Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2013 Mar 27;110(15):5774–5778. doi: 10.1073/pnas.1218846110

Spectroscopic evidence for Fermi liquid-like energy and temperature dependence of the relaxation rate in the pseudogap phase of the cuprates

Seyed Iman Mirzaei a, Damien Stricker a, Jason N Hancock a,b, Christophe Berthod a, Antoine Georges a,c,d, Erik van Heumen a,e, Mun K Chan f, Xudong Zhao f,g, Yuan Li h, Martin Greven f, Neven Barišić f,i,j, Dirk van der Marel a,1
PMCID: PMC3625325  PMID: 23536291

Abstract

Cuprate high-Tc superconductors exhibit enigmatic behavior in the nonsuperconducting state. For carrier concentrations near “optimal doping” (with respect to the highest Tcs) the transport and spectroscopic properties are unlike those of a Landau–Fermi liquid. On the Mott-insulating side of the optimal carrier concentration, which corresponds to underdoping, a pseudogap removes quasi-particle spectral weight from parts of the Fermi surface and causes a breakup of the Fermi surface into disconnected nodal and antinodal sectors. Here, we show that the near-nodal excitations of underdoped cuprates obey Fermi liquid behavior. The lifetime τ(ω, T) of a quasi-particle depends on its energy ω as well as on the temperature T. For a Fermi liquid, 1/τ(ω, T) is expected to collapse on a universal function proportional to (ℏω)2 + (pπkBT)2. Magneto-transport experiments, which probe the properties in the limit ω = 0, have provided indications for the presence of a T2 dependence of the dc (ω = 0) resistivity of different cuprate materials. However, Fermi liquid behavior is very much about the energy dependence of the lifetime, and this can only be addressed by spectroscopic techniques. Our optical experiments confirm the aforementioned universal ω- and T dependence of 1/τ(ω, T), with p ∼ 1.5. Our data thus provide a piece of evidence in favor of a Fermi liquid-like scenario of the pseudogap phase of the cuprates.

Keywords: optical spectroscopy, superconductivity, mass renormalization, self energy


The compound HgBa2CuO4+δ (Hg1201) is the single-layer cuprate that exhibits the highest Tc (97 K). We therefore measured the optical conductivity of strongly underdoped single crystals of Hg1201 Inline graphic. Here we are interested in the optical conductivity of the CuO2 layers. We therefore express the optical conductivity as a 2D sheet conductance Inline graphic, where dc is the interlayer spacing. The real part of the sheet conductance normalized by the conduction quantum Inline graphic is shown in Fig. 1. As seen in the figure, a gap-like suppression below 140 meV is clearly observable for temperatures below Tc and remains visible in the normal state up to ∼250 K. This is a clear optical signature of the pseudogap. We also observe the zero-energy mode due to the free charge carrier response, which progressively narrows upon lowering the temperature. In materials where the charge carrier relaxation is dominated by impurity scattering, the width of this “Drude” peak corresponds to the relaxation rate of the charge carriers. Relaxation processes arising from interactions have the effect of replacing the constant (frequency-independent) relaxation rate with a frequency-dependent one. The general expression for the optical conductivity of interacting electrons is then

graphic file with name pnas.1218846110eq1.jpg

Fig. 1.

Fig. 1.

Optical sheet conductance of underdoped Hg1201 (Tc = 67 K). Temperatures from 10 to 390 K. Data below Tc are shown in gray.

The spectral weight K corresponds to minus the kinetic energy if the frequency integration of the experimental data is restricted to intraband transitions. The effect of electron–electron interactions and coupling to collective modes is described by the memory function Inline graphic, where Inline graphic represents the dynamical (or optical) relaxation rate in the case of a Fermi liquid.

The zero frequency limit of the optical conductivity of Fig. 1 corroborates the recently reported temperature dependence of the dc resistivity (1) as shown in Fig. 2. Because K is practically temperature independent in the normal state (2), the low-temperature T2 dependence of the resistivity is due to the quadratic temperature variation of Inline graphic. The infrared data confirm that Hg1201 exhibits the lowest residual resistance among the cuprates and a change to a linear temperature dependence above T* associated with the sudden closing of a pseudogap (3, 4). Fig. 2B shows this as a clear departure from the T2 curve at ∼5 × 104 K2. The dc transport data, owing to the higher precision, allow for Hg1201 crystals of the same composition and doping to identify T* ∼ 350 K as the temperature above which the resistivity has a linear temperature dependence, and T** ∼ 220 K as the temperature below which the temperature dependence is purely quadratic. Finally, superconducting fluctuations become noticeable at Inline graphic K.

Fig. 2.

Fig. 2.

dc resistivity of underdoped Hg1201 (Tc = 67 K). Inverse optical conductivity extrapolated to zero frequency (green circles) compared with measured dc resistivity (1) (solid red line) of a sample of the same composition and doping (A). The dashed black curve is a T-linear fit to the resistivity data above 350 K. (B) Same dc resistivity data as a function of T2, shown with a T2 fit to the data below 220 K. Definitions and values for T* and T** of ref. 1 are used.

The doping dependences of K and of the coherent spectral weight, defined as K*=K/(1+M1(ω,T)ℏω)|ω=0, are summarized in Fig. 3 for a number of hole-doped cuprates. The theoretical values of K based on the band parameters obtained from local density approximation (LDA) ab initio calculations are about a factor of 2 larger than the measured values, which is due to strong correlation predicted by the Hubbard model for Inline graphic (6). K decreases when the hole doping decreases, but does not extrapolate to zero for zero doping in accordance with the analysis of Comanac et al. (7). In contrast, the coherent spectral weight K* is proportional to the hole doping x: K = xK0, where Inline graphic meV, in agreement with the trend observed for La2−xSr2CuO4 (8) and YBa2Cu3O7−δ (9). This provides strong evidence that a Mott insulator is approached as the doping is reduced. It cannot be determined from these data whether this occurs because (i) the quasi-particle residue is gradually suppressed (10, 11) or (ii) the Fermi surface arcs shrink to zero without vanishing of the nodal spectral weight (12, 13).

Fig. 3.

Fig. 3.

Conduction band spectral weight per CuO2 layer for a large number of cuprate superconductors. Closed symbols: coherent spectral weight K*. The gray line is a linear least-squares fit, K* = 496x, where x is the nominal hole doping. Open symbols: total spectral weight, K. Data are presented for Hg1201 with 10% doping (see Supporting Information). For the other materials see table 1 in ref. 5. The black solid curve represents K calculated from ab initio band parameters for Hg1201 (Supporting Information). For the Bi2Sr2CuO6+δ (Bi2201) sample with hole doping below 0.10, K* could not be calculated, because in this case we obtain Inline graphic for Inline graphic, possibly due to disorder.

The real and imaginary parts of the memory function of underdoped Hg1201 with Inline graphic K are shown in Fig. 4 for temperatures from 10 to 390 K. Inline graphic has a linear slope extrapolating to ω = 0, which becomes less steep at higher temperatures. The maximum at ∼105 meV erodes gradually as temperature increases, but a residue of this structure remains visible even at 390 K. In Fig. 4E we show a plot of the frequency- and temperature-dependent mass enhancement factor Inline graphic. Above 60 meV Inline graphic is a monotonously decreasing function of temperature. For ω that is not too large, Inline graphic is, roughly speaking, a plateau at low temperature, with a weak maximum at T(max) followed by a linear-like decrease at higher temperature. T(max) increases when ω decreases and for Inline graphic extrapolates to 212 K ∼ T**, indicating another way of identifying T**. The increase of Inline graphic from about 3 at 390 K to 5 at T**, taken together with the strong temperature dependence of Inline graphic near its maximum at 105 meV, indicates that the charge carriers become increasingly renormalized when the temperature decreases. Our results also corroborate the observation in ref. 14 that the integrated optical conductivity does not decrease when T decreases, so that no opening of an optical pseudogap is seen when, at T*, part of the Fermi surface is removed by a pseudogap, despite the emergence at this temperature of a novel ordered state with two Ising-like magnetic collective modes at 54 and 39 meV as observed with inelastic neutron scattering (15).

