclusions ∼ TK T Cimp kB 2 2 , (18) The discovery of the cuprate superconductors has T + TK 2 sparked a widespread interest in materials which do not µB χimp ∼ , (19) seem to lie with in the traditional Fermi-liquid frame- T + TK work which we have relied on for understanding the 0 T>TK , effect of interactions in metals. What is emerging from ρimp ∼ (20) ρo T ≤ TK ; this is a striking richness in the types of metallic be- haviour that can appear. We see metals where the These have been chosen to capture the essence of electron dissolves into its magnetic and electric com- the results for the Kondo model shown graphically ponents and systems when no quasiparticle excitation in Box 1 while obeying certain important constraints R ∞ is left. We also see the possibility of unusual states (such as the total impurity entropy Cimp/T dT be- 0 appearing which have no simple analogue outside in- ing independent of TK.) Using a uniform distribu- teracting systems. tion of Kondo temperatures (P (TK )dTK =1/To)we In this article I have attempted to give a flavour of can straight forwardly calculate the expected proper- some of the ideas which are currently being explored ties by averaging over the impurity distribution (eg. R T both to understand the cuprates but, often more fruit- ∼ o C(T ) Nimp/To 0 CimpdTK ). One finds that fully, in understanding equally fascinating problems in   2 other metallic compounds. It may well be that the so- ∼ T To C(T ) NimpkB ln 1+ 2 , (21) lution to the mystery of the cuprate metals also lies in 2To T some of the physics discussed here. However my suspi- 2 µB cion is that nature is playing stranger tricks and that χ(T ) ∼ Nimp ln (1 + To/T ) , (22) To there is a new theory of interacting metals just as pro- ρ(T ) ∼ ρo (1 − T/To)forT

Landau, L. D., 1958, Sov. Phys.–JETP, 8, 70. Shen, Z.-X., 1997, unpublished, Adriatico mini- Laughlin, R. B., 1983 Phys. Rev. Lett., 50, 1395. workshop, Trieste. Leinaas, J. M., and Myrheim, J., 1977, Il Nuovo Ci- Sommerfeld, A., 1928, Z. Phys., 47,1. mento, 37,1. Stewart, G. R., Fisk, Z., Willis, J. O., and Smith, J. L., von L¨ohneysen, H., 1996, J. Phys.:Condens. Matter, 8, 1984, Phys. Rev. Lett., 52, 679. 9689. Taillefer, L., and Lonzarich, G. G, 1988, Phys. Rev. von L¨ohneysen, H., Sieck, M., Stockert, O., and Waf- Lett., 60, 1570. fenschmidt, M., 1996, Physica B, 223-224, 471. Takagi, H., Batlogg, B., Kao, H. L., Kwo, J., Cava, R. Luttinger, J. M., 1963, J. Math. Phys., 15, 609. J., Krajewski, J. J., and Peck, W. F., 1992, Phys. MacLaughlin, D. E., Bernal, O. O., and Lukefahr, H. Rev. Lett., 69, 2975. G., 1996, J. Phys.:Condens. Matter, 8, 9855. Tarucha, S., Honda, T., and Saku, T., 1995, Solid State Mahan, G. D., 1990, Many Particle Physics, (New Comm., 94, 413. York: Plenum). Tsui, D. C., St¨ormer, L. H., and Gossard, A. C., 1982, Majorana, E., 1937, Il Nuovo Cimento, 14 171. Phys. Rev. Lett., 48, 1559. Maple, M. B., Dickey, R. P., Herrmann, J., de Andrade, Tsvelik, A. M., and Wiegmann, P. B., 1985, J. Stat. M. C., Freeman, E. J., Gajewski, D. A., and Chau, Phys., 38, 125. R., 1996, J. Phys.:Condens. Matter, 8, 9773. Uemura, Y. J., Le, L. P., Luke, G. M., Sternlieb, B. J., Mathur, N. D., Grosche, F. M., Julian, S. R., Walker, Wu, W. D., Brewer, J. H., Riseman, T. M., Seaman, I. R., Freye, D. M., Haselwimmer, R. K. W., and