Fermi liquid theory

Last updated

Fermi liquid theory (also known as Landau's Fermi-liquid theory) is a theoretical model of interacting fermions that describes the normal state of the conduction electrons in most metals at sufficiently low temperatures. [1] The theory describes the behavior of many-body systems of particles in which the interactions between particles may be strong. The phenomenological theory of Fermi liquids was introduced by the Soviet physicist Lev Davidovich Landau in 1956, and later developed by Alexei Abrikosov and Isaak Khalatnikov using diagrammatic perturbation theory. [2] The theory explains why some of the properties of an interacting fermion system are very similar to those of the ideal Fermi gas (collection of non-interacting fermions), and why other properties differ.

Contents

Fermi liquid theory applies most notably to conduction electrons in normal (non-superconducting) metals, and to liquid helium-3. [3] Liquid helium-3 is a Fermi liquid at low temperatures (but not low enough to be in its superfluid phase). An atom of helium-3 has two protons, one neutron and two electrons, giving an odd number of fermions, so the atom itself is a fermion. Fermi liquid theory also describes the low-temperature behavior of electrons in heavy fermion materials, which are metallic rare-earth alloys having partially filled f orbitals. The effective mass of electrons in these materials is much larger than the free-electron mass because of interactions with other electrons, so these systems are known as heavy Fermi liquids. Strontium ruthenate displays some key properties of Fermi liquids, despite being a strongly correlated material that is similar to high temperature superconductors such as the cuprates. [4] The low-momentum interactions of nucleons (protons and neutrons) in atomic nuclei are also described by Fermi liquid theory. [5]

Description

The key ideas behind Landau's theory are the notion of adiabaticity and the Pauli exclusion principle. [6] Consider a non-interacting fermion system (a Fermi gas), and suppose we "turn on" the interaction slowly. Landau argued that in this situation, the ground state of the Fermi gas would adiabatically transform into the ground state of the interacting system.

By Pauli's exclusion principle, the ground state of a Fermi gas consists of fermions occupying all momentum states corresponding to momentum with all higher momentum states unoccupied. As the interaction is turned on, the spin, charge and momentum of the fermions corresponding to the occupied states remain unchanged, while their dynamical properties, such as their mass, magnetic moment etc. are renormalized to new values. [6] Thus, there is a one-to-one correspondence between the elementary excitations of a Fermi gas system and a Fermi liquid system. In the context of Fermi liquids, these excitations are called "quasiparticles". [1]

Landau quasiparticles are long-lived excitations with a lifetime that satisfies where is the quasiparticle energy (measured from the Fermi energy). At finite temperature, is on the order of the thermal energy , and the condition for Landau quasiparticles can be reformulated as .

For this system, the many-body Green's function can be written [7] (near its poles) in the form

where is the chemical potential and is the energy corresponding to the given momentum state.

The value is called the quasiparticle residue and is very characteristic of Fermi liquid theory. The spectral function for the system can be directly observed via angle-resolved photoemission spectroscopy (ARPES), and can be written (in the limit of low-lying excitations) in the form:

where is the Fermi velocity. [8]

Physically, we can say that a propagating fermion interacts with its surrounding in such a way that the net effect of the interactions is to make the fermion behave as a "dressed" fermion, altering its effective mass and other dynamical properties. These "dressed" fermions are what we think of as "quasiparticles". [2]

Another important property of Fermi liquids is related to the scattering cross section for electrons. Suppose we have an electron with energy above the Fermi surface, and suppose it scatters with a particle in the Fermi sea with energy . By Pauli's exclusion principle, both the particles after scattering have to lie above the Fermi surface, with energies . Now, suppose the initial electron has energy very close to the Fermi surface Then, we have that also have to be very close to the Fermi surface. This reduces the phase space volume of the possible states after scattering, and hence, by Fermi's golden rule, the scattering cross section goes to zero. Thus we can say that the lifetime of particles at the Fermi surface goes to infinity. [1]