Fig. 4.

Fig. 4.

Optical self-energy as a function of frequency and temperature. (A) Real and (B) imaginary part of the memory function as a function of ℏω for underdoped Hg1201 Inline graphic. Spectra are shown in 10-K intervals for temperatures from 10 to 390 K. Thick lines are used to highlight the 10-, 70-, 280-, and 390-K data. (C) Effective mass Inline graphic as a function of temperature for selected values of ℏω in 10-meV intervals from 10 to 60 meV. (D) Relaxation rate Inline graphic as a function of temperature for selected values of ℏω in 10-meV intervals from 10 to 150 meV. Thick lines are used to highlight the data at selected energies. In C the approximate temperatures of the maxima T(max) are indicated with an arrow for ω with the corresponding color. (E) shows the same T(max) versus ω. The solid line is a linear fit, which extrapolates to Inline graphic for Inline graphic. All data in the superconducting state are in gray. The temperature range (370 K) of C and D is chosen to match the frequency range of A and B (200 meV) according to the scaling relation Inline graphic.

Turning now to the dynamical relaxation rate Inline graphic, we observe from Fig. 4 that its frequency dependence exhibits an upward curvature for all temperatures. Also, the temperature dependence has a T2 component at the lowest frequencies. Earlier indications for T2 dependence of the scattering rate came from the dc Inline graphic resistivity (1, 16, 17). For frequencies above 50 meV this component is either absent or completely masked by the onset of superconductivity (gray segments of the temperature traces). Although Inline graphic has no maximum as a function of temperature, the curves have an inflection point which shifts from roughly 200 to 100 K when the frequency is raised from 10 to 50 meV. The saturation of Inline graphic and the merger of the resistivity with a T2 dependence indicate that the system enters a Fermi liquid-like state at ∼200 K. We notice that at temperatures above Tc the initial rise is given by a linear slope as a function of ω2 (Fig. 5, Inset). For an ideal Fermi liquid, Inline graphic in the relevant range of ω and T is, to a good approximation,

graphic file with name pnas.1218846110eq2.jpg

where Inline graphic is proportional to the quasi-particle residue, C is a constant with units of inverse energy, and Inline graphic (ref. 18 and Supporting Information). To check possible Fermi liquid characteristics of the data, we introduce a single parameter ξ defined as Inline graphic, and we investigate M2 as a function of ξ. As shown in Fig. 6 for three underdoped cuprate materials [Hg1201, ortho-II Y Ba2Cu3O6.5 (Y123) (ref. 19), and Bi2201 (ref. 20)] with hole concentration Inline graphic, the M2 data of the normal state collapse in the low-energy range on a single scaling curve for Inline graphic. This value of p was obtained by searching for the best scaling collapse for Inline graphic in steps of 0.1 (Supporting Information). Comparing the functional form of Inline graphic for these three materials, we make the following observations: (i) Going from Hg1201 to Bi2201 (Fig. 6, Left and Right, respectively) in this plot, the residual Inline graphic value of Inline graphic increases from 0 to about 80 meV. Indeed, it is generally thought that the relatively low values of Tc in single-layer Bi2201 have to do with strong scattering by disorder (1, 21). (ii) We also notice that in the case of Bi2201 some negative curvature appears at the lowest energies, which is an indication that the Fermi liquid characteristics are affected to some extent, and appear to be relatively fragile with respect to disorder. (iii) The implications of the loss of scaling above 100 meV in the Bi2201 data are not entirely clear. In principle there is no reason to expect scaling, because this is clearly beyond the range of “universal” Fermi liquid behavior. However, the single-parameter scaling seems to persist into this regime for the other two materials (Y123 and Hg1201), leading to the speculation that impurity scattering also contributes to the disappearance of scaling above 100 meV for the Bi2201 sample.

Fig. 5.

Fig. 5.

Dynamical relaxation rate of underdoped Hg1201. Imaginary part of the memory function of underdoped Hg1021 Inline graphic for temperatures between 10 and 390 K in 20-K steps as a function of ω2. Thick lines are used to highlight the 70-, 280-, and 390-K data. (Inset) Zoom of the low-ω range showing a linear fit; temperatures are from 70 to 270 K in 20-K steps.

Fig. 6.

Fig. 6.

Collapse of the frequency and temperature dependence of the relaxation rate of underdoped cuprate materials. Normal state Inline graphic as a function of Inline graphic with Inline graphic. (A) Hg1201 Inline graphic. (B) Y123 Inline graphic, spectra by Hwang et al. (19) (digitized data of Fig. 6 represented here as a function of ξ2). (C) Bi2201 Inline graphic; data of van Heumen et al. (20) represented here as a function of ξ2. The data displayed in A and C are in 10-K intervals with color coding indicated for temperatures in 40-K steps. In between these steps the color evolves gradually as a function of temperature. In B the color coding is given for all temperatures displayed.

The most important observation borne out by these data is that the frequency dependence of Inline graphic and Inline graphic follows by and large the behavior expected for a Fermi liquid: At low frequencies and temperatures Inline graphic is indeed a linear function of ω, and Inline graphic scales with Inline graphic. Other hints at possible Fermi liquid behavior came from the recent discovery of quantum oscillations at low temperature and high magnetic field in underdoped YBa2Cu3O6+δ (13) and YBa2Cu4O8 (22), from the observation of the Fermi–Dirac statistics underlying the quantum oscillations (23), and from the two-fluid analysis (24) of NMR data (25). We note that recent theories (e.g., refs. 2628) have emphasized the possible relevance of Fermi liquid concepts––or a hidden form of these in the superconducting regime (29)––to the metallic state of hole-doped cuprates. Our experimental observations provide a strong incentive for further theoretical work in this direction. We highlight two striking aspects of the data: (i) The slope Inline graphic for Inline graphic decreases significantly as a function of increasing temperature; and (ii) Inline graphic. We speculate that these issues are related to the progressive filling-in of the pseudogap as a function of increasing temperature. Already in a two-fluid picture of a nodal Fermi liquid in parallel to an antinodal liquid, nonuniversal features (for Fermi liquids) are introduced in the optical conductivity, because the properties at the Fermi surface change gradually from Fermi liquid at the nodes (30) to strongly incoherent and pseudogapped at the hot spots near the antinodes (31). In fact, also in other compounds p is found to be different from 2 (3234). Recently, Maslov and Chubukov interpreted this as a combination of Fermi liquid scattering and an additional source of elastic scattering from magnetic moments or resonant levels (35).

Theoretically, it is expected that the T2- and ω2 dependence of Inline graphic is limited to ℏω and pπkBT lower than some energy scale ξ0, which in the context of single-parameter scaling behavior of a Fermi liquid is proportional to the effective Fermi energy. Electronic correlations strongly reduce this energy scale compared with the bare Fermi energy. For most materials the issue of the Fermi liquid-like frequency dependence of Inline graphic has remained largely unexplored. This is related to the difficulty that in cases such as the heavy fermion materials where this type of coupling dominates, the range of Fermi liquid behavior is smaller than 10 meV, making it particularly difficult to obtain the required measurement accuracy in an infrared experiment. Clean underdoped cuprates present in this respect a favorable exception because, as can be seen from Figs. 5 and 6, the relevant energy scale ξ0 is about 100 meV for a doping level around 10%. Above this energy, Inline graphic crosses over to a more linear trend as a function of both ω and T. This suggests that in cuprates the range of applicability of Fermi liquid behavior is limited by a different scattering mechanism that develops at high T and high ω as the pseudogap gets filled.

The ξ2-dependence of the relaxation rate can be understood as follows: An electron at a distance ξ above the Fermi energy can, as a result of electron–electron interactions, decay to a final state ξ−Ω by creating an electron–hole pair of energy Ω. The density of states of electron–hole pairs is the spin (charge) susceptibility Inline graphic, where spin (charge) refers to electron–hole pairs carrying (no) net spin. Inline graphic can be strongly renormalized, but the property that Inline graphic in the limit Inline graphic is generic for Fermi liquids (24). Integration of the susceptibility multiplied with the interaction vertex Inline graphic over all possible decay channels from zero to ξ leads us to conclude that indeed Inline graphic, as reported experimentally in the present article. In this description the cross-over ξ0 corresponds to the energy where Inline graphic is truncated, leading to a leveling off of M2 for Inline graphic. The strong temperature dependence of Inline graphic is also a natural consequence of this description; it was shown in ref. 36 that, in leading orders of temperature, Inline graphic of a correlated Fermi liquid decreases as a function of temperature.

In summary, we have shown from optical spectroscopy measurements that the ungapped near-nodal excitations of underdoped cuprate superconductors obey Fermi liquid behavior when materials with reduced amount of disorder are considered. This observation, which is at variance with some established paradigms, provides leads toward understanding of the metallic state and high-temperature superconductivity in these materials.

Materials and Methods

Sample Preparation.

Single crystals were grown using a flux method, characterized, and heat treated to the desired doping level as described in refs. 37 and 38. The conductivity data in Fig. 1 are of a sample which has an onset critical temperature of 67 K and a transition width of 2 K. The crystal surface is oriented along the ab plane with a dimension of about 1.51 × 1.22 mm2. Hg1201 samples are hygroscopic. Therefore, the last stage of the preparation of the sample surface is done under a continuous flow of nitrogen, upon which the sample is transferred to a high-vacuum chamber (10−7 mbar) within a few minutes. Before each measurement the surface is carefully checked for any evidence of oxidation.

Comparison with dc Resistivity.

Transport measurements have been performed using the four-terminal method. Due to the irregular shape of the cleaved samples the absolute value of the dc resistivity can only be determined with about 20% accuracy. However, we obtained very high relative accuracy of the temperature dependence of the dc resistivity, as seen from identical temperature dependences of samples of the same composition and doping, regardless of having significantly different dimensions and shapes. An independent check of the dc resistivity was obtained from the Inline graphic limit of the experimental infrared optical conductivity (Fig. 2). The dc resistivity had to be scaled by a factor of 0.66 to match the optical data, most likely due to the aforementioned influence of the irregular shape of the crystals on the absolute value of the measured dc resistances. The excellent match of the two temperature dependences demonstrates the high quality of both dc resistivity and optical conductivity data.

Supplementary Material

Supporting Information

Acknowledgments

We thank A. Chubukov, A. J. Leggett, T. Giamarchi, T. M. Rice, and T. Timusk for discussions and communications. This work was supported by the Swiss National Science Foundation (SNSF) through Grant 200020-140761 and the National Center of Competence in Research, Materials with Novel Electronic Properties. The crystal growth and characterization work was supported by the US Department of Energy, Office of Basic Energy Sciences. X.Z. acknowledges support from the National Natural Science Foundation, China; N.B. acknowledges support from the European Commission under the [Intra-European Fellowships (IEF)] Marie Curie Fellowship Programme; and E.v.H. acknowledges support through the Vernieuwingsimpuls (VENI) program funded by the Netherlands Organisation for Scientific Research.

Footnotes

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1218846110/-/DCSupplemental.

References

  • 1.Barišić N, et al. 2012. Universal sheet resistance of the cuprate superconductors. arXiv:1207.1504.
  • 2.Norman M, Chubukov A, van Heumen E, Kuzmenko AB, van der Marel D. Optical integral in the cuprates and the question of sum-rule violation. Phys Rev B. 2007;76(22):220509. [Google Scholar]
  • 3.Bucher B, Steiner P, Karpinski J, Kaldis E, Wachter P. Influence of the spin gap on the normal state transport in YBa2Cu4O8. Phys Rev Lett. 1993;70(13):2012–2015. doi: 10.1103/PhysRevLett.70.2012. [DOI] [PubMed] [Google Scholar]
  • 4.Ando Y, Komiya S, Segawa K, Ono S, Kurita Y. Electronic phase diagram of high-Tc cuprate superconductors from a mapping of the in-plane resistivity curvature. Phys Rev Lett. 2004;93(26 Pt 1):267001. doi: 10.1103/PhysRevLett.93.267001. [DOI] [PubMed] [Google Scholar]
  • 5.van Heumen E, et al. Optical determination of the relation between the electron-boson coupling function and the critical temperature in high-Tc cuprates. Phys Rev B. 2009;79(18):184512. [Google Scholar]
  • 6.Rozenberg MJ, et al. Optical conductivity in Mott-Hubbard systems. Phys Rev Lett. 1995;75(1):105–108. doi: 10.1103/PhysRevLett.75.105. [DOI] [PubMed] [Google Scholar]
  • 7.Comanac A, de’ Medici L, Capone M, Millis AJ. Optical conductivity and the correlation strength of high-temperature copper-oxide superconductors. Nat Phys. 2008;4(4):287–290. [Google Scholar]
  • 8.Uchida S, et al. Optical spectra of La2-xSrxCuO4: Effect of carrier doping on the electronic structure of the CuO2 plane. Phys Rev B Condens Matter. 1991;43(10):7942–7954. doi: 10.1103/physrevb.43.7942. [DOI] [PubMed] [Google Scholar]
  • 9.Padilla WJ, et al. Constant effective mass across the phase diagram of high-Tc cuprates. Phys Rev B. 2005;72(6):060511. [Google Scholar]
  • 10.Brinkman WF, Rice TM. Application of Gutzwiller’s variational method to the metal-insulator transition. Phys Rev B. 1970;2(10):4302–4304. [Google Scholar]
  • 11.Fournier D, et al. Loss of nodal quasiparticle integrity in underdoped YBa2Cu3O6+x. Nat Phys. 2010;6(11):905–911. [Google Scholar]
  • 12.Kanigel A, et al. Evolution of the pseudogap from Fermi arcs to the nodal liquid. Nat Phys. 2006;2(7):447–451. [Google Scholar]
  • 13.Doiron-Leyraud N, et al. Quantum oscillations and the Fermi surface in an underdoped high-Tc superconductor. Nature. 2007;447(7144):565–568. doi: 10.1038/nature05872. [DOI] [PubMed] [Google Scholar]
  • 14.Santander-Syro AF, et al. Absence of a loss of in-plane infrared spectral weight in the pseudogap regime of Bi2Sr2CaCu2O8−δ. Phys Rev Lett. 2001;88(9):097005. doi: 10.1103/PhysRevLett.88.097005. [DOI] [PubMed] [Google Scholar]
  • 15.Li Y, et al. Two Ising-like magnetic excitations in a single-layer cuprate superconductor. Nat Phys. 2012;8(5):404–410. [Google Scholar]
  • 16.Ando Y, Kurita Y, Komiya S, Ono S, Segawa K. Evolution of the Hall coefficient and the peculiar electronic structure of the cuprate superconductors. Phys Rev Lett. 2004;92(19):197001. doi: 10.1103/PhysRevLett.92.197001. [DOI] [PubMed] [Google Scholar]
  • 17.Hussey NE. What drives pseudogap physics in high-Tc cuprates? A view from the (resistance)bridge. J Phys Chem Solids. 2011;72(5):529–532. [Google Scholar]
  • 18.Berthod C, et al. 2013. Non-Drude universal scaling laws for the optical response of local Fermi liquids Phys Rev B 87(11):115109.
  • 19.Hwang J, et al. A-axis optical conductivity of detwinned ortho-II YBa2Cu3O6.50. Phys Rev B. 2006;73:014508. [Google Scholar]
  • 20.van Heumen E, Meevasana W, Kuzmenko AB, Eisaki H, van der Marel D. Doping-dependent optical properties of Bi2201. New J Phys. 2009;11(5):055067. [Google Scholar]
  • 21.Eisaki H, et al. Effect of chemical inhomogeneity in bismuth-based copper oxide superconductors. Phys Rev B. 2004;69(8):064512. [Google Scholar]
  • 22.Bangura AF, et al. Small Fermi surface pockets in underdoped high temperature superconductors: Observation of Shubnikov-de Haas oscillations in YBa2Cu4O8. Phys Rev Lett. 2008;100(4):047004. doi: 10.1103/PhysRevLett.100.047004. [DOI] [PubMed] [Google Scholar]
  • 23.Sebastian SE, et al. Fermi-liquid behavior in an underdoped high-Tc superconductor. Phys Rev B. 2010;81(14):140505. [Google Scholar]
  • 24.Barzykin V, Pines D. Universal behaviour and the two-component character of magnetically underdoped cuprate superconductors. Adv Phys. 2009;58(1):1–65. [Google Scholar]
  • 25.Haase J, et al. Two-component uniform spin susceptibility of superconducting HgBa2CuO2+δ single crystals measured using 63Cu and 199Hg nuclear magnetic resonance. Phys Rev B. 2012;85(10):104517. [Google Scholar]
  • 26.Gull E, Ferrero M, Parcollet O, Georges A, Millis AJ. Momentum-space anisotropy and pseudogaps: A comparative cluster dynamical mean-field analysis of the doping-driven metal-insulator transition in the two-dimensional Hubbard model. Phys Rev B. 2010;82(15):155101. [Google Scholar]
  • 27.Shastry B. Extremely correlated Fermi liquids. Phys Rev Lett. 2011;107(5):056404. doi: 10.1103/PhysRevLett.107.056403. [DOI] [PubMed] [Google Scholar]
  • 28.Bergeron D, Hankevych V, Kyung B, Tremblay AM. Optical and dc conductivity of the two-dimensional Hubbard model in the pseudogap regime and across the antiferromagnetic quantum critical point including vertex corrections. Phys Rev B. 2011;84(8):085128. [Google Scholar]
  • 29.Anderson PW. Hidden Fermi liquid: The secret of high-Tc cuprates. Phys Rev B. 2008;78(17):174505. [Google Scholar]
  • 30.Ioffe L, Millis A. Zone-diagonal-dominated transport in high-Tc cuprates. Phys Rev B. 1998;58(17):11631–11637. [Google Scholar]
  • 31.Stojkovic BP, Pines D. Theory of the longitudinal and Hall conductivities of the cuprate superconductors. Phys Rev B. 1997;55(13):8576–8595. [Google Scholar]
  • 32.Basov DN, Averitt RD, van der Marel D, Dressel M, Haule K. Electrodynamics of correlated electron materials. Rev Mod Phys. 2011;83(2):471. [Google Scholar]
  • 33.Dressel M. Quantum criticality in organic conductors? Fermi liquid versus non-Fermi-liquid behaviour. J Phys Condens Matter. 2011;23(29):293201. doi: 10.1088/0953-8984/23/29/293201. [DOI] [PubMed] [Google Scholar]
  • 34.Nagel U, et al. Optical spectroscopy shows that the normal state of URu2Si2 is an anomalous Fermi liquid. Proc Natl Acad Sci USA. 2012;109(47):19161–19165. doi: 10.1073/pnas.1208249109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Maslov D, Chubukov A. First-Matsubara-frequency rule in a Fermi liquid. Part II: Optical conductivity and comparison to experiment. Phys Rev B. 2012;86(15):155137. [Google Scholar]
  • 36.Chubukov A, Maslov D. Singular corrections to the Fermi-liquid theory. Phys Rev B. 2004;69(12):121102. [Google Scholar]
  • 37.Zhao X, et al. Crystal growth and characterization of the model high- temperature superconductor HgBa2CuO4+δ. Adv Mater (Deerfield Beach Fla) 2006;18(24):3243–3247. [Google Scholar]
  • 38.Barišić N, et al. Demonstrating the model nature of the high-temperature superconductor HgBa2CuO4+δ. Phys Rev B. 2008;78(5):054518. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supporting Information

Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES