CCMS Summer 2007 Lecture Series Fermi- and non-fermi Liquids Lecture 3: Fermi-liquid Theory

Similar documents
Landau s Fermi Liquid Theory

Landau-Fermi liquid theory

V. FERMI LIQUID THEORY

Condensed matter theory Lecture notes and problem sets 2012/2013

Identical Particles. Bosons and Fermions

Landau-Fermi liquid theory

Landau Theory of Fermi Liquids : Equilibrium Properties

PHZ 7427 SOLID STATE II: Electron-electron interaction and the Fermi-liquid theory

Strongly correlated systems in atomic and condensed matter physics. Lecture notes for Physics 284 by Eugene Demler Harvard University

Preface Introduction to the electron liquid

Nucleation in a Fermi liquid at negative pressure

ψ s a ˆn a s b ˆn b ψ Hint: Because the state is spherically symmetric the answer can depend only on the angle between the two directions.

Physics 127a: Class Notes

Fermi Liquid and BCS Phase Transition

Correlations between spin accumulation and degree of time-inverse breaking for electron gas in solid

In-class exercises. Day 1

J10M.1 - Rod on a Rail (M93M.2)

We already came across a form of indistinguishably in the canonical partition function: V N Q =

Lecture notes for QFT I (662)

E&M. 1 Capacitors. January 2009

ELECTRON-ELECTRON INTERACTIONS IN TWO-DIMENSIONAL SPIN-ORBIT COUPLED SYSTEMS

Luigi Paolasini

Donoghue, Golowich, Holstein Chapter 4, 6

P3317 HW from Lecture and Recitation 10

Physics Nov Bose-Einstein Gases

Atomic Structure. Chapter 8

LOCAL MOMENTS NEAR THE METAL-INSULATOR TRANSITION

1 Superfluidity and Bose Einstein Condensate

Minimal Update of Solid State Physics

The Dulong-Petit (1819) rule for molar heat capacities of crystalline matter c v, predicts the constant value

Lecture notes on many-body theory

Lecture 3 (Part 1) Physics 4213/5213

+ 1. which gives the expected number of Fermions in energy state ɛ. The expected number of Fermions in energy range ɛ to ɛ + dɛ is then dn = n s g s

The 3 dimensional Schrödinger Equation

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Physics Department Statistical Physics I Spring Term 2013

Phonon II Thermal Properties

The non-interacting Bose gas

Nonlinear screening and percolation transition in 2D electron liquid. Michael Fogler

AST1100 Lecture Notes

Introduction. Chapter The Purpose of Statistical Mechanics

Electromagnetism II. Instructor: Andrei Sirenko Spring 2013 Thursdays 1 pm 4 pm. Spring 2013, NJIT 1

Def.: Magnetism the property of a material to be attracted to (paramagnetic response) or repelled by (diamagnetic response) a magnetic field

Physics 127c: Statistical Mechanics. Weakly Interacting Fermi Gas. The Electron Gas

Ideas on non-fermi liquid metals and quantum criticality. T. Senthil (MIT).

III.H Zeroth Order Hydrodynamics

with a proper choice of the potential U(r). Clearly, we should include the potential of the ions, U ion (r):.

Lecture 11 - Phonons II - Thermal Prop. Continued

Landau Bogolubov Energy Spectrum of Superconductors

Phys 622 Problems Chapter 5

QUALIFYING EXAMINATION, Part 1. 2:00 PM 5:00 PM, Thursday September 3, 2009

Physics 4213/5213 Lecture 1

ɛ(k) = h2 k 2 2m, k F = (3π 2 n) 1/3

Statistical Mechanics

Quantum Theory of Matter

7.4. Why we have two different types of materials: conductors and insulators?

(a) Write down the total Hamiltonian of this system, including the spin degree of freedom of the electron, but neglecting spin-orbit interactions.