Similarities to Fermi gas

The Fermi liquid is qualitatively analogous to the non-interacting Fermi gas, in the following sense: The system's dynamics and thermodynamics at low excitation energies and temperatures may be described by substituting the non-interacting fermions with interacting quasiparticles, each of which carries the same spin, charge and momentum as the original particles. Physically these may be thought of as being particles whose motion is disturbed by the surrounding particles and which themselves perturb the particles in their vicinity. Each many-particle excited state of the interacting system may be described by listing all occupied momentum states, just as in the non-interacting system. As a consequence, quantities such as the heat capacity of the Fermi liquid behave qualitatively in the same way as in the Fermi gas (e.g. the heat capacity rises linearly with temperature).

Differences from Fermi gas

The following differences to the non-interacting Fermi gas arise:

Energy

The energy of a many-particle state is not simply a sum of the single-particle energies of all occupied states. Instead, the change in energy for a given change in occupation of states contains terms both linear and quadratic in (for the Fermi gas, it would only be linear, , where denotes the single-particle energies). The linear contribution corresponds to renormalized single-particle energies, which involve, e.g., a change in the effective mass of particles. The quadratic terms correspond to a sort of "mean-field" interaction between quasiparticles, which is parametrized by so-called Landau Fermi liquid parameters and determines the behaviour of density oscillations (and spin-density oscillations) in the Fermi liquid. Still, these mean-field interactions do not lead to a scattering of quasi-particles with a transfer of particles between different momentum states.

The renormalization of the mass of a fluid of interacting fermions can be calculated from first principles using many-body computational techniques. For the two-dimensional homogeneous electron gas, GW calculations [9] and quantum Monte Carlo methods [10] [11] [12] have been used to calculate renormalized quasiparticle effective masses.

Specific heat and compressibility

Specific heat, compressibility, spin-susceptibility and other quantities show the same qualitative behaviour (e.g. dependence on temperature) as in the Fermi gas, but the magnitude is (sometimes strongly) changed.

Interactions

In addition to the mean-field interactions, some weak interactions between quasiparticles remain, which lead to scattering of quasiparticles off each other. Therefore, quasiparticles acquire a finite lifetime. However, at low enough energies above the Fermi surface, this lifetime becomes very long, such that the product of excitation energy (expressed in frequency) and lifetime is much larger than one. In this sense, the quasiparticle energy is still well-defined (in the opposite limit, Heisenberg's uncertainty relation would prevent an accurate definition of the energy).

Structure

The structure of the "bare" particles (as opposed to quasiparticle) many-body Green's function is similar to that in the Fermi gas (where, for a given momentum, the Green's function in frequency space is a delta peak at the respective single-particle energy). The delta peak in the density-of-states is broadened (with a width given by the quasiparticle lifetime). In addition (and in contrast to the quasiparticle Green's function), its weight (integral over frequency) is suppressed by a quasiparticle weight factor . The remainder of the total weight is in a broad "incoherent background", corresponding to the strong effects of interactions on the fermions at short time scales.

Distribution

The distribution of particles (as opposed to quasiparticles) over momentum states at zero temperature still shows a discontinuous jump at the Fermi surface (as in the Fermi gas), but it does not drop from 1 to 0: the step is only of size .

Electrical resistivity

In a metal the resistivity at low temperatures is dominated by electron–electron scattering in combination with umklapp scattering. For a Fermi liquid, the resistivity from this mechanism varies as , which is often taken as an experimental check for Fermi liquid behaviour (in addition to the linear temperature-dependence of the specific heat), although it only arises in combination with the lattice. In certain cases, umklapp scattering is not required. For example, the resistivity of compensated semimetals scales as because of mutual scattering of electron and hole. This is known as the Baber mechanism. [13]

Optical response

Fermi liquid theory predicts that the scattering rate, which governs the optical response of metals, not only depends quadratically on temperature (thus causing the dependence of the DC resistance), but it also depends quadratically on frequency. [14] [15] [16] This is in contrast to the Drude prediction for non-interacting metallic electrons, where the scattering rate is a constant as a function of frequency. One material in which optical Fermi liquid behavior was experimentally observed is the low-temperature metallic phase of Sr2RuO4. [17]

Instabilities

The experimental observation of exotic phases in strongly correlated systems has triggered an enormous effort from the theoretical community to try to understand their microscopic origin. One possible route to detect instabilities of a Fermi liquid is precisely the analysis done by Isaak Pomeranchuk. [18] Due to that, the Pomeranchuk instability has been studied by several authors [19] with different techniques in the last few years and in particular, the instability of the Fermi liquid towards the nematic phase was investigated for several models.

Non-Fermi liquids

Non-Fermi liquids are systems in which the Fermi-liquid behaviour breaks down. The simplest example is a system of interacting fermions in one dimension, called the Luttinger liquid. [3] Although Luttinger liquids are physically similar to Fermi liquids, the restriction to one dimension gives rise to several qualitative differences such as the absence of a quasiparticle peak in the momentum dependent spectral function, and the presence of spin-charge separation and of spin-density waves. One cannot ignore the existence of interactions in one dimension and has to describe the problem with a non-Fermi theory, where Luttinger liquid is one of them. At small finite spin temperatures in one dimension the ground state of the system is described by spin-incoherent Luttinger liquid (SILL). [20]

Another example of non-Fermi-liquid behaviour is observed at quantum critical points of certain second-order phase transitions, such as heavy fermion criticality, Mott criticality and high- cuprate phase transitions. [8] The ground state of such transitions is characterized by the presence of a sharp Fermi surface, although there may not be well-defined quasiparticles. That is, on approaching the critical point, it is observed that the quasiparticle residue .

In optimally doped cuprates and iron-based superconductors, the normal state above the critical temperature shows signs of non-Fermi liquid behaviour, and is often called a strange metal. In this region of phase diagram, resistivity increases linearly in temperature and the Hall coefficient is found to depend on temperature. [21] [22]

Understanding the behaviour of non-Fermi liquids is an important problem in condensed matter physics. Approaches towards explaining these phenomena include the treatment of marginal Fermi liquids; attempts to understand critical points and derive scaling relations; and descriptions using emergent gauge theories with techniques of holographic gauge/gravity duality. [23] [24] [25]

See also

Related Research Articles

<span class="mw-page-title-main">BCS theory</span> Microscopic theory of superconductivity

In physics, theBardeen–Cooper–Schrieffer (BCS) theory is the first microscopic theory of superconductivity since Heike Kamerlingh Onnes's 1911 discovery. The theory describes superconductivity as a microscopic effect caused by a condensation of Cooper pairs. The theory is also used in nuclear physics to describe the pairing interaction between nucleons in an atomic nucleus.

<span class="mw-page-title-main">Bose–Einstein condensate</span> State of matter

In condensed matter physics, a Bose–Einstein condensate (BEC) is a state of matter that is typically formed when a gas of bosons at very low densities is cooled to temperatures very close to absolute zero. Under such conditions, a large fraction of bosons occupy the lowest quantum state, at which microscopic quantum mechanical phenomena, particularly wavefunction interference, become apparent macroscopically. More generally, condensation refers to the appearance of macroscopic occupation of one or several states: for example, in BCS theory, a superconductor is a condensate of Cooper pairs. As such, condensation can be associated with phase transition, and the macroscopic occupation of the state is the order parameter.

<span class="mw-page-title-main">Exciton</span> Quasiparticle which is a bound state of an electron and an electron hole

An electron and an electron hole that are attracted to each other by the Coulomb force can form a bound state called an exciton. It is an electrically neutral quasiparticle that exists mainly in condensed matter, including insulators, semiconductors, some metals, but also in certain atoms, molecules and liquids. The exciton is regarded as an elementary excitation that can transport energy without transporting net electric charge.

