arxiv:cond-mat/ v2 [cond-mat.str-el] 25 Jun 2003

Similar documents
Tuning order in cuprate superconductors

Spin-orbital separation in the quasi-one-dimensional Mott insulator Sr 2 CuO 3 Splitting the electron

Strongly Correlated Systems:

The Hubbard model in cold atoms and in the high-tc cuprates

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

Landau s Fermi Liquid Theory

Quantum phase transitions in Mott insulators and d-wave superconductors

5. Superconductivity. R(T) = 0 for T < T c, R(T) = R 0 +at 2 +bt 5, B = H+4πM = 0,

A FERMI SEA OF HEAVY ELECTRONS (A KONDO LATTICE) IS NEVER A FERMI LIQUID

Landau Bogolubov Energy Spectrum of Superconductors

Can superconductivity emerge out of a non Fermi liquid.

Talk online at

Quantum phase transitions

Origin of the anomalous low temperature upturn in resistivity in the electron-doped cuprates.

Emergent Frontiers in Quantum Materials:

arxiv:cond-mat/ v1 8 May 1997

Dual vortex theory of doped antiferromagnets

arxiv:cond-mat/ v1 8 Mar 1995

Quantum dynamics in many body systems

Quantum spin liquids and the Mott transition. T. Senthil (MIT)

Quantum Phase Transitions

Superconducting fluctuations, interactions and disorder : a subtle alchemy

arxiv:cond-mat/ v1 [cond-mat.supr-con] 28 May 2003

Quantum phase transitions and the Luttinger theorem.

Classifying two-dimensional superfluids: why there is more to cuprate superconductivity than the condensation of charge -2e Cooper pairs

Landau-Fermi liquid theory

Landau-Fermi liquid theory

Quantum Criticality and Black Holes

arxiv:cond-mat/ v1 [cond-mat.str-el] 30 Dec 1997

Signatures of the precursor superconductivity above T c

Order and quantum phase transitions in the cuprate superconductors

Superfluidity. v s. E. V. Thuneberg Department of Physical Sciences, P.O.Box 3000, FIN University of Oulu, Finland (Dated: June 8, 2012)

arxiv: v2 [cond-mat.supr-con] 7 Jun 2015

Superconductivity and Quantum Coherence

Photoemission Studies of Strongly Correlated Systems

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

Angle-Resolved Two-Photon Photoemission of Mott Insulator

Quantum Oscillations in underdoped cuprate superconductors

Identical Particles. Bosons and Fermions

Investigating the mechanism of High Temperature Superconductivity by Oxygen Isotope Substitution. Eran Amit. Amit Keren

Helium-3, Phase diagram High temperatures the polycritical point. Logarithmic temperature scale

High temperature superconductivity

Metals: the Drude and Sommerfeld models p. 1 Introduction p. 1 What do we know about metals? p. 1 The Drude model p. 2 Assumptions p.

On the Higgs mechanism in the theory of

Magnetism and Superconductivity in Decorated Lattices

Electron Doped Cuprates

A BCS Bose-Einstein crossover theory and its application to the cuprates

Theory of the Nernst effect near the superfluid-insulator transition

R measurements (resistivity, magnetoresistance, Hall). Makariy A. Tanatar

Small and large Fermi surfaces in metals with local moments

Strongly correlated Cooper pair insulators and superfluids

ELECTRON-HOLE ASYMMETRY IS THE KEY TO SUPERCONDUCTIVITY

ANTIFERROMAGNETIC EXCHANGE AND SPIN-FLUCTUATION PAIRING IN CUPRATES

Talk online: sachdev.physics.harvard.edu

Bardeen Bardeen, Cooper Cooper and Schrieffer and Schrieffer 1957

arxiv: v3 [cond-mat.str-el] 28 Aug 2013

Anomalous quantum criticality in the electron-doped cuprates

V, I, R measurements: how to generate and measure quantities and then how to get data (resistivity, magnetoresistance, Hall). Makariy A.

Quantum theory of vortices in d-wave superconductors

arxiv:cond-mat/ v1 [cond-mat.str-el] 21 Sep 2004

Purely electronic transport in dirty boson insulators

Quantum Oscillations, Magnetotransport and the Fermi Surface of cuprates Cyril PROUST

General relativity and the cuprates

ɛ(k) = h2 k 2 2m, k F = (3π 2 n) 1/3

Tunneling Spectroscopy of PCCO

arxiv: v1 [cond-mat.other] 11 Sep 2008

Quantum bosons for holographic superconductors

Origin of anomalous low-temperature downturns in the thermal conductivity of cuprates

Chapter 2 Superconducting Gap Structure and Magnetic Penetration Depth

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

Lecture 6. Fermion Pairing. WS2010/11: Introduction to Nuclear and Particle Physics

Superconductivity by kinetic energy saving?

arxiv:cond-mat/ v1 16 Jun 1993

V. FERMI LIQUID THEORY

arxiv:cond-mat/ v1 [cond-mat.supr-con] 26 Jan 2007

Superconductivity Induced Transparency

Citation PHYSICAL REVIEW LETTERS (2000), 85( RightCopyright 2000 American Physical So

requires going beyond BCS theory to include inelastic scattering In conventional superconductors we use Eliashberg theory to include the electron-

Superconductivity and Superfluidity

Superconducting Stripes

Holographic superconductors

Dynamics of fluctuations in high temperature superconductors far from equilibrium. L. Perfetti, Laboratoire des Solides Irradiés, Ecole Polytechnique

Condensed matter theory Lecture notes and problem sets 2012/2013

Splitting of a Cooper pair by a pair of Majorana bound states

Relativistic magnetotransport in graphene

Landau quantization, Localization, and Insulator-quantum. Hall Transition at Low Magnetic Fields

Some open questions from the KIAS Workshop on Emergent Quantum Phases in Strongly Correlated Electronic Systems, Seoul, Korea, October 2005.

From BEC to BCS. Molecular BECs and Fermionic Condensates of Cooper Pairs. Preseminar Extreme Matter Institute EMMI. and

Quantum oscillations & black hole ringing

Supplementary Figures

Neutron scattering from quantum materials

Contents Preface Physical Constants, Units, Mathematical Signs and Symbols Introduction Kinetic Theory and the Boltzmann Equation

Pseudogap and Conduction Dimensionalities in High-T c. Superconductors arxiv:cond-mat/ v1 [cond-mat.supr-con] 7 Mar 2002.

Topological Kondo Insulators!

Quantum matter & black hole ringing

Universal Features of the Mott-Metal Crossover in the Hole Doped J = 1/2 Insulator Sr 2 IrO 4

arxiv:cond-mat/ v1 4 Aug 2003

A DCA Study of the High Energy Kink Structure in the Hubbard Model Spectra

Lattice modulation experiments with fermions in optical lattices and more

arxiv:nucl-th/ v1 14 Jan 2002

Transcription:

Magnetic field-induced Landau Fermi Liquid in high-t c metals M.Ya. Amusia a,b, V.R. Shaginyan a,c 1 arxiv:cond-mat/0304432v2 [cond-mat.str-el] 25 Jun 2003 a The Racah Institute of Physics, the Hebrew University, Jerusalem 91904, Israel; b A.F. Ioffe Physical-Technical Institute, 194021 St. Petersburg, Russia; c Petersburg Nuclear Physics Institute, Gatchina, 188300, Russia Abstract We consider the behavior of strongly correlated electron liquid in high-temperature superconductors within the framework of the fermion condensation model. We show that at low temperatures the normal state recovered by the application of a magnetic field larger than the critical field can be viewed as the Landau Fermi liquid induced by the magnetic field. In this state, the Wiedemann-Franz law and the Korringa law are held and the elementary excitations are the Landau Fermi Liquid quasiparticles. Contrary to what might be expected from the Landau theory, the effective mass of quasiparticles depends on the magnetic field. The recent experimental verifications of the Wiedemann-Franz law in heavily hole-overdoped, overdoped and optimally doped cuprates and the verification of the Korringa law in the electron-doped copper-oxide superconductor strongly support the existence of fermion condensate in high-t c metals. PACS: 74.25.Fy; 74.72.-h; 74.20.-z Keywords: High temperature superconductivity; Wiedemann-Franz law; Korringa law 1 E mail: vrshag@thd.pnpi.spb.ru 1