C. L., Maple, M. B., Ishikawa, M., Hinks, D. G., Lonzarich, G. G., 1998, Nature, 394, 39. Jorgensen, J. D., Saito, G., and Yamochi, H., 1991, Milliken, F. P., Umbach, C. P., and Webb, R. A., 1996, Phys. Rev. Lett., 66, 2665. Solid State Comm., 97, 309. Varma, C. M., Littlewood, P. B., Schmittrink, S., Abra- Millis, A. J., 1993, Phys. Rev. B, 48, 7183. hams, E., Ruckenstain, A. E., 1989, Phys. Rev. Lett., Miranda, E., Dobrosavljevi´c, V., and Kotliar, G., 1996, 63, 1996. J. Phys.:Condens. Matter, 8, 9871. Voit, J., 1993, Phys. Rev. B, 47, 6740. Miranda, E., Dobrosavljevi´c, V., and Kotliar, G., 1997, Voit, J., 1998, to be published, cond-mat/9711064. Phys. Rev. Lett., 78, 290. Wheatley, J. C., 1970, Progress in low temperature Moriya, T., 1985, Spin Fluctuations in Itinerent Elec- physics, VI, 77, Ed. C. J. Gortor, North-Holland tron Magnets, (Berlin: Springer-Verlag). Publishing Co. Nozi`eres, P., 1974. J. Low Temp. Phys., 17, 31. Willett, R. L. 1997, Adv. Phys, 46, 447. Nozi`eres, P., and Blandin, A., 1980, J. Phys. (Paris), Wilson, K. G., 1975, Rev. Mod. Phys., 47, 773. 41, 193. Wzietek, P., Creuzet, F. Bourbonnais, C., Jerome, D. Ott, H. R., Rudigier, H., Fisk, Z., and Smith, J. L., Bechgaard, K., Batail, P., 1993, J. Phys. (Paris) I, 1983, Phys. Rev. Lett., 50, 1595. 3, 171. Pauli, W., 1927, Z. Phys., 41, 81. Yacoby, A., Stormer, H. L., Wingreen, N. S., Pfeiffer, Pfleiderer, C., McMullan, G. J., Julian, S. R., and Lon- L. N., Baldwin, K. W., and West, K. W., 1996, Phys. zarich, G. G., 1997, Phys. Rev. B, 55, 8330. Rev. Lett., 77, 4612. Pietrus, T., et al. 1995, Physica B, 206-207, 317. Zwick, F., Brown, S., Margaritondo, G., Merlic, C., Pines, D. and Nozi`eres, P., 1966, The Theory of Quan- Onellion, M., Voit, J., and Grioni, M., 1997, Phys. tum Liquids I,(NewYork:W.A.BenjaminInc.). Rev. Lett., 79, 3982. Ralph, D. C., Ludwig, A. W. W., von Delft, J., and Buhrman, R. A., 1994, Phys. Rev. Lett., 72, 1064. Reizer, M. Yu, 1989, Phys. Rev. B, 40, 11571. Reizer, M. Yu, 1991, Phys. Rev. B, 44, 5476. Schofield, A. J., 1997, Phys. Rev. B, 55, 5627. Shaked, H., Keane, P. M., Rodriguez, J. C., Owen, F. F., Hitterman, R. L., and Jorgensen, J. D,, 1994, Crystal Structures of the High-Tc Superconducting Copper-Oxides, (Amsterdam: Elsevier Science B. V.). Sengupta, A. M. and Georges, A., 1994, Phys. Rev. B, 49, 10020. Silin, V. P., 1957, Sov. Phys.—JETP, 6, 387; ibid 6, 985. Non-Fermi liquids A. J. Schofield 23

Andrew Schofield is a Royal Society University Re- search Fellow at the University of Cambridge and a Fellow of Gonville and Caius College. He completed his Ph. D. in 1993 at Cambridge where he applied gauge models to understand how strongly interact- ing electrons might give rise to the unusual normal- state properties of the cuprate superconductors. He then spent two years as a Research Associate at Rut- gers university, USA, where he also worked on the theory of heavy fermion materials. At present his research concerns the consequences of strong elec- tron correlations in a variety of systems, with a par- ticular interest in developing theories of non-Fermi liquid behaviour in the cuprates and in metals near quantum critical points.