The quantum Hall effect is a quantized version of the Hall effect which is observed in two-dimensional electron systems subjected to low temperatures and strong magnetic fields, in which the Hall resistance Rxy exhibits steps that take on the quantized values

<span class="mw-page-title-main">Fermi–Dirac statistics</span> Statistical description for the behavior of fermions

Fermi–Dirac statistics is a type of quantum statistics that applies to the physics of a system consisting of many non-interacting, identical particles that obey the Pauli exclusion principle. A result is the Fermi–Dirac distribution of particles over energy states. It is named after Enrico Fermi and Paul Dirac, each of whom derived the distribution independently in 1926. Fermi–Dirac statistics is a part of the field of statistical mechanics and uses the principles of quantum mechanics.

<span class="mw-page-title-main">Fermi gas</span> Physical model of gases composed of many non-interacting identical fermions

A Fermi gas is an idealized model, an ensemble of many non-interacting fermions. Fermions are particles that obey Fermi–Dirac statistics, like electrons, protons, and neutrons, and, in general, particles with half-integer spin. These statistics determine the energy distribution of fermions in a Fermi gas in thermal equilibrium, and is characterized by their number density, temperature, and the set of available energy states. The model is named after the Italian physicist Enrico Fermi.

In condensed matter physics, a quasiparticle is a concept used to describe a collective behavior of a group of particles that can be treated as if they were a single particle. Formally, quasiparticles and collective excitations are closely related phenomena that arise when a microscopically complicated system such as a solid behaves as if it contained different weakly interacting particles in vacuum.

<span class="mw-page-title-main">Polaron</span> Quasiparticle in condensed matter physics

A polaron is a quasiparticle used in condensed matter physics to understand the interactions between electrons and atoms in a solid material. The polaron concept was proposed by Lev Landau in 1933 and Solomon Pekar in 1946 to describe an electron moving in a dielectric crystal where the atoms displace from their equilibrium positions to effectively screen the charge of an electron, known as a phonon cloud. This lowers the electron mobility and increases the electron's effective mass.

<span class="mw-page-title-main">Luttinger liquid</span> Theoretical model describing interacting fermions in a one-dimensional conductor

A Luttinger liquid, or Tomonaga–Luttinger liquid, is a theoretical model describing interacting electrons in a one-dimensional conductor. Such a model is necessary as the commonly used Fermi liquid model breaks down for one dimension.

In condensed matter physics, the Fermi surface is the surface in reciprocal space which separates occupied from unoccupied electron states at zero temperature. The shape of the Fermi surface is derived from the periodicity and symmetry of the crystalline lattice and from the occupation of electronic energy bands. The existence of a Fermi surface is a direct consequence of the Pauli exclusion principle, which allows a maximum of one electron per quantum state. The study of the Fermi surfaces of materials is called fermiology.

Jellium, also known as the uniform electron gas (UEG) or homogeneous electron gas (HEG), is a quantum mechanical model of interacting electrons in a solid where the positive charges are assumed to be uniformly distributed in space; the electron density is a uniform quantity as well in space. This model allows one to focus on the effects in solids that occur due to the quantum nature of electrons and their mutual repulsive interactions without explicit introduction of the atomic lattice and structure making up a real material. Jellium is often used in solid-state physics as a simple model of delocalized electrons in a metal, where it can qualitatively reproduce features of real metals such as screening, plasmons, Wigner crystallization and Friedel oscillations.

<span class="mw-page-title-main">Hofstadter's butterfly</span> Fractal describing the theorised behaviour of electrons in a magnetic field

In condensed matter physics, Hofstadter's butterfly is a graph of the spectral properties of non-interacting two-dimensional electrons in a perpendicular magnetic field in a lattice. The fractal, self-similar nature of the spectrum was discovered in the 1976 Ph.D. work of Douglas Hofstadter and is one of the early examples of modern scientific data visualization. The name reflects the fact that, as Hofstadter wrote, "the large gaps [in the graph] form a very striking pattern somewhat resembling a butterfly."