The Landau Fermi Liquid (LFL) theory belongs to the most important achievements in physics [1]. The LFL theory has revealed that the low-energy elementary excitations of a Fermi liquid look like the spectrum of an ideal Fermi gas. These excitations are described in terms of quasiparticles with an effective mass M, charge e and spin 1/2. The quasiparticles define the major part of the low-temperature properties of Fermi liquids. Note that this powerful theory gives theoretical grounds for the theory of conventional superconductivity. On the other hand, despite outstanding results obtained in the condensed matter physics, the understanding of the rich and striking behavior of the high-temperature superconductors remains, as years before, among the main problems of the condensed matter physics. There is also a fundamental question about whether or not the properties of this electron liquid in high-t c metals can be understood within the framework of the LFL theory. It was reported recently that in the normal state obtained by applying a magnetic field greater than the upper critical filed B c, in a hole-doped cuprates at overdoped concentration (Tl 2 Ba 2 CuO 6+δ ) [2] and at optimal doping concentration (Bi 2 Sr 2 CuO 6+δ ) [3], there are no any sizable violations of the Wiedemann-Franz (WF) law. In the electron-doped copper oxide superconductor Pr 0.91 LaCe 0.09 Cu0 4 y (T c =24 K) when superconductivity is removed by a strong magnetic field, it was found that the spin-lattice relaxation rate 1/T 1 follows the T 1 T = constant relation, known as Korringa law [5], down to temperature of T = 0.2 K [4]. At elevated temperatures and applied magnetic fields of 15.3 T perpendicular to the CuO 2 plane, 1/T 1 T as a function of T is a constant below T = 55 K. At 300 K > T > 50 K, 1/T 1 T decreases with increasing T [4]. Recent measurements for strongly overdoped non-superconducting La 1.7 Sr 0.3 CuO 4 have shown that the resistivity ρ exhibits T 2 behavior, ρ = ρ 0 + ρ with ρ = AT 2, and the WF law is verified to hold perfectly [6]. Since the validity of the WF law and of the Korringa law are a robust signature of LFL, these experimental facts demonstrate that the observed elementary excitations cannot be distinguished from the Landau quasiparticles. This imposes strong constraints for models describing the hole-doped and electron-doped high-temperature superconductors. For example, in the cases of a Luttinger liquid [7], spin-charge separation (see e.g. [8]), and in some solutions of t J model [9] a violation of the WF law was predicted. In this Letter, we consider the behavior of strongly correlated electron liquid in high-temperature metals within the framework of the fermion condensation model [10, 11, 12]. We show that at temperatures T 0 the normal state recovered by the application of a magnetic field larger than the critical field B c can be viewed as LFL induced by the magnetic field. In this state, the WF law and the Korringa law are held and the elementary excitations are LFL quasiparticles. We show that contrary to what might be expected from the Landau theory, the effective mass of quasiparticles depends on the magnetic field. At T 0, the ground state energy E gs [κ(p), n(p)] of an electron liquid is a functional of the order parameter of the superconducting state κ(p) and of the quasiparticle occupation numbers n(p). Here we assume that the electron system is two-dimensional, while all results can be transported to the case of three-dimensional system. This energy is determined by the known equation of the weak-coupling theory of superconductivity, see e.g. [13] E gs = E[n(p)] + λ 0 V (p 1,p 2 )κ(p 1 )κ (p 2 ) dp 1dp 2 (2π) 4. (1) Here E[n(p)] is the ground-state energy of a normal Fermi liquid, n(p) = v 2 (p) and κ(p) = 2

v(p) 1 v 2 (p). It is assumed that the pairing interaction λ 0 V (p 1,p 2 ) produced, for instance, by electron-phonon interaction in high-t c metals is weak. Minimizing E gs with respect to κ(p) we obtain the equation connecting the single-particle energy ε(p) to the superconducting gap (p), ε(p) µ = (p) 1 2v2 (p), (2) 2κ(p) where the µ is the chemical potential, and single-particle energy ε(p) is determined by the Landau equation ε(p) = δe[n(p)]. (3) δn(p) Note that E[n(p)], ε[n(p)], and the Landau amplitude F L (p,p 1 ) = δe 2 [n(p)]/δn(p)δ(p 1 ) implicitly depend on the density x which defines the strength of F L. The equation for the superconducting gap (p) is of the form (p) = λ 0 V (p,p 1 )κ(p 1 ) dp 1 4π 2 = 1 λ 0 V (p,p 1 ) 2 (p 1 ) (ε(p 1 ) µ) 2 + 2 (p 1 ) dp 1 4π 2. (4) If λ 0 0, than the maximum value of the superconducting gap 1 0, and Eq. (2) reduces to the equation [10, 12, 14] ε(p) µ = 0, if 0 < n(p) < 1; p i p p f. (5) A new state of electron liquid with the fermion condensate (FC) [10, 15] is defined by Eq. (5) and characterized by a flat part of the spectrum in the (p i p f ) region. Apparently, the momenta p i and p f have to satisfy p i < p F < p f, where p F is the Fermi momentum. When the amplitude F L (p = p F, p 1 = p F ) as a function of the density x becomes sufficiently small, the flat part vanishes, Eq. (5) has the only trivial solution ε(p = p F ) = µ, and the quasiparticle occupation numbers are given by the step function, n(p) = θ(p F p) [10, 11]. Note, that a formation of the flat part of the spectrum has been recently confirmed in Ref. [16, 17]. Now we can study the relationships between the state defined by Eq. (5) and the superconductivity. At T 0, Eq. (5) defines a particular state of a Fermi liquid with FC, for which the modulus of the order parameter κ(p) has finite values in the (p i p f ) region, whereas 1 0 unlike the conventional theory of superconductivity, which demands that 1 vanishes simultaneously with the order parameter κ(p). Such a situation can take place because the order parameter κ(p), as it follows from Eqs. (3) and (5), is determined by the Landau amplitude F L, while 1 is given by Eq. (4) and tends to zero at λ 0 0. This state can be considered as superconducting, with an infinitely small value of 1, so that the entropy of this state is equal to zero. It is obvious that this state being driven by the quantum phase transition disappears at T > 0 [12]. Any quantum phase transition, which takes place at temperature T = 0, is determined by a control parameter other than temperature, for example, by pressure, by magnetic field, or by the density of mobile charge carriers x. The quantum phase transition occurs at a quantum critical point. At some density x x FC, when the Landau amplitude F L becomes sufficiently weak, and p i p F p f, Eq. (5) 3

determines the critical density x FC at which the fermion condensation quantum phase transition (FCQPT) takes place leading to the formation of FC [10, 12]. It follows from Eq. (5) that the system becomes divided into two quasiparticle subsystems: the first subsystem in the (p i p f ) range is characterized by the quasiparticles with the effective mass MFC 1/ 1, while the second one is occupied by quasiparticles with finite mass ML and momenta p < p i. The density of states near the Fermi level tends to infinity, N(0) MFC 1/ 1 [12]. If λ 0 0, then 1 becomes finite. It is seen from Eq. (4) that the superconducting gap depends on the single-particle spectrum ε(p). On the other hand, it follows from Eq. (2) that ε(p) depends on (p) provided that at 1 0 the spectrum is given by Eq. (5). Let us assume that λ 0 is small so that the particle-particle interaction λ 0 V (p,p 1 ) can only lead to a small perturbation of the order parameter κ(p) determined by Eq. (5). Upon differentiation both parts of Eq. (2) with respect to the momentum p, we obtain that the effective mass M FC = dε(p)/dp p=pf becomes finite [12] M FC p F p f p i 2 1. (6) It is clear from Eq. (6) that the effective mass and the density of states N(0) M FC 1/ 1 are finite. As a result, we are led to the conclusion that in contrast to the conventional theory of superconductivity the single-particle spectrum ε(p) strongly depends on the superconducting gap and we have to solve Eqs. (3) and (4) in a self-consistent way. Since the particle-particle interaction is small, the order parameter κ(p) is governed mainly by FC and we can solve Eq. (4) analytically taking the Bardeen-Cooper-Schrieffer approximation for the particle-particle interaction: λ 0 V (p,p 1 ) = λ 0 if ε(p) µ ω D, i.e. the interaction is zero outside this region, with ω D being the characteristic phonon energy. As a result, the maximum value of the superconducting gap is given by [14] 1 (0) λ 0p F (p f p F ) 2π ln ( 1 + 2 ) 2βε F p f p F p F ln ( 1 + 2 ). (7) Here, the Fermi energy ε F = p 2 F/2M L, and the dimensionless coupling constant β is given by the relation β = λ 0 M L /2π. Taking the usual values of β as β 0.3, and assuming (p f p F )/p F 0.2, we get from Eq. (7) a large value of 1 (0) 0.1ε F, while for normal metals one has 1 (0) 10 3 ε F. Now we determine the energy scale E 0 which defines the region occupied by quasiparticles with the effective mass M FC E 0 = ε(p f ) ε(p i ) 2 (p f p F )p F M FC 2 1. (8) We have returned back to the Landau Fermi liquid theory since high energy degrees of freedom are eliminated and the quasiparticles are introduced. The only difference between LFL, which serves as a basis when constructing the superconducting state, and Fermi liquid after FCQPT is that we have to expand the number of relevant low energy degrees of freedom by introducing new type of quasiparticles with the effective mass M FC given by Eq. (6) and the energy scale E 0 given by Eq. (8). Therefore, the single-particle spectrum ε(p) of system in question is characterized by two effective masses M L and M FC and by the scale E 0, which define the low temperature properties including the line shape of quasiparticle excitations [12, 14]. We note that both the 4

effective mass M FC and the scale E 0 are temperature independent at T < T c, where T c is the critical temperature of the superconducting phase transition [14]. Obviously, we cannot directly relate these new quasiparticle excitations with the quasiparticle excitations of an ideal Fermi gas because the system in question has undergone FCQPT. Nonetheless, the main basis of the Landau Fermi liquid theory survives FCQPT: the low energy excitations of a strongly correlated liquid with FC are quasiparticles. As it was shown above, properties of these new quasiparticles are closely related to the properties of the superconducting state. We may say that the quasiparticle system in the range (p i p f ) becomes very soft and is to be considered as a strongly correlated liquid. On the other hand, the system s properties and dynamics are dominated by a strong collective effect having its origin in FCQPT and determined by the macroscopic number of quasiparticles in the range (p i p f ). Such a system cannot be perturbed by the scattering of individual quasiparticles and has features of a quantum protectorate [12, 18, 19]. As any phase transition, FCQPT is related to the order parameter, which induces a broken symmetry. As we have seen above, the order parameter is the superconducting order parameter κ(p), while 1 being proportional to the coupling constant (see Eqs. (5) and (7)) can be small. Therefore, the existence of such a state, that is electron liquid with FC, can be revealed experimentally. Since the order parameter κ(p) is suppressed by the critical magnetic field B c, when B 2 c 2 1. If the coupling constant λ 0 0, the weak critical magnetic field B c 0 will destroy the state with FC converting the strongly correlated Fermi liquid into LFL. In this case the magnetic field play a role of the control parameter determining the effective mass [20] M FC(B) 1 B. (9) Note, that the outlined above behavior was observed experimentally in the heavy-electron metal YbRh 2 Si 2 [21]. If λ 0 is finite, the critical field is also finite, and Eq. (9) is valid at B > B c. In that case, the effective mass MFC(B) is finite, and the system is driven back to LFL and has the LFL behavior induced by the magnetic field. At a constant magnetic field, the low energy elementary excitations are characterized by MFC (B) and cannot be distinguished from Landau quasiparticles. As a result, at T 0, the WF law is held in accordance with experimental facts [2, 3]. On the hand, in contrast to the LFL theory, the effective mass MFC (B) depends on the magnetic field. Equation (9) shows that by applying a magnetic field B > B c the system can be driven back into LFL with the effective mass MFC (B) which is finite and independent of the temperature. This means that the low temperature properties depend on the effective mass in accordance with the LFL theory. In particular, the resistivity ρ(t) as a function of the temperature behaves as ρ(t) = ρ 0 + ρ(t) with ρ(t) = AT 2, and the factor A (MFC (B))2. At finite temperatures, the system persists to be LFL, but there is the crossover from the LFL behavior to the non-fermi liquid behavior at temperature T (B) B. At T > T (B), the effective mass starts to depend on the temperature M FC 1/T, and the resistivity possesses the non-fermi liquid behavior with a substantial linear term, ρ(t) = at +bt 2 [12, 20]. Such a behavior of the resistivity was observed in the cuprate superconductor Tl 2 Ba 2 CuO 6+δ (T c < 15 K) [22]. At B=10 T, ρ(t) is a linear function of the temperature between 120 mk and 1.2 K, whereas at B=18 T, the temperature dependence of the resistivity is consistent with ρ(t) = ρ 0 +AT 2 over the same temperature range 5

[22]. In LFL, the nuclear spin-lattice relaxation rate 1/T 1 is determined by the quasiparticles near the Fermi level whose population is proportional to M T, so that 1/T 1 T M is a constant [4, 5]. As it was shown, when the superconducting state is removed by the application of a magnetic field, the underlying ground state can be seen as the field induced LFL with effective mass depending on the magnetic field. As a result, the rate 1/T 1 follows the T 1 T = constant relation, that is the Korringa law is held. Unlike the behavior of LFL, as it follows from Eq. (9), 1/T 1 T M FC(B) decreases with increasing the magnetic field at T < T (B). While, at T > T (B), we observe that 1/T 1 T is a decreasing function of the temperature, 1/T 1 T M FC 1/T. These observations are in a good agreement with the experimental facts [4]. Since T (B) is an increasing function of the magnetic field, the Korringa law retains its validity to higher temperatures at elevated magnetic fields. We can also conclude, that at temperature T 0 T (B 0 ) and elevated magnetic fields B > B 0, the system shows a more pronounced metallic behavior since the effective mass decreases with increasing B, see Eq. (9). Such a behavior of the effective mass can be observed in the de Haas van Alphen-Shubnikov studies, 1/T 1 T measurements, and the resistivity measurements. These experiments can shed light on the physics of high-t c metals and reveal relationships between high-t c metals and heavy-electron metals. The existence of FCQPT can also be revealed experimentally because at densities x > x FC, or beyond the FCQPT point, the system should be LFL at sufficiently low temperatures [23]. Recent experimental data have shown that this liquid exists in heavily overdoped non-superconducting La 1.7 Sr 0.3 CuO 4 [6]. It is remarkable that up to T = 55 K the resistivity exhibits the T 2 behavior with no additional linear term, and the WF law is verified to within the experimental resolution [6]. While at elevated temperatures, a strong deviations from the LFL behavior are observed. We anticipate that the system can be again driven back to the LFL behavior by the application of sufficiently strong magnetic fields [23]. In summary, we have shown that the Landau Fermi liquid properties of high-t c metals observed at low temperatures in optimally hole-doped, overdoped and electron-doped cuprates by the application of a magnetic field higher than the critical field can be explained within the frameworks of the fermion condensation theory of high-t c superconductivity. In that case of LFL obtained by applying magnetic field, the effective mass depends on the magnetic field, the WF law and the Korringa law are held. These observations are in good agreement with experimental facts. The recent experimental verifications of the WF law in heavily overdoped, overdoped and optimally doped cuprates and the verification of the Korringa law in the electron-doped copper-oxide superconductor give strong support in favor of the existence of FC in high-t c metals. MYA is grateful to the Hebrew University Intramural fund of the Hebrew University for financial support. VRS is grateful to Racah Institute of Physics for the hospitality during his stay in Jerusalem. This work was supported in part by the Russian Foundation for Basic Research, project No 01-02-17189. References [1] L. D. Landau, Sov. Phys. JETP 3 (1956) 920. [2] C. Proust et al., Phys. Rev. Lett. 89 (2002) 147003. 6

[3] R. Bel et al., cond-mat/0304111. [4] G.-q. Zheng et al., Phys. Rev. Lett. 90 (2003) 197005. [5] J. Korringa, Physica (Utrecht) 16 (1950) 601. [6] S. Nakamae et al., cond-mat/0212283. [7] C.L. Kane, M.P.A. Fisher, Phys. Rev. Lett. 76 (1996) 3192. [8] T. Senthil, M.P.A. Fisher, Phys. Rev. B 62 (2000) 7850. [9] A. Houghton, S. Lee, J.P. Marston, Phys. Rev B 65 (2002) 22503. [10] V.A. Khodel, V.R. Shaginyan, JETP Lett. 51 (1990) 553. [11] V.A. Khodel, V.R. Shaginyan, V.V. Khodel, Phys. Rep. 249 (1994) 1. [12] M.Ya. Amusia, V.R. Shaginyan, Phys. Rev. B 63 (2001) 224507. [13] D.R. Tilley, J. Tilley, Superfluidity and Superconductivity, Bristol, Hilger, 1985. [14] M.Ya. Amusia, S.A. Artamonov, V.R. Shaginyan, JETP Lett. 74 (2001) 435; V.R. Shaginyan, Physica B 312-313C (2002) 413. [15] G.E. Volovik, JETP Lett. 53 (1991) 222. [16] I.E. Dzyaloshinskii, J. Phys. I (France) 6 (1996) 119. [17] V.Yu. Irkhin, A.A. Katanin, M.I. Katsnelson, Phys. Rev. Lett. 89 (2002) 076401. [18] R.B. Laughlin, D. Pines, Proc. Natl. Acad. Sci. USA 97 (2000) 28. [19] P.W. Anderson, cond-mat/0007185; cond-mat/0007287. [20] Yu.G. Pogorelov, V.R. Shaginyan, JETP Lett. 76 (2002) 532. [21] P. Gegenwart et al., Phys. Rev. Lett. 89 (2002) 056402. [22] A.P. Mackenzie et al., Phys. Rev. B 53 (1996) 5848. [23] V.R. Shaginyan, JETP Lett. 77 (2003) 99; JETP Lett. 77 (2003) 178. 7