Main

In the phase diagram of several organic, heavy-fermion and iron-based superconductors, superconductivity forms a dome around the quantum critical point (QCP), where a phase of antiferromagnetic order ends. The spin fluctuations associated with that QCP are believed to cause both pairing and scattering4. The scattering is anomalous in that it produces a resistivity with a linear temperature dependence as T → 0, instead of the conventional T2 dependence of a Fermi liquid3,5,8. In hole-doped cuprates, the same features—a Tc dome and a T-linear resistivity—are also observed1, but not at the critical doping where the Néel temperature TN for the onset of long-range antiferromagnetic order vanishes. Instead, they are observed near the doping level p* where the pseudogap phase ends (Extended Data Fig. 1) and the essential nature of this phase remains unknown—it is the central enigma of cuprates.

The thermodynamic signature of a QCP is a diverging electronic mass. For an antiferromagnetic QCP in two dimensions, for example, the mass is expected to follow m*  log(1/|x − x*|) and the specific heat C to follow C/T  log(1/|x − x*|), as the system is moved towards its QCP at x* by varying some tuning parameter x, such as pressure or concentration3. This is what is observed in the iron-based superconductor BaFe2(As1−xPx)2 at its antiferromagnetic QCP (tuned by P concentration), both in the carrier mass m* measured via quantum oscillations and in the electronic specific heat estimated from the jump at Tc5,7. At x = x*, we expect that C/T  log(1/T), as observed in the heavy-fermion metal CeCu6−xAux at its antiferromagnetic QCP (tuned by Au concentration)3,6. This logarithmic divergence of C/T as T → 0 is the true sign of an energy scale that vanishes at x*.

In the cuprate YBa2Cu3Oy (YBCO), a study of quantum oscillations9 has revealed an increase in m* as p approaches the pseudogap critical doping p* ≈ 0.19, suggesting that p* may be a QCP. The low value of m* at p = 0.10–0.12 is consistent with specific heat data10 (modulo a residual value attributed to CuO chains) and its increase versus p is consistent with the dramatic increase in the magnitude of the specific heat jump at Tc—that is, ΔC/Tc11,12—which reaches its maximal value at p* (Fig. 1). However, to conclude that there is quantum criticality, what is missing is a direct measurement of the normal-state specific heat of a cuprate as T → 0, across p*. This is what we report here.

Fig. 1: Specific heat data in various cuprates.
figure 1

Electronic specific heat coefficient γ (= Cel/T at T → 0) in non-superconducting Tl-2201 at p = 0.33 ± 0.02 (solid orange diamond, left axis)18 and in the field-induced normal state of YBCO at p = 0.10, 0.11 and 0.12 (solid blue circles, left axis10; a background zero-field component attributed to CuO chains is removed from these γ values). Also shown are the values of γ obtained from the effective mass m* measured by quantum oscillations in YBCO at 0.08 < p < 0.16 (open blue circles, left axis)9, Tl-2201 at p = 0.29 ± 0.02 (open orange diamond, left axis)19 and Hg-1201 at p ≈ 0.1 (open green square, left axis)22. The jump in specific heat at Tc, ΔC/Tc (defined in Extended Data Fig. 3), measured in Ca-doped YBCO (open blue triangles)12 is plotted for comparison (right axis). (We note that ΔC/Tc is not a direct measurement of γ, and that it is affected by pair breaking due to disorder, more so as Tc decreases.) The doping p is taken from the original papers9,10,12,19,22, except for Tl-2201 at p = 0.33, where it is estimated from the fact that the sample is non-superconducting18 and that the the Tc dome ends at p = 0.31 (ref. 19). The units are expressed per mole of planar Cu. The dashed blue line marks the pseudogap critical point p* in YBCO. The solid line is a guide to the eye. The vertical and horizontal error bars are taken from the original literature, except for the horizontal error bars on the two Tl-2201 points, which reflect the uncertainty in our estimates of the doping for those samples.

We have measured C(T) in the closely related cuprate materials La1.8−xEu0.2SrxCuO4 (Eu-LSCO) and La1.6−xNd0.4SrxCuO4 (Nd-LSCO). Because of their low Tc (<20 K), superconductivity can be fully suppressed with a readily accessible magnetic field (about 15 T). The two materials have the same crystal structure and phase diagram, with very similar boundaries for the pseudogap phase, T*(p) (Extended Data Fig. 1) and superconducting phase, Tc(p) (Extended Data Fig. 2). In Nd-LSCO, resistivity and Hall effect measurements yield p* = 0.23 ± 0.01 (ref. 2). At p = 0.24, ρ(T) is linear in T as T → 0, the signature of quantum criticality in electrical transport. A very similar behaviour is observed in Eu-LSCO (Extended Data Fig. 2). We deduce that p* ≈ 0.23 also in Eu-LSCO.

The specific heat of five crystals of Eu-LSCO and seven crystals of Nd-LSCO was measured below 10 K (Supplementary Fig. 1). The normal-state specific heat is obtained by applying a magnetic field of either H = 8 T or 18 T (Extended Data Figs. 3, 4). In Fig. 2a, c, we show normal-state data at H = 18 T in Eu-LSCO and Nd-LSCO, respectively, plotted as C/T versus T2. We note that the magnetic moment on the Nd produces a Schottky anomaly in the specific heat of Nd-LSCO (Cmag), not present in Eu-LSCO (since Eu has no moment). At H = 0, this anomaly leads to a large increase of C/T at low T. However, a magnetic field moves this Schottky contribution to higher temperature, so that it becomes negligible below about 5 K at 18 T (Extended Data Fig. 5c).

Fig. 2: Specific heat of Eu-LSCO and Nd-LSCO.
figure 2

a, Specific heat of Eu-LSCO measured in a field H = 18 T, plotted as C/T versus T2, for four different dopings, as indicated. The dashed line is a linear fit to the data at p = 0.11 (blue) for T < 5 K; it yields γ = 2.8 mJ K−2 mol−1 and β = 0.22 mJ K−4 mol−1, where C/T = γ + βT2. (Data for our five crystals of Eu-LSCO are displayed in Supplementary Fig. 1.) b, Electronic specific heat of Eu-LSCO, Cel(T), obtained as described in Methods, plotted as Cel/T versus logT, from data at H = 8 T (p = 0.11 and 0.24) and at H = 18 T (p = 0.21). (The dashed line is a linear extrapolation of the p = 0.21 data.) At p = 0.11, Cel/T = γ, a constant, while at p = 0.24 ≈ p*, Cel/T ≈ log(1/T), the thermodynamic signature of a quantum critical point. c, Specific heat of Nd-LSCO measured in a field H = 18 T, plotted as C/T versus T2, for four dopings, as indicated. (Data for our seven crystals of Nd-LSCO are displayed in Supplementary Fig. 1.) The dashed line is a linear fit to the data at p = 0.12, C/T = γ + βT2 (below 5 K), giving γ = 3.6 mJ K−2 mol−1 and β = 0.215 mJ K−4 mol−1. d, Electronic specific heat of Nd-LSCO, obtained as described in Methods, plotted as Cel/T versus logT. This is done separately for the data below (H = 8 T) and above (H = 18 T) the dashed line. At p = 0.24 ≈ p*, Cel/T ≈ log(1/T), just as in Eu-LSCO.

In Fig. 3a, we plot the raw data at T = 2 K (and H = 18 T), C/T versus p, for all 12 crystals. The 12 data points fall on the same smooth curve, demonstrating a high level of quantitative fidelity and reproducibility. Because the magnetic and nuclear Schottky contributions (Cmag and Cnuclear; see Methods) are both negligible at 2 K, we have C = Cel + Cph, the sum of the electronic and phononic contributions. In Eu-LSCO at p = 0.11 and p = 0.16, the data obey C/T = γ + βT2 below T ≈ 5 K (Fig. 2a). The residual linear term γ is electronic and the second term is due to phonons, with β ≈ 0.22 mJ K−4 mol for both dopings. The same is true in Nd-LSCO at p = 0.12 (Fig. 2c) and p = 0.15 (Supplementary Fig. 1c), again with β ≈ 0.22 mJ K−4 mol for both dopings. In Extended Data Fig. 6d, we plot β versus p for all our samples. We see that β is constant between p = 0.1 and p = 0.4, within error bars. As a result, Cph/T is a small constant background in that range (see dashed lines in Fig. 3a).

Fig. 3: Doping dependence of the electronic specific heat.
figure 3

a, Raw data for C/T in Eu-LSCO (squares) and Nd-LSCO (circles) at T = 2 K and H = 18 T (solid red symbols) and at T = 1 K and H = 8 T (solid orange squares). We also include data points for non-superconducting LSCO at H = 0 (diamonds), at T = 2 K (red) and T = 1 K (orange), at p = 0.33 (ref. 13). Open circles are (C – Cmag)/T at T = 2 K, obtained on polycrystalline samples of Nd-LSCO (Extended Data Figs. 7, 8). The dashed lines indicate the phonon contribution, Cph/T = βT2, to the specific heat of Nd-LSCO, at T = 2 K (red) and T = 1 K (orange) (Extended Data Fig. 6d). Error bars are explained inMethods. Horizontal error bars reflect the uncertainty in the doping p for each sample, as reported in ref. 2 for the Nd-LSCO crystals and ref. 13 for the LSCO sample. For the Eu-LSCO crystals, the uncertainty is taken to be ±0.005; for the Nd-LSCO powders, it is ±0.01. Vertical error bars reflect the level of measurement reproducibility and the relative importance of addenda and sample contributions. For the Nd-LSCO powders, an additional uncertainty comes from having to subtract the magnetic Schottky anomaly. b, Normal-state electronic specific heat Cel of Eu-LSCO (squares) and Nd-LSCO (red circles, crystals), at T = 0.5 K, plotted as Cel/T versus p (see Extended Data Fig. 10a for details). We also include γ for Nd-LSCO (purple circles, polycrystals; see Extended Data Figs. 7, 8) and non-superconducting LSCO from published work (diamonds; p < 0.06, ref. 32; p = 0.33, ref. 13), obtained by extrapolating C/T = γ + βT2 to T = 0 from data below 10 K. The jump in specific heat at Tc, ΔC/Tc, measured in Eu-LSCO (blue triangles; Extended Data Fig. 3) is plotted for comparison (right axis). We see that ΔC/Tc and Cel/T scale nicely. The dashed line marks the pseudogap critical point p* in Nd-LSCO (Extended Data Fig. 1). The solid lines are a guide to the eye. The error bars are the same as in a. The vertical error bars on the blue triangles also include an uncertainty in estimating the specific heat jump ΔC relative to its background above Tc (see Extended Data Fig. 3).

To investigate what happens above p*, we measured the specific heat of three Nd-LSCO samples with p = 0.27, 0.36 and 0.40—dopings at which the material is no longer superconducting. Because it is difficult to grow single crystals at such high doping, the samples are polycrystalline powder. As a result, a field cannot be used to shift the magnetic Schottky anomaly (Cmag) up in temperature (because the effect of a field is highly anisotropic). Nevertheless, we can reliably subtract Cmag from C and obtain Cel + Cph, as demonstrated in Extended Data Figs. 7, 8 for our five polycrystalline samples. In all cases, the zero-field data obey (C – Cmag)/T = γ + βT2, with values of γ and β that are consistent with those obtained in single crystals. In Extended Data Fig. 7a, we show that polycrystalline and single-crystal samples yield exactly the same data, at the same doping (here p = 0.12). Plotting the values of (C – Cmag)/T at T = 2 K in Fig. 3a, we see that C/T drops by a factor of three in going from p* up to p = 0.4. Our values in Nd-LSCO are consistent with published data on La2−xSrxCuO4 (LSCO) at p = 0.33 (ref. 13; Fig. 3a).

The thermodynamic signature of the pseudogap critical point in Eu-LSCO and Nd-LSCO is therefore seen to be a huge peak in Cel/T at p*, at low temperature, not just a drop below p* (Fig. 3a). There are two standard explanations for such a peak: a van Hove singularity in the band structure and a quantum critical point. Because there is indeed a van Hove singularity in Nd-LSCO at p = pvHs ≈ p* (ref. 14), we have considered the first scenario carefully. Fortunately, the band structure of Nd-LSCO is well known and simple14, so reliable calculations can be performed. In a perfectly clean two-dimensional metal, Cel/T versus p does show a sharp cusp at p = pvHs in the T = 0 limit (Extended Data Fig. 9). However, when the actual three-dimensional dispersion of the real Fermi surface and the actual level of disorder scattering of the real samples are included, the peak due to the van Hove singularity is much reduced and broadened (Extended Data Fig. 9). The large, sharp peak we observe is therefore clearly caused by electronic effects beyond the band structure.

We now turn to the temperature dependence of Cel. In Fig. 2a, we see that C/T at p = 0.24 deviates strongly from the γ + βT2 behaviour as T → 0. To investigate this deviation in detail, measurements were carried out in a 3He refrigerator with a field of 8 T, just enough to reach the normal state at p = 0.11, 0.16 and 0.24 (Extended Data Fig. 4). The data are plotted as C/T versus T in Extended Data Fig. 5a. Below 1 K, we observe a nuclear Schottky anomaly (Cnuclear), which rises as C/T  T–3. Because Cnuclear and Cph are both almost independent of doping, we can readily subtract them from the measured C and obtain the purely electronic contribution Cel (see Methods). The result is plotted as Cel/T versus logT in Fig. 2b. At p = 0.11, we see that Cel/T is constant, with Cel/T = γ = 2.8 mJ K−2 mol. In contrast, at p = 0.24, we see that Cel/T grows linearly from about 10 K all the way down to 0.5 K, our lowest temperature. We arrive at the key finding that Cel/T  log(1/T) at p ≈ p*.

The fact that Cel/T keeps increasing as T → 0 at p = 0.24 is confirmed directly by looking at ΔC/Tc, the jump in C/T at Tc, the two being linked by entropy balance. In Extended Data Fig. 3, we see that ΔC/Tc increases by a factor of ten in going from p = 0.11 to p = 0.24, in tandem with the increase in Cel/T at T = 0.5 K (Fig. 3b).

In Extended Data Fig. 5d (and Supplementary Fig. 1c), we show Nd-LSCO data obtained in the 3He refrigerator at H = 8 T, a field sufficient to suppress superconductivity at p = 0.12, 0.22, 0.23, 0.24 and 0.25. As described in Methods, we obtain Cel(T) as we did for Eu-LSCO. The result is displayed in Fig. 2d (and Supplementary Fig. 1d), plotted as Cel/T versus logT. The Nd-LSCO data for Cel(T) are seen to be in excellent quantitative agreement with the Eu-LSCO data (Fig. 2b). In particular, they confirm our key finding that Cel/T  log(1/T) at p ≈ p*. We therefore find that the cuprates Eu-LSCO and Nd-LSCO exhibit the classic thermodynamic signature of a QCP, as observed in the antiferromagnetic heavy-fermion metals CeCu6−xAux (ref. 6), YbRh2Si2 (ref. 15) and CeCoIn5 (ref. 16).

In Fig. 3b, we plot the value of Cel/T for our 12 crystals as a function of doping, estimated at T = 0.5 K. We also plot the extrapolated γ values for our five polycrystalline samples of Nd-LSCO (p = 0.07, 0.12, 0.27, 0.36 and 0.40; Extended Data Figs. 7b, 8b) and the LSCO crystal at p = 0.33 (ref. 13). The huge cusp-like peak in Cel/T versus p is very similar to that seen in the iron-based superconductor BaFe2(As1−xPx)2 at its antiferromagnetic QCP, with C/T  log(1/|x − x*|) (ref. 7). Taken together, the p and T dependence of Cel provides compelling evidence for a QCP in Eu-LSCO and Nd-LSCO.

The strong similarity of our data on Nd-LSCO and Eu-LSCO with data on other cuprates indicates that the signatures of a QCP reported here are likely to be a generic feature of hole-doped cuprates. In Extended Data Fig. 10b, we compare our values of Cel/T in Eu-LSCO and Nd-LSCO at 10 K (Extended Data Fig. 10a) with γ values in LSCO obtained from fits to C/T = γ + βT2 between about 4 K and about 8 K, where C was measured on LSCO powders made non-superconducting by adding high levels of Zn impurities17. (We note that in this early work the temperature dependence of Cel/T was not investigated at very low temperature.) A clear peak in γ versus p is observed also in LSCO, similar to that found here in Nd-LSCO, although centred at a slightly lower doping—consistent with the lower p* in that material (Extended Data Fig. 1).