Basic Concepts and Tools in Statistical Physics

Loop corrections in Yukawa theory based on S-51

Thermodynamics, Gibbs Method and Statistical Physics of Electron Gases

Interacting Fermi Gases

Condensed Matter Physics Prof. G. Rangarajan Department of Physics Indian Institute of Technology, Madras

Optical Properties of Lattice Vibrations

Intermediate valence in Yb Intermetallic compounds

Multi-Electron Atoms II

First-Principles Calculation of Exchange Interactions

Physics Dec The Maxwell Velocity Distribution

Spin Superfluidity and Graphene in a Strong Magnetic Field

Brief review of Quantum Mechanics (QM)

Quantum Condensed Matter Physics Lecture 5

Physics 228 Today: April 22, 2012 Ch. 43 Nuclear Physics. Website: Sakai 01:750:228 or

Lecture 11: Periodic systems and Phonons

Magnets, 1D quantum system, and quantum Phase transitions

Ferromagnetism of electron gas. Abstract

SOLID STATE PHYSICS. Second Edition. John Wiley & Sons. J. R. Hook H. E. Hall. Department of Physics, University of Manchester

Intermission: Let s review the essentials of the Helium Atom

Lecture 6: Fluctuation-Dissipation theorem and introduction to systems of interest

STONY BROOK UNIVERSITY DEPARTMENT OF PHYSICS AND ASTRONOMY. Comprehensive Examination. Classical Mechanics. August 25, 2014

Phase Transitions in Condensed Matter Spontaneous Symmetry Breaking and Universality. Hans-Henning Klauss. Institut für Festkörperphysik TU Dresden

Lecture 4 Recap: normal metals and the clectron-phonon interaction *

The 4th Windsor Summer School on Condensed Matter Theory Quantum Transport and Dynamics in Nanostructures Great Park, Windsor, UK, August 6-18, 2007

Plan of the lectures

3. General properties of phase transitions and the Landau theory

Statistical. mechanics

Numerical study of localization in antidot lattices

On the Heisenberg and Schrödinger Pictures

PHY 396 K. Solutions for problem set #11. Problem 1: At the tree level, the σ ππ decay proceeds via the Feynman diagram

M01M.1 Massive Spring

The Ginzburg-Landau Theory

Physics 221A Fall 1996 Notes 13 Spins in Magnetic Fields

BCS-BEC Crossover. Hauptseminar: Physik der kalten Gase Robin Wanke

Topological insulator part I: Phenomena

Quantum Physics in the Nanoworld

Dr Victoria Martin, Spring Semester 2013

Fermi liquids and fractional statistics in one dimension

Renormalization Group Methods for the Nuclear Many-Body Problem

Statistical Mechanics Victor Naden Robinson vlnr500 3 rd Year MPhys 17/2/12 Lectured by Rex Godby

Physics 607 Exam 2. ( ) = 1, Γ( z +1) = zγ( z) x n e x2 dx = 1. e x2

Lecture Models for heavy-ion collisions (Part III): transport models. SS2016: Dynamical models for relativistic heavy-ion collisions

P3317 HW from Lecture 15 and Recitation 8

SECOND PUBLIC EXAMINATION. Honour School of Physics Part C: 4 Year Course. Honour School of Physics and Philosophy Part C C3: CONDENSED MATTER PHYSICS

Transcription:

CCMS Summer 2007 Lecture Series Fermi- and non-fermi Liquids Lecture 3: Fermi-liquid Theory Dmitrii L. Maslov maslov@phys.ufl.edu (Dated: July 22, 2007) 1