In Materials Science, heavy fermion materials are a specific type of intermetallic compound, containing elements with 4f or 5f electrons in unfilled electron bands. Electrons are one type of fermion, and when they are found in such materials, they are sometimes referred to as heavy electrons. Heavy fermion materials have a low-temperature specific heat whose linear term is up to 1000 times larger than the value expected from the free electron model. The properties of the heavy fermion compounds often derive from the partly filled f-orbitals of rare-earth or actinide ions, which behave like localized magnetic moments.

A Peierls transition or Peierls distortion is a distortion of the periodic lattice of a one-dimensional crystal. Atomic positions oscillate, so that the perfect order of the 1-D crystal is broken.

A composite fermion is the topological bound state of an electron and an even number of quantized vortices, sometimes visually pictured as the bound state of an electron and, attached, an even number of magnetic flux quanta. Composite fermions were originally envisioned in the context of the fractional quantum Hall effect, but subsequently took on a life of their own, exhibiting many other consequences and phenomena.

<span class="mw-page-title-main">Friedel oscillations</span>

Friedel oscillations, named after French physicist Jacques Friedel, arise from localized perturbations in a metallic or semiconductor system caused by a defect in the Fermi gas or Fermi liquid. Friedel oscillations are a quantum mechanical analog to electric charge screening of charged species in a pool of ions. Whereas electrical charge screening utilizes a point entity treatment to describe the make-up of the ion pool, Friedel oscillations describing fermions in a Fermi fluid or Fermi gas require a quasi-particle or a scattering treatment. Such oscillations depict a characteristic exponential decay in the fermionic density near the perturbation followed by an ongoing sinusoidal decay resembling sinc function. In 2020, magnetic Friedel oscillations were observed on a metal surface.

<span class="mw-page-title-main">Luttinger's theorem</span>

In condensed matter physics, Luttinger's theorem is a result derived by J. M. Luttinger and J. C. Ward in 1960 that has broad implications in the field of electron transport. It arises frequently in theoretical models of correlated electrons, such as the high-temperature superconductors, and in photoemission, where a metal's Fermi surface can be directly observed.

The term Dirac matter refers to a class of condensed matter systems which can be effectively described by the Dirac equation. Even though the Dirac equation itself was formulated for fermions, the quasi-particles present within Dirac matter can be of any statistics. As a consequence, Dirac matter can be distinguished in fermionic, bosonic or anyonic Dirac matter. Prominent examples of Dirac matter are graphene and other Dirac semimetals, topological insulators, Weyl semimetals, various high-temperature superconductors with -wave pairing and liquid helium-3. The effective theory of such systems is classified by a specific choice of the Dirac mass, the Dirac velocity, the gamma matrices and the space-time curvature. The universal treatment of the class of Dirac matter in terms of an effective theory leads to a common features with respect to the density of states, the heat capacity and impurity scattering.

The Pomeranchuk instability is an instability in the shape of the Fermi surface of a material with interacting fermions, causing Landau’s Fermi liquid theory to break down. It occurs when a Landau parameter in Fermi liquid theory has a sufficiently negative value, causing deformations of the Fermi surface to be energetically favourable. It is named after the Soviet physicist Isaak Pomeranchuk.

Kohn–Luttinger superconductivity is a theoretical mechanism for unconventional superconductivity proposed by Walter Kohn and Joaquin Mazdak Luttinger based on Friedel oscillations. In contrast to BCS theory, in which Cooper pairs are formed due to electron–phonon interaction, Kohn–Luttinger mechanism is based on fact that screened Coulomb interaction oscillates as and can create Cooper instability for non-zero angular momentum .