In no other cuprate has a direct measurement of the normal-state specific heat at low temperature been performed over a wide range including p*. To continue our survey of cuprates, we must therefore piece together different data from different materials. This is what we do in Fig. 1. Starting on the overdoped side, we have γ = 6.6 ± 1 mJ K−2 mol−1 in Tl2Ba2CuO6+δ (Tl-2201) at p ≈ 0.33 (ref. 18), not far from the value of γ = 7.6 ± 0.6 mJ K−2 mol−1 obtained from the effective mass measured by quantum oscillations, m* = (5.2 ± 0.4)me, in Tl-2201 at p = 0.29 ± 0.02 (ref. 19). Since those are similar to the values found in Nd-LSCO and in LSCO at p ≈ 0.3 (Fig. 3b), it is reasonable to suppose that other cuprates, such as YBCO, Bi2Sr2CaCu2O8 (Bi2212) and HgBa2CuO4+δ (Hg-1201), would also have γ ≈ 7 mJ K−2 mol−1 at p ≈ 0.3. (In none of these three materials has m*, γ or ΔC/Tc been measured beyond p ≈ 0.2312,20. Note that p* = 0.22 in Bi221221.) On the underdoped side, γ ≈ 3 mJ K−2 mol−1 in YBCO at p ≈ 0.1 (ref. 10) and quantum oscillations in YBCO9 and Hg-120122 at p ≈ 0.1 yield γ = 2.5 and 4.0 mJ K−2 mol−1, respectively (per mole of planar Cu), compared to γ = 2.8 and 3.6 mJ K−2 mol−1 in Eu-LSCO at p = 0.11 and Nd-LSCO at p = 0.12, respectively (Fig. 3b)—all in good agreement (Fig. 1).

We emphasize that γ in Tl-2201 at p ≈ 0.33 is only a factor 1.7 larger than γ in Hg-1201 at p ≈ 0.1. In other words, the opening of the pseudogap between the two has only reduced the density of states by a factor of about 1.7. This is a much smaller reduction than that observed in going from p* to p = 0.1 in any hole-doped cuprate11,12,17,20,23. If we assume that the reduction below p* in Tl-2201 is comparable to that in other cuprates, then γ in Tl-2201 must first rise from p ≈ 0.33 to p* before it falls below p*. This would imply that a peak in Cel/T at p* is a generic property of cuprates. (We note, however, that the only attempt to extract the doping dependence of γ in Tl-2201, based on zero-field data, found a constant γ (ref. 18). A direct measurement of the normal-state Cel is needed to resolve this apparent contradiction.)

In Fig. 1, we also plot the specific heat jump at Tc, ΔC/Tc, as a function of p, previously measured in Ca-doped YBCO12. We see that ΔC/Tc drops by a factor of about 10 in going from p* to p ≈ 0.1, consistent with the drop in condensation energy23. Since ΔC/Tc  γ, this implies that γ ≈ 25 mJ K−2 mol−1 at p* in YBCO (Fig. 1), comparable to our value of Cel /T = 22 mJ K−2 mol−1 at T = 0.5 K in Nd-LSCO at p* (Fig. 3b). This high value of γ in YBCO must then drop by a factor of about 3–4 above p*, if it is to reach the common value γ ≈ 7 mJ K−2 mol−1 at p ≈ 0.3 (Fig. 1). In summary, while further work is needed to establish that there is indeed a peak in Cel/T versus p for cuprates other than Nd-LSCO, Eu-LSCO and LSCO, most existing data are not incompatible with a universal peak in hole-doped cuprates.

Our finding that Cel/T peaks at p* is a change of paradigm in our understanding of cuprates—it reveals a mechanism of strong mass enhancement above p*, associated with a QCP at p*. Our observation of a continuous logarithmic increase of the electronic specific heat down to temperatures as low as 0.5 K raises the fundamental question of what energy scale, associated with most of the low-temperature entropy, might vanish at p*, and what corresponding length scale would then diverge at p*.

Given the remarkable similarity with the signatures of an antiferromagnetic QCP, as observed in heavy-fermion metals6 and iron-based superconductors7, it is tempting to attribute the quantum criticality of cuprates to antiferromagnetic spin correlations. However, unlike in electron-doped cuprates where the antiferromagnetic correlation length diverges as T → 0 (ref. 24) close to the critical doping x* where the Fermi surface is reconstructed and ρ(T) is T-linear25, there is no evidence of a diverging antiferromagnetic correlation length in hole-doped cuprates near p* (ref. 26).

In Nd-LSCO, incommensurate spin-density-wave order is observed by neutron diffraction27 to vanish with increasing p at p ≈ p*, but there is no divergent spin-density-wave correlation length or vanishing energy scale. Indeed, muon spin relaxation finds that the magnetic order in Nd-LSCO vanishes before p*, at p ≈ 0.20, which indicates that the order is not static at p = 0.20 (ref. 28), or above.

It may therefore be that the quantum criticality of hole-doped cuprates is of an entirely new kind, possibly involving topological order29. Alternatively, we may be dealing with two closely intertwined mechanisms: one mechanism for the pseudogap, possibly short-range antiferromagnetic correlations, and a second—separate but coupled—mechanism for the quantum criticality, possibly nematic order30 or current-loop order31.

Methods

Samples

Eu-LSCO

Single crystals of La2−yxEuySrxCuO4 (Eu-LSCO) were grown at the University of Tokyo with a Eu content y = 0.2, using a travelling-float-zone technique. Five samples were cut in the shape of small rectangular platelets, of typical dimensions 1 mm × 1 mm × 0.5 mm and mass of about 1 mg, from boules with nominal Sr concentrations x = 0.08, 0.11, 0.16, 0.21 and 0.24. The hole concentration p is given by p = x, with a maximal error bar of ±0.005. The critical temperature Tc of our samples, defined as the onset of the drop in the magnetization, is plotted in Extended Data Fig. 2a.

Nd-LSCO

Single crystals of La2−yxNdySrxCuO4 (Nd-LSCO) were grown at the University of Texas at Austin with a Nd content y = 0.4, using a travelling-float-zone technique. Seven samples were cut in the shape of small rectangular platelets, of typical dimensions 1 mm × 1 mm × 0.5 mm and mass of about 1 mg, from boules with nominal Sr concentrations x = 0.12, 0.15, 0.20, 0.22, 0.23, 0.24 and 0.25. The hole concentration p is given by p = x, with an error bar ±0.003, except for our sample with p = 0.24, for which the error bar is ±0.005 (ref. 2). The Tc values of all samples are plotted in Extended Data Fig. 2b. The rapid decrease in Tc from p = 0.22 to p = 0.25 provides a sensitive measure of the relative doping of samples in that range.

Powder samples of Nd-LSCO were produced at McMaster University with y = 0.4 and x = 0.07, 0.12, 0.27, 0.36 and 0.40, and shaped into sintered pellets. Samples with x = 0.27, 0.36 and 0.40 are not superconducting, and so their normal-state specific heat can be measured in zero field. The uncertainty on p from x is ±0.01.

Specific heat measurements

To perform the series of specific heat measurements reported in this article, we implemented an original alternating-current modulation method that leads to an absolute accuracy of ΔC/C ≈ ±4% with a relative sensitivity ΔC/C ≈ ±0.01% on samples whose masses are of the order of 1.0 mg (down to 0.1 mg). These figures are valid for the temperature range 0.5 K < T < 20 K and for the magnetic field range 0 T < H < 18 T.

Calorimetric setup

The calorimetric chips were prepared from bare Cernox chips (1010 for T < 2 K, 1050 for T > 2 K). First, a shallow groove is made in the central part with a wire saw to obtain two independent films; one used as a heater (of resistance Rh) and the other one as a thermometer (of resistance Rt; Supplementary Fig. 2). The split chip is then attached to a small copper ring with PtW (7%) wires, 25 μm or 50 μm in diameter and 1 mm to 2 mm in length, glued with a minute amount of Ag epoxy. The choice of wires is important since it defines the external thermal conductance and the frequency range where the measurements will have their optimal accuracy (between 0.5 Hz and 20 Hz). We used PtW wires since their thermal conductivity is insensitive to magnetic fields (<1% in 18 T).

Modulation technique

In the simplest approximation, when an alternative current Iac is applied at a frequency ω on the heater side, temperature oscillations Tac = Pac/(Ke + jC2ω) are induced at 2ω, where Ke is the external thermal conductance and C the total heat capacity. Introducing the phase φ of Tac relative to the power Pac, we obtain: C = Pacsin(−φ)/[2ω|Tac|]where Pac = Rh|Iac|2/2. These oscillations can then be measured by applying a direct current Idc across the thermometer: Vact(2ω) = [(dRt/dT)Tac(2ω)]Idc. The main drawback of this simple approach is that internal time corrections (due to finite thermal conductances between the different parts of the chips: thermometer, heater, substrate and sample) are not taken into account. To overcome this difficulty, especially at the lowest temperature where the thermal conductivity between the sensing layer of the Cernox and the sapphire substrate is the main limiting factor, we have also used the heater side to measure the temperature oscillations: Vach(3ω) = [(dRh/dT)Tac(2ω)]Iac(ω).

When all the internal conductances are larger than the external one Ke, the Tac measured on the thermometer side (at 2ω) must be the same as that measured on the heater side (at 3ω). Any difference points to a thermal gradient within the chips, which can then be minimized by adjusting the measurement parameters. This procedure improves the absolute accuracy and enables a much better estimate of the error bars.

Thermometry

The first step is to know precisely the temperature of the main heat sink on which the measuring chip is attached. This is achieved with commercial calibrated Cernox sensors 1010 and 1050, used respectively in the ranges 0.5–5 K and 1.5–20 K. The calibration has been further improved in-house with a superconductive fixed point device and a cerium–magnesium–nitrate (CMN) thermometer to reach an accuracy of ±1% in the absolute temperature, within the temperature range considered here. All thermometers have then been thoroughly calibrated in field, from 0.2 K to 4 K against a Ge sensor placed in a compensated area of an 18 T superconducting magnet and between 2 K and 20 K against a capacitor. The output of this protocol is a quintic bivariable (T and H) spline interpolation sheet, which was used to determine and control the temperature between 0.5 K and 20 K up to 18 T with a relative accuracy ΔT/T ≈ ±0.2%.

Addenda and test

To subtract the heat capacity of the sample mount (chip + a few micrograms of Apiezon grease used to glue the sample onto the back of the chip), the empty chip (with grease) was measured before each sample measurement. This background heat capacity is of the order of Cadd/T ≈ 5 nJ K−2 at 1 K (3 K) for the 1010 (1050) chips, which represents 10% to 50% of the heat capacity of the samples.

To benchmark our measurement system and technique, we measured a piece of ultrapure copper, of mass 1 mg, whose heat capacity at low T is comparable to that of our Eu-LSCO and Nd-LSCO samples. In zero magnetic field, the reproducibility between different runs and between different chips shows an absolute accuracy of at least ΔC/C ≈ ±3% compared to the values tabulated by the National Bureau of Standards (NBS), across the range from 0.5 K to 10 K (Supplementary Fig. 3). In magnetic fields up to 18 T, the apparent change ΔC/C versus H is very smooth and does not exceed 1%.

All factors considered, the absolute accuracy of each of the specific heat runs is estimated to be ΔC/C ≈ ±4%, or better. For 0.5 K < T < 5 K, the specific heat is dominated by the electronic contribution and ΔCel/Cel ≈ ΔC/C at T = 0.5 K and 1.0 K. But this error bar increases with temperature, owing to the rapid increase in the phonon contribution (and the Schottky term in Nd-LSCO), which makes the electronic term a smaller fraction of the total signal. The error bars plotted in Fig. 3b and Extended Data Fig. 10a show how this translates into an uncertainty on the absolute value of Cel/T for each sample separately, at each temperature.

Extracting the electronic specific heat

There are three contributions to the specific heat C of Eu-LSCO and Nd-LSCO: from electrons (Cel), phonons (Cph) and nuclei (Cnuclear). In Nd-LSCO, there is an additional contribution from the magnetic moment on the Nd3+ ions (Cmag). Because Eu ions are not magnetic, this term is absent in Eu-LSCO. To extract the electronic term of interest here, we proceed as outlined in Extended Data Fig. 6, and as described in detail below.

Nuclear Schottky term, Cnuclear

The nuclear hyperfine term is a Schottky anomaly peaked at very low temperature. Above the peak, Cnuclear dies off rapidly, as Cnuclear ≈ 1/T2. In Eu-LSCO and Nd-LSCO, Cnuclear is clearly visible in all samples as a rapid upturn below 1 K (Extended Data Fig. 5 and Supplementary Fig. 1). In a field of 8 T, Cnuclear becomes negligible above T ≈ 1 K (Extended Data Figs. 5b, 6a); in 18 T, above about 2 K (Extended Data Fig. 5c). By subtracting the raw data for C versus T in Eu-LSCO at p = 0.16 from the raw data C(T) at each other doping, we see that the rapid upturn below 1 K is almost entirely removed (Fig. 2b, Supplementary Fig. 1b, Extended Data Fig. 6c), at least down to 0.5 K. (Below T = 0.5 K, we do observe that the subtraction is not perfect, reflecting a small difference in the nuclear Schottky anomaly from sample to sample (within ±20% at 0.3 K), and this is why we report data only for T = 0.5 K and above.) This shows that Cnuclear is very weakly sample-dependent and doping-dependent. (The nuclear Schottky anomaly is believed to come from the Eu ions33, and the Eu content of all our Eu-LSCO samples is kept fixed.)

In summary, to remove Cnuclear from the measured C we can work at T = 1 K (in a field of 8 T) or at T = 2 K in 18 T, and higher temperatures. Below 1 K, we can remove Cnuclear reliably down to 0.5 K by subtracting a reference curve, namely p = 0.16 in Eu-LSCO and p = 0.12 in Nd-LSCO, since Cnuclear varies from sample to sample by less than ±5% at 0.5 K. In Extended Data Fig. 6b, we see the effect of such a subtraction: (CCnuclear)/T is constant below 1 K at p = 0.11 and p = 0.16, while it rises monotonically as T → 0 at p = 0.24. (Note that 8 T is enough to suppress superconductivity in Eu-LSCO down to the lowest temperature for all dopings except p = 0.21, where we use 18 T and are limited to T > 2 K (Fig. 2b).)

Magnetic Schottky term, Cmag

Compared to Eu-LSCO, the specific heat of Nd-LSCO has one additional contribution, Cmag, a Schottky peak due to the magnetic moment on the Nd3+ ions (Extended Data Fig. 4b). A field moves this magnetic Schottky peak up in temperature, so that Cmag becomes very small below 2 K in 8 T and very small below 6 K in 18 T (Extended Data Fig. 5c). We apply the same subtraction procedure as for Eu-LSCO, using our Nd-LSCO sample with p = 0.12 as the reference. (The subtraction procedure works well to remove both Cmag and Cnuclear because the Nd content of all our Nd-LSCO samples is kept fixed.)

At high doping (p > 0.25), single crystals are difficult to grow and we have therefore resorted to powder samples of Nd-LSCO, with p = 0.27, 0.36 and 0.40. Because these samples are not superconducting, the normal state specific heat can be measured in zero field. We note, however, that the magnetic Schottky term depends on the field direction relative to the c axis of the crystal structure. As a result, we could not use a field to suppress Cmag in the powder samples, made of micro-crystallites of all orientations. In Extended Data Fig. 7a, we compare directly the raw data from our powder sample at p = 0.12 with the raw data from our single-crystal sample at p = 0.12, both at H = 0. The two sets of raw data are seen to be in excellent agreement. In the single crystal, we remove Cmag by applying a field of 18 T; a fit to C/T = γ + βT2 below about 5 K yields γ = 3.6 mJ K−2 mol−1 and β = 0.215 mJ K−4 mol−1 (Fig. 2c). In the powder sample, we fit the zero-field raw data to a Schottky anomaly that goes as Cmag ≈ 1/T2 (in the range 3 K < T < 7 K). We subtract Cmag from the raw powder data and fit the difference to (C – Cmag)/T = γ + βT2 (Extended Data Fig. 7b). The resulting values of γ (4 ± 1 mJ K−2 mol−1) and β (0.225 ± 0.015 mJ K−4 mol−1) are in good agreement with those quoted above for the p = 0.12 single crystal (see Fig. 3b and Extended Data Fig. 6d).

Performing the same fit and subtraction to the zero-field raw powder data at p = 0.27, 0.36 and 0.40, reported in Extended Data Fig. 8a, yields curves that are roughly parallel when plotted as (C – Cmag)/T versus T2 (Extended Data Fig. 8b). A fit to (C – Cmag)/T = γ + βT2 yields the values of γ plotted in Fig. 3b (purple circles) and the values of β plotted in Extended Data Fig. 6d (orange circles). The fact that each curve in Extended Data Fig. 8b can be fitted to γ + βT2 shows that the electronic term Cel/T is constant in temperature. This is illustrated in Extended Data Fig. 8c, where we subtract the phonon term Cph = βT3.

In Extended Data Fig. 8d, we compare the three curves of Extended Data Fig. 8b to raw data on Nd-LSCO (red; from Fig. 2c) and Eu-LSCO (orange; from Fig. 2a) at p = 0.24, measured at H = 18 T, a field large enough to entirely remove superconductivity. In Eu-LSCO, there is no magnetic Schottky anomaly (Cmag = 0), and so the orange curve C/T at p = 0.24 can be compared directly to the curves (CCmag)/T versus T2 at p = 0.27, 0.36 and 0.40. We see that the orange curve is not straight, showing a clear upward deviation from linearity at low T. Such a deviation is not seen at p = 0.36 and 0.40. At p = 0.27, there might be a slight upward deviation, which would not be inconsistent with this doping being close to p*.

In the Nd-LSCO crystal at p = 0.24 (red curve in Extended Data Fig. 8d), a field of 18 T pushes the magnetic Schottky anomaly above about 5 K. Below about 5 K, the raw curve in Nd-LSCO is essentially identical to the raw curve in Eu-LSCO, confirming that the low-T upward deviation is present in Nd-LSCO also. In conclusion, the log(1/T) behaviour seen at p = 0.24 goes away as p increases above p*.

Phonon term, Cph

At 18 T, superconductivity is completely eliminated from all samples (Extended Data Fig. 4a). At T > 2 K, Cnuclear is negligible. In Nd-LSCO, Cmag is negligible below about 6 K. So in the range 2–6 K at 18 T, we have C = Cel + Cph. In Eu-LSCO at p = 0.11 and 0.16, and in Nd-LSCO at p = 0.12, the raw data in that T range are well described by C/T = γ + βT2 (Fig. 2a, c), with very similar values of β (Extended Data Fig. 6d). Alternatively, we can fit the 8 T data from those two samples of Eu-LSCO over the full range from T = 0.5 K to 10 K by using (CCnuclear)/T = γ + Cph/T, where Cph/T = βT2 + δT4, as shown in Extended Data Fig. 6. The two approaches yield very similar values of γ and β (identical within error bars). Fitting single-crystal data away from p* = 0.23, namely Eu-LSCO at p = 0.08, 0.11 and 0.16 in 8 T and Nd-LSCO at p = 0.12, 0.15 and 0.20 in 18 T, to (CCnuclear)/T = γ + βT2 + δT4 yields the values of β plotted in Extended Data Fig. 6d. For the five powder samples of Nd-LSCO, we fit the zero-field data to (CCmag)/T = γ + βT2 + δT4 (Extended Data Figs. 7b, 8b). In Extended Data Fig. 6d, we see that the doping dependence of β is weak, and in excellent agreement with the very weak doping dependence of the phonon energy, Eph, measured by neutron scattering (performed by Q.M., M.D., H.A.D. and B.D.G.), that is, β ≈ 1/Eph3.

Doping dependence of the electronic term, Cel versus p

In Extended Data Fig. 6a, we see that Cnuclear in 8 T is negligible at T = 1 K (and above). In Extended Data Fig. 6b, we see that Cph is negligible at T = 1 K (and below). Therefore the raw data at T = 1 K and H = 8 T, in Eu-LSCO, directly give the electronic specific heat at 1 K, that is, Cel = C to a very good approximation (except at p = 0.21, where 8 T is not enough to fully suppress superconductivity). The huge rise in C from p = 0.11 to p = 0.24 (Fig. 3a) is therefore entirely due to Cel.

At high doping, beyond p*, the Nd-LSCO powder data show a clear decrease in γ as p increases (Extended Data Fig. 8b). The result is therefore an unambiguous peak, at p = p*, in Cel versus p in the T = 0 limit (Fig. 3a). This is the first key signature of a QCP.

Temperature dependence of the electronic term, Cel versus T

Extended Data Fig. 6 shows that Cel/T at p = 0.16 is constant as a function of T, all T dependence being due to Cnuclear/T and Cph/T. The same is true at p = 0.11. At p = 0.24, however, there is an additional T dependence not due to Cnuclear or Cph, which comes from Cel(T). The simplest way to reveal this T dependence of Cel(T) is to subtract the raw curve at p = 0.16 from the raw curve at p = 0.24, as done in Fig. 2b. Or, equivalently, subtract the fit to (CCnuclearCph)/T performed on the p = 0.16 data, as done in Extended Data Fig. 6c. Both approaches yield a clean log(1/T) dependence for Cel/T at p = 0.24, from 0.5 K to 10 K (see dotted line in Extended Data Fig. 6c).

These approaches assume that Cnuclear is the same in the two samples, which is true to better than 5%, and also that Cph at p = 0.24 is the same as Cph at p = 0.16, which is true to better than 4%, as established by the neutron data: Eph changes from 14.6 meV at p = 0.12 to 14.8 meV at p = 0.24. In other words, within error bars, we find that Cel/T ≈ log(1/T) up to at least 10 K. This is the second key signature of a QCP.

In Extended Data Fig. 10a, we plot the value of Cel/T at T = 0.5 K, 2 K and 10 K, obtained using the first procedure of simply subtracting two raw curves, requiring no fitting at all. As we have just shown, the values of Cel/T obtained in this way are accurate and reliable, within the error bars.

The fact that Cel/T keeps increasing as T → 0 at p = 0.24 is confirmed directly by looking at ΔC/Tc, the jump in Cel/T at Tc, the two being linked by entropy balance. In Extended Data Fig. 3, we see that ΔC/Tc increases by a factor of ten in going from p = 0.11 to p = 0.24, in tandem with the increase in Cel/T at T = 0.5 K (Fig. 3b)—separate evidence from raw data (with no analysis or subtraction) that Cel/T at p = 0.24 must increase below 10 K.

In summary, the high level of quantitative consistency we find between the values of Cel/T obtained in our five crystals of Eu-LSCO and those obtained in our seven crystals of Nd-LSCO (Fig. 3b) is a strong validation of both the experimental technique and the data analysis. It confirms that the electronic specific heat Cel(T) we report is reproducible and accurate (within the quoted error bars). Furthermore, the excellent agreement between the values of Cel(T) plotted in Fig. 3b and the raw values of C(T) at T = 1 K and 2 K plotted in Fig. 3a confirms the fidelity of our analysis, as does the parallel growth observed in ΔC/Tc and Cel/T (Fig. 3b).

van Hove singularity

In hole-doped cuprates, the large Fermi surface in the overdoped regime undergoes a change of topology from hole-like to electron-like at some material-dependent critical doping pvHs. According to ARPES, this change of topology occurs between p = 0.15 and p = 0.22 in LSCO34 and between p = 0.20 and p = 0.24 in Nd-LSCO14, that is, close to p* in both cases (Extended Data Fig. 1; ref. 35). If the Fermi surface were strictly two-dimensional, this would, in the clean limit, lead to a van Hove singularity in the density of states, producing a cusp in Cel/T versus p at pvHs and a log(1/T) dependence of Cel/T as T → 0, analogous to the behaviour we report in Eu-LSCO and Nd-LSCO.

However, the change of Fermi-surface topology at pvHs ≈ p* cannot in fact be responsible for that behaviour, for two reasons: because of the strong three-dimensional character of the Fermi surface in Nd-LSCO (and Eu-LSCO), and because of the high level of disorder in our samples. Each of these two mechanisms broadens the van Hove singularity, removes the cusp versus p and cuts off the log(1/T) divergence at low T.

To be quantitative, we have calculated the specific heat of Nd-LSCO associated with its known band structure, using the following one-band model36:

$$\begin{array}{c}{\xi }_{{\boldsymbol{k}}}=-\mu -2t({\rm{c}}{\rm{o}}{\rm{s}}({k}_{x}a)+{\rm{c}}{\rm{o}}{\rm{s}}({k}_{y}b))-4t^{\prime} {\rm{c}}{\rm{o}}{\rm{s}}({k}_{x}a){\rm{c}}{\rm{o}}{\rm{s}}({k}_{y}b)-2t^{\prime\prime} ({\rm{c}}{\rm{o}}{\rm{s}}(2{k}_{x}a){\rm{c}}{\rm{o}}{\rm{s}}(2{k}_{y}b))\\ -2{t}_{z}{\rm{c}}{\rm{o}}{\rm{s}}({k}_{z}c/2){[{\rm{c}}{\rm{o}}{\rm{s}}({k}_{x}a)-{\rm{c}}{\rm{o}}{\rm{s}}({k}_{y}b)]}^{2}{\rm{c}}{\rm{o}}{\rm{s}}({k}_{x}a/2){\rm{c}}{\rm{o}}{\rm{s}}({k}_{y}b/2)\end{array}$$

with t = 0.189 meV, t′ = –0.17t, t′′ = 0.05t. Those parameters agree with the experimental band structure measured by ARPES37, and are such that the van Hove singularity is located at p = 0.23. The three-dimensional character is controlled by the interlayer hopping parameter tz. The doping is adjusted by tuning the chemical potential μ.

The temperature-dependent specific heat is given by38:

$${C}_{{\rm{el}}}\left(T\right)=\underset{-\infty }{\overset{\infty }{\int }}{\rm{d}}\epsilon \frac{\partial f\left(\epsilon \right)}{\partial T}\epsilon N\left(\epsilon \right)$$

with

$$N(\epsilon )=\frac{1}{V}\sum _{{\boldsymbol{k}}}\frac{1}{{\rm{\pi }}}\frac{\hbar /2\tau }{{(\epsilon -{\xi }_{{\boldsymbol{k}}})}^{2}+{(\hbar /2\tau )}^{2}}$$

where f(ε) is the Fermi–Dirac function, N(ε) is the density of states, and τ is the quasiparticle lifetime.

The calculated specific heat is shown in Extended Data Fig. 9 (blue), as a function of doping (at T → 0) on the top and as a function of temperature (at p = pvHs) on the bottom. In the left panels, we show the clean-limit two-dimensional result, with ħ/τ = 0 and tz = 0. In the middle panels, we show the effect of disorder, with ħ/τ = t/25, the value needed to produce the measured residual resistivity of our Nd-LSCO and Eu-LSCO samples at p = 0.24, namely ρ0 ≈ 30 μΩ cm (Extended Data Fig. 2). In the right panels, we further add the effect of three-dimensional dispersion, with tz = 0.13t, the value needed to produce the measured resistivity anisotropy of Nd-LSCO at p = 0.24, namely ρc/ρa ≈ 250 (ref. 39).

Direct comparison with our data shows that band structure effects associated with pvHs in Nd-LSCO do produce a broad background bump in Cel versus p, but they cannot account for the large and sharp peak we observe in Cel/T at T = 0.5 K (Extended Data Fig. 9). Moreover, the fact that we see Cel/T ≈ log(1/T) persisting down to 0.5 K at p = p* completely excludes a van Hove mechanism, which yields a constant Cel/T when kBT < ħ/τ or when kBT < tz, that is, below about 20 K in our samples (Extended Data Fig. 9).

Recent calculations of Cel/T versus p at T = 0 using the real three-dimensional electronic structure measured by ARPES in LSCO and Eu-LSCO40 yield curves that are essentially identical, quantitatively, with our own curves in Extended Data Fig. 9, showing again that the van Hove singularity cannot account for the large peak observed in Nd-LSCO, Eu-LSCO or LSCO. Even when the authors of ref. 40 assume no disorder (1/τ = 0), they find that the three-dimensional dispersion alone limits the increase in Cel/T from p = 0.40 to p = 0.24 to no more than a factor of 2, whereas our data show a fourfold increase (Fig. 3b).

We note that strong disorder does not alter the log(1/T) dependence of Cel/T coming from a QCP, as demonstrated by its persistence down to about 20 mK in samples of CeCu5.9Au0.1 (ref. 6) that have a residual resistivity ρ0 larger than that of our own samples at p = 0.24 (Extended Data Fig. 2).