Notation 1 Here and thereafter, L1 stands for Lecture 1: Basic Notions of the Many-body Physics and L2 stands for Lecture 2: Examples of Interacting Fermi systems. I. GENERAL CONCEPTS All Fermi systems (metals, degenerate semiconductors, normal He 3, neutron stars, etc.) belong to the categories of either moderately or strongly interacting systems. For example, in metals r s in the range from 2 to 5. (There are only few exceptions of this rule; for example, bismuth, in which the large value of the background dielectric constant brings the value of r s to 0.3 and GaAs heterostructures in which the small value of the effective mass 0.07 of the bare mass leads to the higher value of the Fermi energy and thus to r s < 1). On the other hand, as we learned from the section on Wigner crystallization, the critical r s for Wigner crystallization is very high about 100 in 3D and about 40 in 2D. Thus almost all Fermi systems occurring in Nature are too strongly interacting to be described by the weak-coupling theory (Coulomb gas) but too weakly interacting to solidify. Landau put forward a hypothesis that an interacting Fermi system is qualitatively similar to the Fermi gas [6]. Although original Landau s formulation refers to a translationally invariant system of particles interacting via short-range forces, e.g., normal He 3, later on his arguments were extended to metals (which have only discrete symmetries) and to charged particles. Experiment gives a strong justification to this hypothesis. The specific heat of almost all fermionic systems (in solids, one need to subtract off the lattice contribution to get the one from electrons) scales linearly with temperature: C (T ) = γ T. In a free Fermi gas, γ = γ = (π 2 /3) ν F = (1/3) mk F. In a band model, when non-interacting electrons move in the presence of a periodic potential, one should use the appropriate value of the density of states at the Fermi level for a given lattice structure. In reality, the coefficient γ can differ significantly from the band value but the linearity of C(T ) in T is well-preserved. In those cases when one can change number density continuously (for example, by applying pressure to normal He 3 ), γ is found to vary. One is then tempted to assume that the interacting Fermi liquid is composed of some effective particles (quasi-particles) that behave as free fermions albeit their masses and other characteristics are different from the non-interacting values. 2

A. Quasi-particles The concept of quasi-particles is central to the Landau s theory of Fermi liquids. The ground state of a Fermi gas is a completely filled Fermi sphere. The spectrum of excited states can be classified in terms of how many fermions were promoted from states below the Fermi surface to the ones above. For example, the first excited state is the one with an electron above the Fermi sphere and the hole below. The energy of this state, measured from the ground state, is ξ p = p 2 /2m ε F. The net momentum of the system is p. The next state correspond to two fermions above the Fermi sphere, etc. If the net momentum of the system is p, then p = p 1 + p 2, where p 1 and p 2 are the momenta of individual electrons. We see that in a free system, any excited macroscopic state is a superposition of single-particle states. This is not so in an interacting system. Even if we promote only one particle to a state above the Fermi surface, the energy of this state would not be equal to p 2 /2m ξ p because the interaction will change the energy of all other fermions. However, Landau assumed that excited states with energies near the Fermi level, that is, weakly excited states, can be described as a superposition of elementary excitations which behave as free particles, although the original system may as well be a strongly interacting one. An example of such a behavior are familiar phonons in a solid. Suppose that we have a gas of sodium atoms (which are fermions) which essentially do not interact because of low density. The elementary excitations in an ideal gas simply coincide with real atoms. Now we condense the gas into a metal. Individual atoms are not free to move on their own. Instead, they can only participate in a collective oscillatory motion, which is a sound wave. For small frequencies, the sound wave can be thought of consisting of elementary quanta of free particles phonons. The spectrum of each phonon branch is ω i (q) = s i q, where s i is the speed of sound and the oscillatory energy is E = i d 3 q (2π) 3 ω in i, where n i is the number of excited phonons at given temperature. If the number of phonons is varied, so is the total energy δe = i d 3 q (2π) 3 ω iδn i. Quite similarly, Landau assumed that the variation of the total energy of a single-component 3

Fermi liquid (or single band metal) can be written as δe = (2π) 3 ε pδn p. (1.1) (For the sake of simplicity, I also assume that the system is isotropic, i.e., the energy and n depend only on the magnitude but not the direction of the momentum but the argument can be extended to anisotropic systems as well). In this formula, n p is the distribution function of quasi-particles which are elementary excitations of the interacting system. The variation of n p occurs when, e.g., a new particle is added to the system. However, these quasi-particles are not completely free and, therefore, unlike in the phonon analogy, the total energy is not equal to the sum of individual energies. However, the variation of the total energy δe is related to δn p via (1.1). In fact, this relation serves as a definition of the quasiparticle energy via the variational derivative of E : ε p = δe δn p. One more and crucial assumption is that the quasi-particles of an interacting Fermi system are fermions, which is not at all obvious. For example, regardless of the statistics of individual atoms which can be either fermions or bosons, phonons are always bosons. Landau s argument was that if quasi-particles were bosons they could accumulate without a restriction in every quantum state. That means that an excited state of a quantum system has a classical analog. Indeed, an excited state of many coupled oscillators is a classical sound wave. Fermions don t have macroscopic states so quasi-particles of a Fermi systems must be fermions (to be precise they should not be bosons; proposals for particles of a statistics intermediate between bosons and fermions anyons have been made). On quite general grounds, one can show that quasi-particles must have spin 1/2 regardless of (half-integer) spin of original particles [3],[4]. (Thus, quasi-particles of a system composed of fermions with spin S = 3/2 still have spin 1/2.) Therefore, the quasiparticle energy and the occupation number are the operators in the spin space represented by 2 2 matrices ˆε and ˆn. As long as one has a system of fermions with doubly (S z = ±1/2) degenerate states, the entropy on purely combinatorial grounds is S/vol = Tr (2π) 3 [ˆn p ln ˆn p (1 ˆn p ) ln (1 ˆn p )], which is the same expression in for a Fermi gas. The equlibrium occupation numbers are obtained by equating the variation of S at fixed volume to zero. As for a Fermi gas, this 4

