Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Letter
  • Published:

Thermodynamic signatures of quantum criticality in cuprate superconductors

Abstract

The three central phenomena of cuprate (copper oxide) superconductors are linked by a common doping level p*—at which the enigmatic pseudogap phase ends and the resistivity exhibits an anomalous linear dependence on temperature, and around which the superconducting phase forms a dome-shaped area in the phase diagram1. However, the fundamental nature of p* remains unclear, in particular regarding whether it marks a true quantum phase transition. Here we measure the specific heat C of the cuprates Eu-LSCO and Nd-LSCO at low temperature in magnetic fields large enough to suppress superconductivity, over a wide doping range2 that includes p*. As a function of doping, we find that Cel/T is strongly peaked at p* (where Cel is the electronic contribution to C) and exhibits a log(1/T) dependence as temperature T tends to zero. These are the classic thermodynamic signatures of a quantum critical point3,4,5, as observed in heavy-fermion6 and iron-based7 superconductors at the point where their antiferromagnetic phase comes to an end. We conclude that the pseudogap phase of cuprates ends at a quantum critical point, the associated fluctuations of which are probably involved in d-wave pairing and the anomalous scattering of charge carriers.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Specific heat data in various cuprates.
Fig. 2: Specific heat of Eu-LSCO and Nd-LSCO.
Fig. 3: Doping dependence of the electronic specific heat.

Similar content being viewed by others

Data availability

The data that support the plots within this paper and other findings of this study are available from the corresponding author upon reasonable request.

References

  1. Taillefer, L. Scattering and pairing in cuprate superconductors. Annu. Rev. Condens. Matter Phys. 1, 51–70 (2010).

    Article  ADS  CAS  Google Scholar 

  2. Collignon, C. et al. Fermi-surface transformation across the pseudogap critical point of the cuprate superconductor La1.6−xNd0.4SrxCuO4. Phys. Rev. B 95, 224517 (2017).

    Article  ADS  Google Scholar 

  3. Löhneysen, H. v. et al. Fermi-liquid instabilities at magnetic quantum phase transitions. Rev. Mod. Phys. 79, 1015–1075 (2007).

    Article  ADS  Google Scholar 

  4. Monthoux, P., Pines, D. & Lonzarich, G. G. Superconductivity without phonons. Nature 450, 1177–1183 (2007).

    Article  ADS  CAS  Google Scholar 

  5. Shibauchi, T., Carrington, A. & Matsuda, Y. A quantum critical point lying beneath the superconducting dome in iron pnictides. Annu. Rev. Condens. Matter Phys. 5, 113–135 (2014).

    Article  ADS  CAS  Google Scholar 

  6. Löhneysen, H. v. et al. Non-Fermi-liquid behavior in a heavy-fermion alloy at a magnetic instability. Phys. Rev. Lett. 72, 3262–3265 (1994).

    Article  ADS  Google Scholar 

  7. Walmsley, P. et al. Quasiparticle mass enhancement close to the quantum critical point in BaFe2(As1−xPx)2. Phys. Rev. Lett. 110, 257002 (2013).

    Article  ADS  CAS  Google Scholar 

  8. Doiron-Leyraud, N. et al. Correlation between linear resistivity and T c in the Bechgaard salts and the pnictide superconductor Ba(Fe1−xCox)2As2. Phys. Rev. B 80, 214531 (2009).

    Article  ADS  Google Scholar 

  9. Ramshaw, B. J. et al. Quasiparticle mass enhancement approaching optimal doping in a high-T c superconductor. Science 348, 317–320 (2015).

    Article  ADS  CAS  Google Scholar 

  10. Kačmarčík, J. et al. Unusual interplay between superconductivity and field-induced charge order in YBa2Cu3Oy. Phys. Rev. Lett. 121, 167002 (2018).

    Article  ADS  Google Scholar 

  11. Loram, J. W. et al. Electronic specific heat of YBa2Cu3O6+x from 1.8 to 300 K. Phys. Rev. Lett. 71, 1740–1743 (1993).