References

  1. 1 2 3 Phillips, Philip (2008). Advanced Solid State Physics. Perseus Books. p. 224. ISBN   978-81-89938-16-1.
  2. 1 2 Cross, Michael. "Fermi Liquid Theory: Principles" (PDF). California Institute of Technology. Retrieved 2 February 2015.
  3. 1 2 Schulz, H. J. (March 1995). "Fermi liquids and non–Fermi liquids". In "proceedings of les Houches Summer School Lxi", ed. E. Akkermans, G. Montambaux, J. Pichard, et J. Zinn-Justin (Elsevier, Amsterdam. 1995 (533). arXiv: cond-mat/9503150 . Bibcode:1995cond.mat..3150S.
  4. Wysokiński, Carol; et al. (2003). "Spin triplet superconductivity in Sr2RuO4" (PDF). Physica Status Solidi. 236 (2): 325–331. arXiv: cond-mat/0211199 . Bibcode:2003PSSBR.236..325W. doi:10.1002/pssb.200301672. S2CID   119378907 . Retrieved 8 April 2012.
  5. Schwenk, Achim; Brown, Gerald E.; Friman, Bengt (2002). "Low-momentum nucleon–nucleon interaction and Fermi liquid theory". Nuclear Physics A. 703 (3–4): 745–769. arXiv: nucl-th/0109059 . doi:10.1016/s0375-9474(01)01673-6. ISSN   0375-9474.
  6. 1 2 Coleman, Piers. Introduction to Many Body Physics (PDF). Rutgers University. p. 143. Archived from the original (PDF) on 2012-05-17. Retrieved 2011-02-14. (draft copy)
  7. Lifshitz, E. M.; Pitaevskii, L.P. (1980). Statistical Physics (Part 2). Landau and Lifshitz. Vol. 9. Elsevier. ISBN   978-0-7506-2636-1.
  8. 1 2 Senthil, Todadri (2008). "Critical Fermi surfaces and non-Fermi liquid metals". Physical Review B . 78 (3): 035103. arXiv: 0803.4009 . Bibcode:2008PhRvB..78c5103S. doi:10.1103/PhysRevB.78.035103. S2CID   118656854.
  9. R. Asgari; B. Tanatar (2006). "Many-body effective mass and spin susceptibility in a quasi-two-dimensional electron liquid" (PDF). Physical Review B. 74 (7): 075301. Bibcode:2006PhRvB..74g5301A. doi:10.1103/PhysRevB.74.075301. hdl: 11693/23741 .
  10. Y. Kwon; D. M. Ceperley; R. M. Martin (2013). "Quantum Monte Carlo calculation of the Fermi-liquid parameters in the two-dimensional electron gas". Physical Review B. 50 (3): 1684–1694. arXiv: 1307.4009 . Bibcode:1994PhRvB..50.1684K. doi:10.1103/PhysRevB.50.1684. PMID   9976356.
  11. M. Holzmann; B. Bernu; V. Olevano; R. M. Martin; D. M. Ceperley (2009). "Renormalization factor and effective mass of the two-dimensional electron gas". Physical Review B. 79 (4): 041308(R). arXiv: 0810.2450 . Bibcode:2009PhRvB..79d1308H. doi:10.1103/PhysRevB.79.041308. S2CID   12279058.
  12. N. D. Drummond; R. J. Needs (2013). "Diffusion quantum Monte Carlo calculation of the quasiparticle effective mass of the two-dimensional homogeneous electron gas". Physical Review B. 87 (4): 045131. arXiv: 1208.6317 . Bibcode:2013PhRvB..87d5131D. doi:10.1103/PhysRevB.87.045131. S2CID   53548304.
  13. Baber, W. G. (1937). "The Contribution to the Electrical Resistance of Metals from Collisions between Electrons". Proc. R. Soc. Lond. A. 158 (894): 383–396. Bibcode:1937RSPSA.158..383B. doi:10.1098/rspa.1937.0027.
  14. R. N. Gurzhi (1959). "MUTUAL ELECTRON CORRELATIONS IN METAL OPTICS". Sov. Phys. JETP. 8: 673–675.
  15. M. Scheffler; K. Schlegel; C. Clauss; D. Hafner; C. Fella; M. Dressel; M. Jourdan; J. Sichelschmidt; C. Krellner; C. Geibel; F. Steglich (2013). "Microwave spectroscopy on heavy-fermion systems: Probing the dynamics of charges and magnetic moments". Phys. Status Solidi B. 250 (3): 439–449. arXiv: 1303.5011 . Bibcode:2013PSSBR.250..439S. doi:10.1002/pssb.201200925. S2CID   59067473.
  16. C. C. Homes; J. J. Tu; J. Li; G. D. Gu; A. Akrap (2013). "Optical conductivity of nodal metals". Scientific Reports. 3 (3446): 3446. arXiv: 1312.4466 . Bibcode:2013NatSR...3E3446H. doi:10.1038/srep03446. PMC   3861800 . PMID   24336241.
  17. D. Stricker; J. Mravlje; C. Berthod; R. Fittipaldi; A. Vecchione; A. Georges; D. van der Marel (2014). "Optical Response of Sr2RuO4 Reveals Universal Fermi-Liquid Scaling and Quasiparticles Beyond Landau Theory". Physical Review Letters. 113 (8): 087404. arXiv: 1403.5445 . Bibcode:2014PhRvL.113h7404S. doi:10.1103/PhysRevLett.113.087404. PMID   25192127. S2CID   20176023.
  18. I. I. Pomeranchuk (1959). "ON THE STABILITY OF A FERMI LIQUID". Sov. Phys. JETP. 8: 361–362.
  19. Actually, this is a subject of investigation, see for example: https://arxiv.org/abs/0804.4422.
  20. M. Soltanieh-ha, A. E. Feiguin (2012). "Class of variational Ansätze for the spin-incoherent ground state of a Luttinger liquid coupled to a spin bath". Physical Review B. 86 (20): 205120. arXiv: 1211.0982 . Bibcode:2012PhRvB..86t5120S. doi:10.1103/PhysRevB.86.205120. S2CID   118724491.
  21. Lee, Patrick A.; Nagaosa, Naoto; Wen, Xiao-Gang (2006-01-06). "Doping a Mott insulator: Physics of high-temperature superconductivity". Reviews of Modern Physics. 78 (1): 17–85. arXiv: cond-mat/0410445 . doi:10.1103/RevModPhys.78.17. ISSN   0034-6861.
  22. Varma, Chandra M. (2020-07-07). "Colloquium : Linear in temperature resistivity and associated mysteries including high temperature superconductivity". Reviews of Modern Physics. 92 (3). arXiv: 1908.05686 . doi:10.1103/RevModPhys.92.031001. ISSN   0034-6861.
  23. Faulkner, Thomas; Polchinski, Joseph (2010). "Semi-Holographic Fermi Liquids". Journal of High Energy Physics. 2011 (6): 12. arXiv: 1001.5049 . Bibcode:2011JHEP...06..012F. CiteSeerX   10.1.1.755.3304 . doi:10.1007/JHEP06(2011)012. S2CID   119243857.
  24. Guo, Haoyu; Gu, Yingfei; Sachdev, Subir (2020). "Linear in temperature resistivity in the limit of zero temperature from the time reparameterization soft mode". Annals of Physics. 418: 168202. arXiv: 2004.05182 . doi:10.1016/j.aop.2020.168202.
  25. Wei, Chenan; Sedrakyan, Tigran A. (2023-08-02). "Strange metal phase of disordered magic-angle twisted bilayer graphene at low temperatures: From flat bands to weakly coupled Sachdev-Ye-Kitaev bundles". Physical Review B. 108 (6). arXiv: 2205.09766 . doi:10.1103/PhysRevB.108.064202. ISSN   2469-9950.