gives ˆn = 1 exp ( ˆε µ T In contrast to the Fermi-gas, however, ˆε is a functional of ˆn itself. ) + 1, (1.2) If the system is not in the presence of the magnetic field and not ferromagnetic, ˆε αβ = εδ αβ, ˆn αβ = nδ αβ. In a general case, instead of (1.1) we have δe = which, for a spin-isotropic liquid, reduces to δe = 2 3 Trˆε (p) δˆn (p), (2π) 3 Trε (p) δn (p). (2π) The occupation number is normalized by the condition δn = d 3 p (2π) 3 Trδn (p) = 2 where N is the total number of real particles. 3 δn (p) = 0, (2π) For T=0, the chemical potential coincides with energy of the topmost state µ = ε (p F ) E F. Another important property (known as Luttinger theorem) is that the volume of the Fermi surface is not affected by the interaction. For an isotropic system, this means the Fermi momenta of free and interacting systems are the same. counting of states is not affected by the interaction, i.e., the relation N = 2 4πp3 F /3 (2π) 3 A simple argument is that the holds in both cases. A general proof of this statement is given in Ref. [ [3]]. B. Interaction of quasi-particles Phonons in a solid do not interact only in the first (harmonic) approximation. Anharmonism results in the phonon-phonon interaction. However, the interaction is weak at low 5

energies not really because the coupling constant is weak but rather because the scattering rate of phonons on each other is proportional to a high power of their frequency: τ 1 ph-ph ω5. As a result, at small ω phonons are almost free quasi-particles. Something similar happens with the fermions. The nominal interaction may as well be strong. However, because of the Pauli principle, the scattering rate is proportional to (ε ε F ) 2 and weakly excited states interact only weakly. In the Landau theory, the interaction between quasi-particles is introduced via a phenomenological interaction function defined by the proportionality coefficient (more precisely, a kernel) between the variation of the occupation number and the corresponding variation in the quasi-particle spectrum δε αβ = (2π) 3 f αγ,βδ ( p, p ) δn γδ ( p ). Function f αγ,βδ ( p, p ) describes the interaction between the quasi-particles of momenta p and p (notice that these are both initial states of the of the scattering process). Spin indices α and β correspond to the state of momentum p whereas indices γ and δ correspond to momentum p. In a matrix form, δˆε = Tr (2π) 3 ˆf ( p, p ) δˆn ( p ), (1.3) where Tr denotes trace over spin indices γ and δ. For a spin-isotropic FL, when δˆε = δ αβ δε and δˆn = δ αβ δn external spin indices (α and β) can also be traced out and Eq.(1.3) reduces to where δε = (2π) 3 f ( p, p ) δn ( p ), f ( p, p ) 1 2 TrTr ˆf ( p, p ). For small deviations from the equilibrium, δn ( p ) is peaked near the Fermi surface. Function ˆf ( p, p ) can be then estimated directly on the Fermi surface, i.e., for p = p = p F. Then ˆf depends only on the angle between p and p. The spin dependence of ˆf can be established on quite general grounds. In a spin-isotropic FL, ˆf can depend only on the scalar product of spin operators but on the products of the individual spin operators with some other vectors. Thus the most general form of ˆf for a spin-isotropic system is ν ˆf ( p, p ) = F s (θ) Î + F a (θ) ˆσ ˆσ, 6

where Î is the unity matrix, ˆσ is the vector of three Pauli matrices, θ is the angle between p and p, and the density of states was introduced just to make functions F (θ) and G (θ) dimensionless. (Star in ν means that we have used a renormalized value of the effective mass so that ν = m k F /π 2, but this is again just a matter of convenience.) Explicitly, ν f αγ,βδ ( p, p ) = F s (θ) δ αβ δ γδ + F a (θ) ˆσ αβ ˆσ γδ. (1.4) In general, the interaction function is not known. However, if the interaction is weak, one can relate ˆf to the pair interaction potential. The microscopic theory of a Fermi liquid [3] shows that the Landau interaction function is related to the interaction vertex Γ αβ,γδ ( pε, p ε p + q, ε + ω, p q, ε ω), which describes scattering of two fermions in initials states with momenta, energies, and spin projections p, ε, α and p, ε, β, respectively into the final states p + q, ε + ω, γ and p, ε + ω, δ via f αβ,γδ ( p, p ) = Z 2 lim Γ αβ,γδ ( pε, p ε p + q, ε + ω, p q, ε ω), q/ω 0 where Z is the renormalization factor of the Green s function. Example 2 At first order in the interaction, Γ is represented by two diagrams, one of which is obtained from the other by a permutation of outgoing lines. If the interaction conserves spin, Γ αβ,γδ ( pε, p ε p + q, ε + ω, p q, ε ω) = δ αγ δ βδ U ( q) δ αδ δ βγ U ( p p + q). Using an SU(2) identity δ αδ δ βγ = 1 2 (δ αγδ βδ + ˆσ αγ ˆσ βδ ), this expression for Γ reduces to Γ αβ,γδ ( pε, p ε p + q, ε + ω, p q, ε ω) = δ αγ δ βδ [U ( q) 1 ] 2 U ( p p + q) 1 2ˆσ αγ ˆσ βδ U ( p p + q). Exchanging indices β and γ and taking the limit q 0 (at this order the vertex does not depend on the energy), we obtain for ˆf f αγ,βδ ( p, p ) = δ αβ δ γδ [U (0) 1 ] 2 U ( p p ) 1 2ˆσ αβ ˆσ γδ U ( p p ). 7

[Because this relation holds only at first order of the perturbation theory, Z still equals to 1.] On the Fermi surface, p p = 2p F sin θ/2. Comparing this expression with Eq.(1.4), we find that F s (θ) = ν [U 1 (0) 1 ] 2 U (2p F sin θ/2) F a (θ) = 1 2 U (2p F sin θ/2). (1.5) Notice that a repulsive interaction (U > 0) corresponds to the attraction in the spin-exchange channel (F a < 0) this is the origin of the ferromagnetic instability. C. General strategy of the Fermi-liquid theory One may wonder what one can achieve introducing unknown phenomenological quantities, such as the interaction function or its charge and spin components. FL theory allows one to express general thermodynamic characteristic of a liquid (effective mass, compressibility, spin susceptibility, etc.) via angular averages of Landau functions F s (θ) and F a (θ). Some quantities depend on the same averages and thus one express such quantities via each other. The relationship between such quantities can be checked by comparison with the experiment. D. Effective mass As a first application of the Landau theory, let s consider the effective mass. In a Galileaninvariant system, the momentum per unit volume coincides with the flux of mass: Tr d 3 p (2π) 3 pˆn = Tr m ˆε (2π) 3 p ˆn, where m is the bare mass. Taking the variation of both sides of this equation, we obtain d 3 p Tr (2π) 3 pδˆn = Tr d 3 [ ] p δˆε (2π) 3 m ˆε ˆn + p p δˆn. (1.6) Now, δˆε p = Tr For a spin-isotropic liquid, Eq.(1.4) reduces to d 3 p (2π) 3 pδn = (2π) 3 m ε p δn + ˆf ( p, p ) (2π) 3 δˆn ( p ). p d 3 [ p (2π) 3 m f ( p, p ] ) (2π) 3 δn ( p ) p 8 n ( p).

In the second term re-label the variables p p, p p, use the fact that f ( p, p ) = f ( p, p) and integrate by parts d 3 p (2π) 3 pδn = (2π) 3 m ε p δn d 3 p (2π) 3 (2π) 3 f ( p, p ) n ( p ) δn ( p). p Because this equation should be satisfied for an arbitrary variation δn ( p), p m = ε p (2π) 3 f ( p, p ) n ( p ). p Near the Fermi surface, the quasi-particle energy can be always written as so that ε (p) = v F (p p F ) ε p = v F ˆp = p F m ˆp, where ˆp is the unit vector in the direction of p, and thus Now Near the Fermi surface, and or p m = p F ˆp m (2π) 3 f ( p, p ) n ( p ) p n ( p ) = ε ( p ) n p p ε. n ε = δ (ε E F ) p m = p F ˆp m + 1 dω 2 ν F 4π f ( p, p ) p F ˆp m p m = p F ˆp m + 1 dω 2 ν F 4π f ( p, p ) p F ˆp m = p F ˆp m + 1 p F dω 2 π 2 4π f ( p, p ) p F ˆp. Setting p = p F and multiplying both sides of the equation by ˆp, we obtain or 1 m = 1 m + 1 p F 2 π 2 1 m = 1 m p F 2π 2 dω f (θ) cos θ 4π dω f (θ) cos θ. (1.7) 4π This is the Landau s formula for the effective mass. Noticing that f (θ) = (1/2)TrTr ˆf (θ) = (2/ν F ) F (θ) = (2π 2 /m p F ), the last equation can be reduced to m m = 1 + dω 4π F (θ) cos θ 1 + F 1. 9

Although we do not know the explicit form of f (θ), some useful conclusions can be made already from the most general form [Eq.(1.7)]. If f (θ) cos θ is negative, 1/m > 1/m m < m; conversely, if f (θ) cos θ is positive, m > m. For the weak short-range interaction, we know that m > m, whereas for a weak long-range (Coulomb) interaction m < m. Now we see that both of these cases are described by the Landau s formula. Recalling the weakcoupling form of F (θ) [Eq.(1.5)], we find dω dω F 1 = [U 4π F (θ) cos θ = (0) 12 ( 4π U 2p F sin θ )] cos θ 2 ( dω 1 = 4π 2 U 2p F sin θ ) cos θ. 2 If U is repulsive and picked at small θ where cos θ is positive, F 1 < 0 and m < m. This case corresponds to a screened Coulomb potential. If U is repulsive and depends on θ only weakly (short-range interaction), the angular integral is dominated by values of θ close to π, where cos θ < 0. Then m > m. In general, forward scattering reduces the effective mass, whereas large-angle scattering enhances it. E. Spin susceptibility 1. Free electrons We start with free electrons. An electron with spin s has a magnetic moment µ = 2µ B s, where µ B = e h/2mc is the Bohr magneton. Zeeman splitting of energy levels is E = E E = µ B H ( µ B H) = 2µ B H. The number of spin-up and -down electrons is found as an integral over the density of states n, = 1 2 EF ±µ B H 0 dεν (ε). The Fermi energy now also depends on the magnetic field but the dependence is only quadratic (why?), whereas the Zeeman terms are linear in H. For weak fields, one can neglect the field-dependence of E F. Total magnetic moment per unit volume magnetization is found by expanding in H M = µ B (n n ) = µ [ EF +µ B B H ] EF µ B H dεν (ε) dεν (ε) 2 0 0 = µ [ EF ] EF B dεν (ε) + µ B Hν F dεν (ε) + µ B Hν F 2 0 0 = µ 2 BHν F. 10

Spin susceptibility χ = M H = µ2 Bν F (1.8) is finite and positive, which means that spins are oriented along the external magnetic field. Notice that if, for some reason the bare g- factor of electrons is not equal to 2, then Eq.(1.8) changes to χ = g 2 µ2 Bν F (1.9) 2. Fermi liquid In a Fermi liquid, the energy of a quasi-particle in a magnetic field is changed not only due to the Zeeman splitting (as in the Fermi gas) but also because the occupation number is changed. This effect brings in an additional term in the energy as the energy is related to the occupation number. This is expressed by following equation δˆε = µ B H ˆσ + Tr (2π) 3 ˆf ( p, p ) δn ( p ). The first term is just the Zeeman splitting, the second one comes from the interaction. Now, δˆn = ˆn ε δˆε, and we have an equation for δˆε δˆε = µ B H ˆσ + Tr ˆf (2π) 3 ( p, p ) ˆn ε δˆε ( p ). Replacing ˆn ε by the delta-function and setting p = p F δˆε (p F ˆp) = µ B H ˆσ Tr ν F Let s try a solution in the following form dω 4π ˆf (p F ˆp, p F ˆp ) δˆε (p F ˆp ). (1.10) δˆε = µ B gh ˆσ, (1.11) 2 where g has the meaning of an effective g factor. For free electrons, g = 2. Recalling that ν F ˆf = F (θ) Î + G (θ) ˆσ ˆσ and substituting (1.11) into (1.11), we obtain µ B dω 2 gh ˆσ = µ BH ˆσ Tr 4π [ F s (θ) Î + F a (θ) ˆσ ˆσ ] ( µ ) B gh ˆσ. 2 11

The term containing F in the integral vanishes because Pauli matrices are traceless: Trσ = 0. The term containing G is re-arranged using the identity Tr(ˆσ ˆσ ) ˆσ = 2ˆσ upon which we get where Therefore, g 2 = 1 g 2 F a 0, F0 a = g = dω 4π F a (θ). 2. 1 + F0 a Substituting this result into Eq.(1.9) and replacing ν F obtain the spin susceptibility of a Fermi liquid by its renormalized value ν F, we χ = µ2 Bν F 1 + F a 0 = χ 1 + F 1 s, 1 + F0 a where χ is the susceptibility of free fermions. Notice that χ differs from χ both because the effective mass is renormalized and because the g-factor differs from 2. When F0 a = 1, the g-factor and susceptibility diverge signaling a ferromagnetic instability. However, even if G 0 = 0, χ is still renormalized in proportion to the effective mass. If m diverges (which is a signature of a metal-insulator transition of Mott type), χ diverges as well. Recall that [Eq.(1.5)] in the weak-coupling limit F a (θ) = 1U (2p 2 F sin θ/2). Thus, for repulsive interactions g > 2 which signals a ferromagnetic tendency. This is another manifestation of general principle that repulsively interacting fermions tend to have spin aligned to minimize the energy of repulsion. For normal He 3, G 0 2/3. F. Zero sound All gases and liquids support sound waves. Even ideal gases have finite compressibilities and therefor finite sound velocities s = P ρ. For example, in an ideal Boltzmann gas P = nt = ρt/m and s = T m = 1 3 v T, 12

where v T = 3T/m is the rms thermal velocity. In an ideal Fermi gas, P = (2/3)E where E = (2/5) ne F is the ground state energy and s = 1 3 v F. It seems that interactions are not essential for sound propagation. This conclusion is not true. In a sound wave, all thermodynamic characteristics density n (r, t), pressure P (r, t), temperature T (r, t) (if the experiment is performed under adiabatic conditions), etc. vary in space and time in sync with the sound wave. Thus we are dealing with a non-equilibrium situation. To describe a non-equilibrium situation by a set of local and time-dependent quantities, one needs to maintain local and temporal equilibrium. When the sound wave arrives to an initially unperturbed region, this region is driven away from its equilibrium state. Collisions between molecules (or fermions in a Fermi gas) has to be frequent enough to establish a new equilibrium state before the sound wave leaves the region. For sound propagation, the characteristic spatial scale is the wavelength λ and characteristic time is the period 2π/ω. If the mean free path and time are l and τ, respectively, the conditions of establishing local and temporal equilibrium are λ l 2π/ω τ. Therefore, collisions (interactions) are essential to ensure that the sound propagation occurs under quasi-equilibrium conditions. For classical gases and liquids these conditions are satisfied for all not extremely high frequencies. For example, the mean free path of molecules in air at P = 1 atm and T = 300 K is about l 10 5 cm and s 300 m/s=3 10 4 cm/s. Condition λ l takes the form ω s l = 3 104 cm/s 10 5 cm = 3 109 s 1. In a liquid, l is of order of the inter-molecule separation few 10 7 cm. As long as λ 10 7 cm, equilibrium is maintained. In a Fermi liquid, the situation is different because l and τ increase as temperature goes down. Local equilibrium is maintained as long as τ (T ) ω 1. Now, τ (T ) = 1/AT 2 so at fixed temperature sound of frequency ω τ 1 (T ) = AT 2 cannot proceed in a quasiequilibrium manner. 13

What happens when we start at small frequency and then increase it? As long as τ (T ) ω 1, a Fermi liquid supports normal sound (in the FL theory it is called first sound). The velocity of this sound is linked to the sound velocity in the ideal Fermi gas but differs from because of renormalization s 1 = v F (1 + F0 s ) 1/2 (1 + F1 s ) 1/2, 3 where F s 0 = dω 4π F s (θ). The first sound corresponds to the oscillations of fermion number density in sync with the wave. Because the number density fixes the radius of the Fermi sphere, p F oscillates in time and space. As the product ωτ increases, sound absorption does so too, at when ωτ becomes impossible. However, as ωτ increases further and becomes 1, another type of sound wave emerges. These waves are called zero sound. These waves can propagate even at T = 0, when τ = as they do not require local equilibrium. Another difference between zero- and first sound is that in the first sound-wave the shape of the Fermi surface remains spherical whereas its radius changes. In the zero-sound wave, the shape of the Fermi surface may be changed without changing its volume. In a simplest zero-sound wave [4], the occupation number is δn = δ (ε E F ) ν (ˆk) e i( k r ωt) ν (ˆk) cos θ = C s 0 /v F cos θ where s 0 > v F is the zero-sound velocity. The Fermi surface is an egg-shaped spheroid, with the narrow end pointing in the direction of the wave propagation. The condition s 0 > v F, always satisfied for zero sound, means absence of Landau damping. As well as s 1, the zero-sound velocity s 0 depends on Landau parameters F s 0 and F s 1, although the functional dependence is more complicated. At the same time, F s 1 can be extracted from the effective mass (measured from the specific heat) and F s 0 is extracted from the compressibility. Having two experimentally determined parameters F s 0 and F s 1, one can substitute them into theoretical expressions for s 1 and s 0 and compare them with measured sound velocities. The agreement (see the plot) is quite satisfactory. This comparison provides a quantitative check for the Fermi-liquid theory. H. J. Schulz, Fermi liquids and non-fermi liquids, in: Proceedings of Les Houches Summer School LXI, edited by E. Akkermans et al., Elsevier (Amsterdam), 1995, p. 553. Available 14

at xxx.lanl.gov/abs/comd-mat/9503150. [1] D. Yue, L. I. Glazman and K. A. Matveev, Phys. Rev. B 49, 1966 (1994). [2] Mahan, Many-body physics [3] A. A. Abrikosov, L. P. Gorkov, and I. E. Dzyaloshinski, Methods of quantum field theory in statistical physics, (Dover Publications, New York, 1963). [4] E.M. Lifshitz and L.P. Pitaevskii, Statistical Physics II, (Course of Theoretical Physics, v. IX), Pergamon Press, Oxford, 1980. [5] D. Pines and P. Nozieres, The Theory of Quantum Liquids, Benjamin, New York, 1969. [6] L. D. Landau, Zh. Eksp. Teor. Fiz. 30, 1058 (1956) [Sov. Phys. JETP 3, 920 (1957)]. Reprinted in: D. Pines, The Many-Body Problem, Benjamin, Reading, MA. [7] B. L. Altshuler and A. G. Aronov, in Electron-electron interactions in disordered conductors, edited by A. L. Efros and M. Pollak (Elsevier, 1985), p. 1. [8] G. Zala, B. N. Narozhny, and I. L. Aleiner, Phys. Rev. B 64, 214204 (2001). [9] E. Abrahams, S. V. Kravchenko, and M. P. Sarachik, Rev. Mod. Phys. 73, 251 (2001). [10] B. L. Altshuler, D. L. Maslov, and V. M. Pudalov, Physica E 9, 209 (2001). 15