    Article  ADS  CAS  Google Scholar 

  12. Loram, J. W. et al. Specific heat evidence on the normal state pseudogap. J. Phys. Chem. Solids 59, 2091–2094 (1998).

    Article  ADS  CAS  Google Scholar 

  13. Nakamae, S. et al. Electronic ground state of heavily overdoped nonsuperconducting La2−xSrxCuO4. Phys. Rev. B 68, 100502 (2003).

    Article  ADS  Google Scholar 

  14. Matt, C. E. et al. Electron scattering, charge order, and pseudogap physics in La1.6−xNd0.4SrxCuO4: an angle-resolved photoemission spectroscopy study. Phys. Rev. B 92, 134524 (2015).

    Article  ADS  Google Scholar 

  15. Trovarelli, O. et al. YbRh2Si2: pronounced non-Fermi-liquid effects above a low-lying magnetic phase transition. Phys. Rev. Lett. 85, 626–629 (2000).

    Article  ADS  CAS  Google Scholar 

  16. Bianchi, A. et al. Avoided antiferromagnetic order and quantum critical point in CeCoIn5. Phys. Rev. Lett. 91, 257001 (2003).

    Article  ADS  CAS  Google Scholar 

  17. Momono, N. et al. Low-temperature electronic specific heat of La2−xSrxCuO4 and La2−xSrxCu1−yZnyO4. Evidence for a d-wave superconductor. Physica C 233, 395–401 (1994).

    Article  ADS  CAS  Google Scholar 

  18. Wade, J. M. et al. Electronic specific heat of Tl2Ba2CuO6+δ from 2 K to 300 K for 0 < δ < 0.1. J. Supercond. 7, 261–264 (1994).

    Article  ADS  CAS  Google Scholar 

  19. Bangura, A. F. et al. Fermi surface and electronic homogeneity of the overdoped cuprate superconductor Tl2Ba2CuO6+δ as revealed by quantum oscillations. Phys. Rev. B 82, 140501 (2010).

    Article  ADS  Google Scholar 

  20. Loram, J. W. et al. Evidence on the pseudogap and condensate from the electronic specific heat. J. Phys. Chem. Solids 62, 59–64 (2001).

    Article  ADS  CAS  Google Scholar 

  21. Loret, B. et al. Raman and ARPES combined study on the connection between the existence of the pseudogap and the topology of the Fermi surface in Bi2Sr2CaCu2O8+δ. Phys. Rev. B 97, 174521 (2018).

    Article  ADS  Google Scholar 

  22. Chan, M. K. et al. Single reconstructed Fermi surface pocket in an underdoped single-layer cuprate superconductor. Nat. Commun. 7, 12244 (2016).

    Article  ADS  CAS  Google Scholar 

  23. Grissonnanche, G. et al. Direct measurement of the upper critical field in cuprate superconductors. Nat. Commun. 5, 3280 (2014).

    Article  CAS  Google Scholar 

  24. Motoyama, E. M. et al. Spin correlations in the electron-doped high-transition-temperature superconductor Nd2−xCexCuOδ. Nature 445, 186–189 (2007).

    Article  ADS  CAS  Google Scholar 

  25. Dagan, Y. et al. Evidence for a quantum phase transition in Pr2−xCexCuO4−δ from transport measurements. Phys. Rev. Lett. 92, 167001 (2004).

    Article  ADS  CAS  Google Scholar 

  26. Storey, J. G. et al. Pseudogap ground state in high-temperature superconductors. Phys. Rev. B 78, 140506(R) (2008).

    Article  ADS  Google Scholar 

  27. Tranquada, J. M. et al. Coexistence of, and competition between, superconductivity and charge-stripe order in La1.62−xNd0.4SrxCuO4. Phys. Rev. Lett. 78, 338–341 (1997).

    Article  ADS  CAS  Google Scholar 

  28. Nachumi, B. et al. Muon spin relaxation study of the stripe phase order in La1.6−xNd0.4SrxCuO4 and related 214 cuprates. Phys. Rev. Lett. 58, 8760–8772 (1998).

    CAS  Google Scholar 

  29. Chatterjee, S. & Sachdev, S. Insulators and metals with topological order and discrete symmetry breaking. Phys. Rev. B 95, 205133 (2017).

    Article  ADS  Google Scholar 

  30. Nie, L., Tarjus, G. & Kivelson, S. A. Quenched disorder and vestigial nematicity in the pseudogap regime of the cuprates. Proc. Natl Acad. Sci. USA 111, 7980–7985 (2014).

    Article  ADS  CAS  Google Scholar 

  31. Varma, C. M. Quantum-critical fluctuations in 2D metals: strange metals and superconductivity in antiferromagnets and in cuprates. Rep. Prog. Phys. 79, 082501 (2016).

    Article  ADS  Google Scholar 

  32. Komiya, S. & Tsukada, I. Doping evolution of the electronic specific heat coefficient in slightly-doped La2−xSrxCuO4 single crystals. J. Phys. Conf. Ser. 150, 052118 (2009).

    Article  Google Scholar 

  33. Ghamaty, S. et al. Low temperature specific heat of Ln2CuO4 (Ln = Pr, Nd, Sm, Eu and Gd) and Nd1.85M0.15CuO4−y (M = Ce and Th). Physica C 160, 217–222 (1989).

    Article  ADS  CAS  Google Scholar 

  34. Yoshida, T. et al. Systematic doping evolution of the underlying Fermi surface of La2−xSrxCuO4. Phys. Rev. B 74, 224510 (2006).

    Article  ADS  Google Scholar 

  35. Cyr-Choinière, O. et al. Pseudogap temperature T of cuprate superconductors from the Nernst effect. Phys. Rev. B 97, 064502 (2018).

    Article  ADS  Google Scholar 

  36. Markiewicz, R. S. et al. One-band tight-binding model parametrization of the high-T c cuprates including the effect of k z dispersion. Phys. Rev. B 72, 054519 (2005).

    Article  ADS  Google Scholar 

  37. Matt, C. E. et al. Direct observation of orbital hybridisation in a cuprate superconductor. Nat. Commun. 9, 972 (2018).

    Article  ADS  CAS  Google Scholar 

  38. Verret, S. et al. Phenomenological theories of the low-temperature pseudogap: Hall number, specific heat, and Seebeck coefficient. Phys. Rev. B 96, 125139 (2017).

    Article  ADS  Google Scholar 

  39. Daou, R. et al. Linear temperature dependence of resistivity and change in the Fermi surface at the pseudogap critical point of a high-T c superconductor. Nat. Phys. 5, 31–34 (2009).

    Article  CAS  Google Scholar 

  40. Horio, M. et al. Three-dimensional Fermi surface of overdoped La-based cuprates. Phys. Rev. Lett. 121, 077004 (2018).

    Article  ADS  CAS  Google Scholar 

Download references

Acknowledgements

We thank J. Chang, M. Horio, M.-H. Julien, S. Kivelson, R. Markiewicz, C. Proust, B. Ramshaw, S. Sachdev, A. Sacuto, J. Tallon, A.-M. Tremblay and C. Varma for discussions. C.M. and T.K. acknowledge support from the Laboratoire d’excellence LANEF (ANR-10-LABX-51-01) and the Laboratoire National des Champs Magnétiques Intenses (LNCMI) in Grenoble. J.K. was supported by the Slovak Research and Development Agency under grant number APVV-16-0372. L.T. acknowledges support from the Canadian Institute for Advanced Research (CIFAR) and funding from the Natural Sciences and Engineering Research Council of Canada (NSERC; PIN: 123817), the Fonds de recherche du Québec–Nature et Technologies (FRQNT), the Canada Foundation for Innovation (CFI), and a Canada Research Chair. This research was undertaken thanks in part to funding from the Canada First Research Excellence Fund. Part of this work was funded by the Gordon and Betty Moore Foundation’s EPiQS Initiative (grant GBMF5306 to L.T.). J.-S.Z. was supported by NSF MRSEC under Cooperative Agreement Number DMR-1720595 in the US. H.T. acknowledges MEXT Japan for a Grant-in-Aid for Scientific Research.

Author information

Authors and Affiliations

Authors

Contributions

B.M., C.G., C.M. and T.K. performed the specific heat measurements. J.K. assisted in the development of the measurement technique. B.M., S.B. and N.D.-L. characterized the samples via resistivity and magnetization measurements. J.-S.Z. prepared the Nd-LSCO crystals. S.P., T.T. and H.T. prepared the Eu-LSCO crystals. Q.M., M.D., H.A.D. and B.D.G. prepared the Nd-LSCO polycrystalline samples and measured the phonon energies of Nd-LSCO with neutron scattering. S.V. calculated the specific heat of Nd-LSCO from its three-dimensional band structure. B.M., C.M., L.T. and T.K. wrote the manuscript, in consultation with all authors. L.T. and T.K. designed the study and supervised the project.

Corresponding authors

Correspondence to L. Taillefer or T. Klein.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended Data Fig. 1 Temperature–doping phase diagram.

Temperature-doping phase diagram of LSCO (black), Nd-LSCO (red) and Eu-LSCO (green), showing the boundary of the phase of long-range commensurate antiferromagnetic order (TN, brown line), the pseudogap temperature T* (blue line) and the superconducting transition temperature Tc of LSCO (grey line) and Nd-LSCO (pink line). T* is detected in two transport properties: resistivity (Tρ, circles) and the Nernst effect (Tν, squares). The open triangles show T* detected by ARPES as the temperature below which the anti-nodal pseudogap opens, in LSCO (black) and Nd-LSCO (red). We see that Tν ≈ Tρ ≈ T*, within error bars. The pseudogap phase ends at a critical doping p* = 0.18 ± 0.01 in LSCO (black diamond) and p* = 0.23 ± 0.01 in Nd-LSCO (red diamond). Figure adapted from figure 10 in ref. 35 (American Physical Society) (error bars defined therein).

Extended Data Fig. 2 Characterization of our Eu-LSCO and Nd-LSCO samples.

a, Tc versus p for our Eu-LSCO samples. Tc is defined as the onset of the drop in the magnetization upon cooling. Error bars on Tc reflect the uncertainty in defining the onset of the drop in the magnetization. b, Same for our Nd-LSCO samples. c, ρ versus T in our Eu-LSCO samples with p = 0.21 (red) and p = 0.24 (blue), at H = 0 and H = 33 T (short section below 40 K). d, Same for our Nd-LSCO samples with p = 0.22 (red) and p = 0.24 (blue)2. The approximately linear ρ(T) as T → 0 at p = 0.24 (blue) shows that 0.24 is close to the critical point p* ≈ 0.23 in both materials. The large upturn in ρ(T) as T → 0 at p = 0.21 and p = 0.22 (red) shows that the pseudogap has opened in both materials (at p < 0.23).

Extended Data Fig. 3 Specific heat jump at Tc in Eu-LSCO.

a, C/T versus T in our Eu-LSCO sample with p = 0.24, at H = 0 (red) and H = 8 T (blue). The inset shows the difference between the two curves in the main panel (red). This is the difference between the superconducting-state C/T and the normal-state C/T. It reveals the jump at Tc, whose peak value, ΔC/Tc, is defined as drawn. The black curve is the magnetization of that sample. At p = 0.24, the bulk Tc = 10.5 ± 0.5 K (dashed line). b, As in a, for our sample with p = 0.21, at H = 0 (red) and H = 18 T (blue). At p = 0.21, the bulk Tc = 14.5 ± 0.5 K. c, As in a, for our sample with p = 0.16, at H = 0 (red) and H = 18 T (blue). At p = 0.16, the bulk Tc = 11.5 ± 0.5 K. d, As in a, for our sample with p = 0.11, at H = 0 (red) and H = 8 T (blue). At p = 0.11, the bulk Tc = 5.0 ± 0.5 K. e, Plot of ΔC/Tc versus Cel/T at T = 0.5 K, the latter being obtained from Fig. 3b (red squares). The error bar on ΔC/Tc comes mostly from the uncertainty on defining the baseline above Tc. The dashed line is a linear fit through the first three data points. The tenfold increase in ΔC/Tc from p = 0.11 to p = 0.24 is independent evidence of a similar increase in Cel/T.

Extended Data Fig. 4 Specific heat as a function of magnetic field.

a, C/T versus H in our Eu-LSCO samples with p = 0.21 (orange) and p = 0.24 (red), at T = 2 K. The upper critical field above which there is no remaining superconductivity is Hc2 = 15 T at p = 0.21 and Hc2 = 9 T at p = 0.24. Note that for p = 0.24, C/T has reached 99% of its normal state value by 8 T. b, C/T versus H in our Nd-LSCO sample with p = 0.23, at T = 2 K, in a semi-log plot. The dashed line shows the expected field dependence of the Schottky contribution associated with Nd ions (Cmag, dashed line). The data are independent of field above H ≈ 9 T. Dotted lines are horizontal.

Extended Data Fig. 5 Specific heat of Eu-LSCO and Nd-LSCO down to base temperature.

a, Specific heat of the four Eu-LSCO samples of Fig. 2a measured in a field H = 8 T, down to 0.4 K. The rapid rise below 1 K is a nuclear Schottky anomaly (Cnuclear). b, Difference between the measured C/T of a and a constant term γ, plotted for each doping as a function of temperature, on a log–log plot (γ = 2.8 and 4.2 mJ K−2 mol−1, at p = 0.11 and 0.16, respectively). The line marked T2 shows that the data at p = 0.11 and p = 0.16 obey ΔC = βT3 in the range from 1.5 K to about 5 K. The line marked T–3 shows that the data at p = 0.11 and p = 0.16 obey ΔC ≈ T–2 below 1 K, as expected for the upper tail of a Schottky anomaly. The ΔC curve at p = 0.16, ΔC(p = 0.16; T), therefore constitutes the non-electronic, and weakly doping-dependent, background for C(T) in Eu-LSCO, made of phonon and Schottky contributions. c, Specific heat of our Nd-LSCO crystal with p = 0.12, plotted as C/T versus T at three different fields, as indicated. At H = 0 (green), we see the large Schottky anomaly associated with Nd ions, varying as Cmag ≈ T–2 at low T. At H = 8 T (red), it is pushed up above 2 K; at H = 18 T (blue), above 5 K. The line is a fit of the 18 T data to γ + βT2 below 5 K. d, Specific heat of the four Nd-LSCO samples of Fig. 2c, plotted as C/T versus T. Below the vertical dashed line, we show low-temperature data taken at H = 8 T on three of these same samples. (See Supplementary Fig. 1 for the complete set of dopings and Extended Data Fig. 6 for further analysis and discussion.).

Extended Data Fig. 6 Analysis of specific heat data and doping dependence of β.

a, Raw data for Eu-LSCO at p = 0.11 (blue), 0.16 (green) and 0.24 (red), plotted as C/T versus logT. The width of the pale band tracking each curve is the uncertainty on the absolute measurement of C (±4%). The solid green line is a fit to Cnuclear ≈ T–2 for the p = 0.16 data. b, Same three curves as in a, from which the same Schottky anomaly, Cnuclear/T, the green line in a, has been subtracted. The straight dotted lines show that (CCnuclear)/T is flat as T → 0 for p = 0.11 and 0.16, while it rises as log(1/T) for p = 0.24. The solid green line is a fit of the green curve at p = 0.16 to (CCnuclear)/T = γ + Cph/T up to 10 K, where Cph/T = βT2 + δT4 is the phonon contribution. c, Same three curves as in b, from which the same phonon contribution Cph/T, the green line in b, has been subtracted. We see that within error bars the resulting Cel/T is constant up to 10 K for p = 0.11 (blue) and 0.16 (green), while it varies as log(1/T) up to 10 K for p = 0.24 (red). The dotted lines are a linear fit to the data. d, Doping dependence of the phonon specific heat parameter β, in Cph/T = βT2 + δT4, obtained from a fit to (CCnuclear)/T = γ + Cph/T up to 10 K, for Eu-LSCO crystals (dark blue squares) and Nd-LSCO crystals (red dots). For crystals, the error bars are the sum of two uncertainties: on the magnitude of the raw data, defined in the legend of Fig. 3, plus on the fitting procedure to extract β, described in a, b and c. For Nd-LSCO powders (orange dots), the values are obtained from Extended Data Figs. 7, 8; the error bars are the sum of two uncertainties: on the magnitude of the raw data, defined in the legend of Fig. 3, plus on the fitting procedure to extract β, described in the legend of Extended Data Figs. 7, 8. The black diamonds are 1/Eph3 (right axis), where Eph is the phonon energy (top of the acoustic branch) measured by neutron scattering on three Nd-LSCO single crystals: Eph = 14.6, 14.7 and 14.8 meV at p = 0.12, 0.19 and 0.24, respectively (measurements performed by Q.M., M.D., H.A.D. and B.D.G.). We see that Eph varies very little with doping, and β ≈ 1/Eph3, justifying our assumption that Cph(T) does not change appreciably between p = 0.11 and p = 0.25.

Extended Data Fig. 7 Specific heat of our polycrystalline samples at p = 0.07 and 0.12.

a, C/T versus T for Nd-LSCO p = 0.12, comparing raw data on crystal and powder, as indicated. The solid red line is a fit to the crystal data, consisting of the sum of three contributions, plotted below: electrons (dash-dotted), phonons (dashed) and Schottky (Cmag ≈ T–2, dotted). b, Specific heat data for our powders with p = 0.07 and 0.12, at H = 0, from which the Schottky term (Cmag) has been subtracted, plotted as (C(H = 0) – Cmag)/T versus T2. The dashed lines are linear fits to the data (γ + βT2). For p = 0.12, the fit yields γ ≈ 4.0 mJ K−2 mol−1, in reasonable agreement with the value obtained by applying 18 T to suppress the Schottky anomaly in our Nd-LSCO crystal with p = 0.12 (Fig. 2c), namely γ = 3.6 ± 0.5 mJ K−2 mol−1 (Fig. 3b). For the two powder samples, the γ values are plotted in Fig. 3b (purple dots) and the β values are plotted in Extended Data Fig. 6d (orange dots). The value of (C(H = 0) – Cmag)/T extrapolated at T = 2 K is plotted in Fig. 3a (open red circles).

Extended Data Fig. 8 Specific heat of our polycrystalline samples at p = 0.27, 0.36 and 0.40.

a, Raw specific heat data for our powders with p = 0.27, 0.36 and 0.40, at H = 0. b, The data in a from which the magnetic Schottky term (Cmag = A/T2) has been subtracted, plotted as (CCmag)/T versus T2. The dotted lines are linear fits to the data (γ + βT2). For the three powder samples, the γ values are plotted in Fig. 3b (purple circles) and the β values are plotted in Extended Data Fig. 6d (orange circles). The value of (C(H = 0) – Cmag)/T extrapolated at T = 2 K is plotted in Fig. 3a (open red circles). c, The data in b from which the phonon term (Cph = βT3) has been subtracted, plotted as (CCmagCph)/T versus T2. All three curves are seen to be constant. d, The data in b, compared to raw single-crystal data at p = 0.24 for Eu-LSCO (orange) and Nd-LSCO (red), taken at H = 18 T. (The two raw curves agree beautifully at T < 5 K, where Cmag in Nd-LSCO is negligible; Extended Data Fig. 5c.) The upper dotted line is a linear guide to the eye showing that the Eu-LSCO curve deviates from linearity at low temperature, unlike the curves at p = 0.27, 0.36 and 0.40.

Extended Data Fig. 9 Calculated specific heat from the Nd-LSCO band structure.

Comparison of the measured specific heat of Nd-LSCO (red; Fig. 3b, with error bars defined in the legend of Fig. 3b) and the specific heat calculated for the band structure of Nd-LSCO (blue; see Methods), with the van Hove point pvHs set to be at p*. The calculations include the three-dimensional dispersion in the Fermi surface (along the c axis) and the disorder scattering, both consistent with the measured properties of our Nd-LSCO samples, namely their resistivity anisotropy (ρc/ρa) and their residual resistivity (see Methods). We see that while the van Hove singularity can give rise to a cusp-like peak at pvHs (top left panel) and a log(1/T) dependence of C/T at pvHs in a perfectly two-dimensional system with no disorder (lower left panel), these features inevitably disappear when the considerable three-dimensional dispersion of the real material and the high disorder of the real samples are included (right panels). The calculations only quantify what is naturally expected: the rise in specific heat due to the van Hove singularity is cut off when kBT < ħΓ, where Γ is the scattering rate, or when kBT < tz, where tz is the c-axis hopping parameter. The fact that we see C/T continuing to increase down to 0.5 K (lower right panel) excludes the van Hove singularity as the underlying mechanism.

Extended Data Fig. 10 Comparing with data on non-superconducting LSCO.

a, Normal-state electronic specific heat Cel of Eu-LSCO (squares; from Fig. 2b) and Nd-LSCO (circles; from Fig. 2d), at T = 0.5 K (red), 2 K (blue) and 10 K (green), plotted as Cel /T versus p. (At p = 0.08, 0.11 and 0.16, the red and green squares are split apart slightly so they can both be seen.) Open symbols are extrapolated values (dashed lines in Fig. 2b, d). Data on Nd-LSCO at p = 0.07, 0.12, 0.27, 0.36 and 0.40 (purple) are γ values obtained on polycrystalline samples, as described in Extended Data Figs. 7, 8. The black, purple and red data points are taken from Fig. 3b, with error bars defined in the legend of Fig. 3b. Error bars on the blue and green data points are defined in the same way as for the red data points (see legend of Fig. 3b). We also include γ for non-superconducting LSCO from published work (diamonds), obtained by extrapolating C/T = γ + βT2 to T = 0 from data below 10 K (p < 0.06, ref. 32; p = 0.33, ref. 13). The vertical dashed line marks the pseudogap critical point p* in Nd-LSCO (Extended Data Fig. 1). All solid lines are a guide to the eye. b, Comparison of Cel/T versus p in our samples of Eu-LSCO and Nd-LSCO at T = 10 K (green squares and circles in a) with published data on non-superconducting LSCO (diamonds). Open diamonds are γ measured in single crystals of LSCO at dopings where there is no superconductivity (p = 0.33, ref. 13; p < 0.06, ref. 32; remainder17). Solid diamonds are data from powders made non-superconducting by Zn substitution17; γ values are obtained from fits to C/T = γ + βT2 between about 4 K and about 8 K. We see that these early data on LSCO are quantitatively consistent with our data on Eu-LSCO and Nd-LSCO, apart from a downward shift in the position of the peak, consistent with a lower p* in LSCO (Extended Data Fig. 1). Lines are a guide to the eye.

Supplementary information

Supplementary Figures

This file contains Supplementary Figures 1-3

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Michon, B., Girod, C., Badoux, S. et al. Thermodynamic signatures of quantum criticality in cuprate superconductors. Nature 567, 218–222 (2019). https://doi.org/10.1038/s41586-019-0932-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-019-0932-x

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing