This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Field-dependent specific heat of the canonical underdoped cuprate superconductor YBa2Cu4O8

Jeffery L. Tallon1,†, and John W. Loram2,‡ 1Robinson Research Institute, and MacDiarmid Institute for Advanced Materials and Nanotechnology, Victoria University of Wellington, P.O. Box 33436, Lower Hutt 5046, New Zealand. 2Cavendish Laboratory, Cambridge University, Cambridge CB3 0HE, United Kingdom.

The cuprate superconductor YBa2Cu4O8, in comparison with most other cuprates, has a stable stoichiometry, is largely free of defects and may be regarded as the canonical underdoped cuprate, displaying marked pseudogap behaviour and an associated distinct weakening of superconducting properties. This cuprate ‘pseudogap’ manifests as a partial gap in the electronic density of states at the Fermi level and is observed in most spectroscopic properties. After several decades of intensive study it is widely believed that the pseudogap closes, mean-field like, near a characteristic temperature, TT^{*}, which rises with decreasing hole concentration, pp. Here, we report extensive field-dependent electronic specific heat studies on YBa2Cu4O8 up to an unprecedented 400 K and show unequivocally that the pseudogap never closes, remaining open to at least 400 K where TT^{*} is typically presumed to be about 150 K. We show from the NMR Knight shift and the electronic entropy that the Wilson ratio is numerically consistent with a weakly-interacting Fermion system for the near-nodal states. And, from the field-dependent specific heat, we characterise the impact of fluctuations and impurity scattering on the thermodynamic properties.

Both the pseudogap Norman1 ; Timusk ; Tstar and the origins of superconductivity in the cuprates remain enigmatic and a source of continuing dispute, especially the former entrant . Still there is no consensus as to pseudogap’s phenomenology, at what doping the ground-state pseudogap ultimately vanishes, whether it really does close at TT^{*}, whether this closure might be a thermodynamic phase transition Bourges ; Xia ; Hashimoto ; He1 ; Shekhter ; Sato and whether it is causatively related to superconductivity entrant . The electronic specific heat captures the entire spectrum of low-energy excitations and in principle can adjudicate in all these matters. The key experimental challenge is to separate the electronic term from the much larger phonon term. In many previous experiments Loram2 ; Loram1 ; Loram4 and in the present report this can be done using a differential technique in which the specific heat is measured relative to a reference sample which, if closely related to the sample itself, automatically backs off most of the phonon contribution. The residual phonon contribution can be identified and removed by measuring a series of doping states in which the residual is found to scale linearly with the mass change of the doping process, usually changing oxygen content. Further details are given under Methods.

Electronic specific heat coefficient

Refer to caption
Figure 1: (a) The measured electronic specific heat coefficient γ(T)\gamma(T) for YBa2Cu4O8 for 0%, 2% and 4% planar Zn concentration at 0, 1, 2, 3, 5, 7, 9, 11 and 13 tesla. The same colour coding is used for each sample. Inset: the same data plotted versus ln(T)\ln(T) highlighting the impurity scattering term (dotted lines). (b) the same data with the impurity term subtracted. The black dashed curve is the fitted normal-state specific heat coefficient, γn\gamma_{\textrm{n}} which satisfies entropy balance.

The measured electronic specific heat coefficient, γ(T)CP(T)/T\gamma(T)\equiv C_{P}(T)/T is shown in Fig. 1(a) for YBa2Cu4O8 (Y124) with 0%, 2% and 4% planar Zn concentration at nine different applied fields, as annotated. As discussed later, the data extends to an unprecedented 400 K but we focus first on the transitions into the superconducting state. Because Zn substitutes only on the CuO2 planes and not on the chains Williams1 these compositions correspond to YBa2Cu4-yZnyO8 with y=y= 0, 0.04 and 0.08. It is immediately evident that TcT_{c} is rapidly suppressed by Zn substitution (as observed previously in other cuprates Tallon_scattering ) and along with this a rapid reduction in the jump height at TcT_{c}, Δγc\Delta\gamma_{\textrm{c}}. At the same time there is a rapid suppression in Δγc\Delta\gamma_{\textrm{c}} with applied field that becomes more extreme in the Zn-substituted samples. Further, at low TT, γ(T)\gamma(T) fans out to higher values with applied field in the ‘pure’ sample but not appreciably in the doped samples. This is due to the Volovik effect Volovik – the field-induced pairbreaking at the nodes due to Doppler shift of quasiparticle energies, as discussed below.

The most notable feature, however, is the low-TT upturn due to impurity scattering. The inset to Fig. 1(a) shows γ(T)\gamma(T) plotted versus ln(T)\ln(T) and this reveals a common underlying energy scale given by the convergence of the dashed lines at 38 K. The dashed lines are subtracted from the raw data to give the γ(T)\gamma(T) versus TT plot in panel (b) and it is this that we proceed to analyze. (There is a small anomaly at 18 K, present in the Zn-doped samples but very weak in the pure; and another at 120 K, present only in the pure sample. The sample variability indicates unidentified impurities and these anomalies are ignored in the following).

Our first task is to identify the normal-state coefficient, γn\gamma_{\textrm{n}}, that would occur in the absence of superconductivity. This is very much constrained by the displayed data for γ(T)\gamma(T) because γn\gamma_{\textrm{n}} must follow each of the three data sets above their respective TcT_{\textrm{c}} values. This is the black dashed curve. It is further tightly constrained by the requirement for entropy balance. Because the area under a γ(T)\gamma(T) curve is entropy then integrating γ(T)\gamma(T) from T=0T=0 to some T0>TcT_{0}>T_{\textrm{c}} must give the same result as integrating γn(T)\gamma_{\textrm{n}}(T) from T=0T=0 to T0T_{0}. The fit function which satisfies these two requirements is

γn(T)=1.93[10.913tanhα(E2kBT)ln[cosh(E/2kBT)]E/2kBT]\gamma_{\textrm{n}}(T)=1.93\left[1-0.913\tanh^{\alpha}\left(\frac{E^{*}}{2k_{B}T}\right)\frac{\ln[\cosh(E^{*}/2k_{B}T)]}{E^{*}/2k_{B}T}\right] (1)

where E=13.44E^{*}=13.44 meV (or T=E/kB=156T^{*}=E^{*}/k_{B}=156 K) and the exponent α=1.7\alpha=1.7. The general form of this equation for γ(T)\gamma(T) arises analytically from inserting into Eq. 3, below, a triangular gap in the density of states (DOS) with a finite DOS at the Fermi level Tstar . The amplitude 0.913 (being less than unity) reflects the finite DOS at EFE_{\textrm{F}}. This residual DOS is manifested in the finite value of γn(0)=0.183\gamma_{\textrm{n}}(0)=0.183 mJ/g.at.K2 and is a signature of the ungapped Fermi arcs, or hole pockets, of a reconstructed Fermi surface Storey2 ; Kunisada .

T=156T^{*}=156 K is typical of values reported for Y124 from transport Bucher and NMR relaxation Raffa measurements but we emphasize this reflects an energy scale not a temperature entrant . An important implication of Fig. 1 is that there is no coupling between superconductivity and the pseudogap in the sense that the onset of superconductivity does not weaken the pseudogap. This is evident from the fact that a single γn(T)\gamma_{\textrm{n}}(T) curve fits all three samples i.e. γn(T)\gamma_{\textrm{n}}(T) is the same for 4% and 0% Zn even in the temperature range below TcT_{\textrm{c}} for 0% Zn so that the onset of superconductivity in the latter case does not alter the underlying pseudogap energy scale, EE^{*}. Close scrutiny of the kk-dependent gap in Bi2212, as measured by angle-resolved photoelectron spectroscopy (ARPES) Vishik , (which allows separation of the antinodal pseudogap from the nodal superconducting gap on the Fermi arcs) confirms that the pseudogap amplitude does not alter on cooling below TcT_{\textrm{c}}.

NMR Knight shift and entropy

Next, we note that the spin susceptibility and electronic entropy are closely related. To see this consider the entropy for a weakly-interacting Fermi liquid Padamsee :

Sn=2kB[fln(f)+(1f)ln(1f)]N(E)dES_{\textrm{n}}=-2k_{B}\int_{-\infty}^{\infty}\![f\ln(f)+(1-f)\ln(1-f)]\,N(E)\,\mathrm{d}E (2)

where f(E)f(E) is the Fermi function and N(E)N(E) is the electronic DOS for one spin direction. This is just a weighted integral of the DOS with the ‘Fermi window’ [fln(f)+(1f)ln(1f)][f\ln(f)+(1-f)\ln(1-f)].

On the other hand, the spin susceptibility for a weakly-interacting Fermion system is:

χs=2μB2f(E)EN(E)dE,\chi_{s}=-2\mu_{B}^{2}\int_{-\infty}^{\infty}\!\frac{\partial f(E)}{\partial E}\,N(E)\,\mathrm{d}E, (3)

Therefore, like the entropy, the susceptibility is an integral of the DOS where the Fermi window is now the function f/E\partial f/\partial E. It turns out that Tf/ET\partial f/\partial E is essentially identical to [fln(f)+(1f)ln(1f)][f\ln(f)+(1-f)\ln(1-f)] if χs\chi_{s} in the former is stretched in temperature by a factor 1.187 Loram_IRC . It is therefore not surprising that S/TS/T and χs\chi_{\textrm{s}} are related. This relationship is expressed by the Wilson ratio, aWa_{\textrm{W}}, such that S(T)/T=aWχs(T)S(T)/T=a_{\textrm{W}}\chi_{\textrm{s}}(T), where

aW=13μ0(πkB2μB)2a_{W}=\frac{1}{3\mu_{0}}\left(\frac{\pi k_{B}^{2}}{\mu_{B}}\right)^{2} (4)
Refer to caption
Figure 2: Dashed curve: The normal-state entropy coefficient, Sn(T)/TS_{\textrm{n}}(T)/T, obtained by integrating the dashed curve for γn(T)\gamma_{\textrm{n}}(T) in Fig. 1(b). Red and black data points: the spin susceptibility, χs(T)\chi_{\textrm{s}}(T), calculated from the planar oxygen Knight shifts K2,c17{}^{17}K_{2,c} and K3,c17{}^{17}K_{3,c} associated with the O2 and O3 oxygen sites. χs(T)\chi_{\textrm{s}}(T) is expressed in entropy units by multiplying by the Wilson ratio, aWa_{W}, for weakly interacting Fermions. The green connected data points are the difference in O2 and O3 spin susceptibilities obtained from K2,c1717K3,c{}^{17}K_{2,c}-^{17}K_{3,c}, showing the abrupt onset of nematic splitting at 200 K within the pseudogap state. The inset shows the cuprate phase diagram with the three red data points of Sato et al. Sato marking the onset of nematicity. The blue star marks the onset of nematic splitting of the O2 and O3 Knight shifts.

We will now test this relationship in the present case of Y124. By integrating γn(T)\gamma_{\textrm{n}}(T) from T=0T=0 to TT we obtain the normal-state entropy and this is plotted as Sn(T)/TS_{\textrm{n}}(T)/T by the black dashed curve in Fig. 2. For fully-oxygenated Y123 Sn(T)/TS_{\textrm{n}}(T)/T is essentially independent of temperature Loram3 , reflecting the fact that the pseudogap has closed at maximal doping (p0.19p\approx 0.19 holes/Cu). But for Y124 there is a large pseudogap present which suppresses S/TS/T at low TT. This is also seen in the TT-dependent NMR Knight shift which is linearly related to the spin susceptibility. To illustrate, we show in Fig. 2 the 17O Knight shift, referenced to the chemical shift, as reported by Tomeno et al. Tomeno . These authors also report the bulk susceptibility as a function of the Knight shift, thus enabling calibration of the spin susceptibility from the Knight shift. As a final step we multiply the spin susceptibility by the Wilson ratio, aWa_{\textrm{W}}, in order to express the TT-dependent part of the Knight shift in entropy/TT units.

It can be seen in Fig. 2 that, not only the shape, but the absolute magnitude concurs remarkably well with the derived Sn(T)/TS_{\textrm{n}}(T)/T suggesting, as already noted for the bulk susceptibility Loram , that the near-nodal states are consistent with a weakly interacting Fermionic system. Of especial interest is the fact that the O2 and O3 Knight shifts begin to diverge below 200 K. The difference in shift, K2,c1717K3,c{}^{17}K_{2,c}-^{17}K_{3,c}, expressed in entropy units, is also shown in the figure (×\times10). This shows an abrupt onset in nematicity, consistent with that reported by the Matsuda group Sato using torque magnetometry. Its location at (pp=0.13, TT=200) is precisely consistent with the three data points in the (pp,TT) plane reported by the Matsuda group for Y123. (For the doping state of 0.13 see Materials). These are plotted in the inset to Fig. 2 by the red data points. Their susceptibility data was presented as evidence for a thermodynamic “phase transition at the onset of the pseudogap” however it is clear from Fig. 2 that the pseudogap is already open far above TnematicT_{\textrm{nematic}}, having already depleted half of the spin susceptibility. We observe no anomaly in γ(T)\gamma(T) at or near 200 K to suggest a phase transition. It must be very weak. We will see below that the pseudogap in fact extends at least to well above 400 K. Consequently this nematic phase transition occurs within a preexisting pseudogap state that extends far above and is not a transition into the pseudogap state, contrary to what has been claimed Sato .

Refer to caption
Figure 3: The measured low-temperature specific heat coefficient, γ(T,H)\gamma(T,H) for pure Y124 plotted as a function of field for TT = 6, 8, 10, 12 and 14 K. Above 1 tesla the data is essentially linear, consistent with the Volovik effect Volovik . The black dash/dot line shows the expected slope for Δ0=30\Delta_{0}=30 meV.

Volovik effect

We now consider the field-dependent low-TT behaviour of γ(T)\gamma(T) for the pure sample. Fig. 3 shows γ(T,H)\gamma(T,H) plotted as a function of μ0H\sqrt{\mu_{0}H} for TT = 6, 8, 10, 12 and 14 K. Above 1 tesla the behaviour is linear, consistent with the expected phenomenology of the Volovik effect. Such behaviour has been seen in the specific heats of single-crystal Y123 Junod and La2-xSrxCuO4 Wang , and in interlayer tunneling in Bi2212 Benseman . For dd-wave symmetry, in the superconducting state the finite ground-state specific heat coefficient is Wang

Δγ(H,0)=4kB23πϕ0nVM15dakF2Δ0μ0H,\Delta\gamma(H,0)=\frac{4k_{\textrm{B}}^{2}}{3\hbar}\sqrt{\frac{\pi}{\phi_{0}}}\frac{nV_{\textrm{M}}}{15d}\frac{a\hbar k_{\textrm{F}}}{2\Delta_{0}}\sqrt{\mu_{0}H}, (5)

in units of mJ/g.at.K2, where VMV_{\textrm{M}} is the molar volume, dd the unit cell length, ϕ0\phi_{0} the flux quantum, Δ0\Delta_{0} is the antinodal amplitude of the dd-wave gap, and kFk_{\textrm{F}} the Fermi wave vector at the node along the (π\pi,π\pi) direction. For Δ0=\Delta_{0}= 30 meV the dash/dot line in Fig. 3 shows the expected slope. The data is very consistent with this expectation. The averaged slope of 0.0443 mJ/g.at.K2T1/2 implies a gap amplitude of Δ0=\Delta_{0}= 34 meV. We have used the value kF=0.436×1010k_{\textrm{F}}=0.436\times 10^{10} m1{}^{-1}\, Sebastian .

Refer to caption
Figure 4: (a) The normal-state entropy, Sn(T)S_{\textrm{n}}(T), obtained by integrating the black dashed curve in Fig. 1(b) and the measured entropy, Ss(T)S_{\textrm{s}}(T), for pure Y124 in zero field. The condensation entropy, ΔSns=Sn(T)Ss(T)\Delta S_{\textrm{ns}}=S_{\textrm{n}}(T)-S_{\textrm{s}}(T) is shown underneath for μ0H\mu_{0}H = 0, 1, 3, 5, 7, 9, 11 and 13 T colour-coded as in (b). (b) The TT-dependence of the condensation free energy ΔFns=FnFs\Delta F_{\textrm{ns}}=F_{\textrm{n}}-F_{\textrm{s}} for pure Y124 obtained by integrating ΔSns(T)\Delta S_{\textrm{ns}}(T) from above TcT_{\textrm{c}} for each field as annotated. Below these are shown the condensation energy for the 2% Zn and 4% Zn samples showing a rapid suppression with scattering and applied field. The condensation energy for fully-oxygenated Y123 is also shown and is much higher.

Free energy and superconducting gap

Fig. 4(a) shows the normal-state and superconducting state entropy below 100 K obtained by integrating γn(T)\gamma_{\textrm{n}}(T) and γs(T)\gamma_{\textrm{s}}(T), respectively (as displayed in Fig. 1(b)) from 0 to TT. The curves shown are for pure Y124 in zero field and they are denoted Sn(T)S_{\textrm{n}}(T) and Ss(T)S_{\textrm{s}}(T), respectively. The difference is the condensation entropy ΔSns=Sn(T)Ss(T)\Delta S_{\textrm{ns}}=S_{\textrm{n}}(T)-S_{\textrm{s}}(T). This is plotted underneath for fields of μ0H\mu_{0}H = 0, 1, 3, 5, 7, 9, 11 and 13 T, colour-coded as in panel (b). Clearly the condensation entropy is rapidly suppressed in field. Another feature of note is the presence of fluctuations around TcT_{\textrm{c}} which broadens the transition somewhat. This will be discussed later.

By integrating ΔSns(T)\Delta S_{\textrm{ns}}(T) from a temperature T0T_{0}, sufficiently above the fluctuation regime that ΔSns(T0)=0\Delta S_{\textrm{ns}}(T_{0})=0, down to a temperature, TT, one obtains the condensation free energy ΔFns(T)\Delta F_{\textrm{ns}}(T). However, a more efficient way of calculating the condensation free energy using just a single integration is given by:

ΔFns(T)=T0TTΔγns(T)dTTT0TΔγns(T)dT,-\Delta F_{\textrm{ns}}(T)=\int_{T_{0}}^{T}\!T\,\Delta\gamma_{\textrm{ns}}(T)\,\mathrm{d}T-T\,\int_{T_{0}}^{T}\!\Delta\gamma_{\textrm{ns}}(T)\,\mathrm{d}T, (6)

where the first term is the condensation internal energy, ΔUns(T)-\Delta U_{\textrm{ns}}(T), and the second term is the condensation entropy term, TΔSns(T)T\Delta S_{\textrm{ns}}(T). The condensation free energy calculated in this way is plotted in Fig. 4(b) for 0%, 2% and 4% Zn and for the various annotated fields. ΔFns(T)\Delta F_{\textrm{ns}}(T) and its components ΔUns(T)\Delta U_{\textrm{ns}}(T) and TΔSns(T)-T\Delta S_{\textrm{ns}}(T) are plotted in Fig. 5. Also plotted in Fig. 4(b) is the condensation energy for fully oxygenated Y123 which rises to a ground-state value of 3400 mJ/g.at - a full six-fold greater than for pure Y124. This shows the full impact of the pseudogap for Y124 in weakening superconductivity. Also evident is the dramatic effect of impurity scattering in further reducing the condensation energy (which is particularly marked in underdoped cuprates where the pseudogap is present Tallon_scattering ). ΔFns(0)\Delta F_{\textrm{ns}}(0) for the 4% Zn-doped sample in zero field is just 28 mJ/g.at - 125 times smaller than for pure Y123. The curves for 2% Zn are dashed below 30 K and this is because ΔSns(T)\Delta S_{\textrm{ns}}(T) is a little noisy at low TT and in some cases does not fall exactly to zero, as it must. We find the first 20 K of the data scales precisely with ΔSns(T)\Delta S_{\textrm{ns}}(T) for the 0% Zn sample, and so we assumed that this scaling continues down to T=0T=0 thus enforcing ΔSns(T)\Delta S_{\textrm{ns}}(T) to fall to zero as T0T\rightarrow 0. Any errors introduced are very small - of the order of the thickness of the curves and of no consequence in the following analysis.

Refer to caption
Figure 5: (a) The TT-dependence of ΔFns(T)\Delta F_{\textrm{ns}}(T) and its components ΔUns(T)\Delta U_{\textrm{ns}}(T) and TΔSns(T)-T\Delta S_{\textrm{ns}}(T) calculated from Eq. 6 using γnγs\gamma_{\textrm{n}}-\gamma_{\textrm{s}} as given in Fig. 1. Most notable is the persistence of fluctuations high above TcT_{\textrm{c}} in both ΔUns(T)\Delta U_{\textrm{ns}}(T) and TΔSns(T)-T\Delta S_{\textrm{ns}}(T) and the almost complete suppression in ΔFns(T)\Delta F_{\textrm{ns}}(T). (b) The superconducting order parameter, Δ(T)\Delta^{\prime}(T), calculated from 2ΔUns(T)TΔSns(T)2\Delta U_{\textrm{ns}}(T)-T\Delta S_{\textrm{ns}}(T) using Eq. 7 for 0% and 2% Zn at fields given by the colour coding in panel (a).

As mentioned, Fig. 5(a) shows ΔFns(T)\Delta F_{\textrm{ns}}(T) and its components ΔUns(T)\Delta U_{\textrm{ns}}(T) and TΔSns(T)-T\Delta S_{\textrm{ns}}(T). It is striking that fluctuations persist high above TcT_{\textrm{c}} in both ΔUns(T)\Delta U_{\textrm{ns}}(T) and TΔSns(T)-T\Delta S_{\textrm{ns}}(T) while they are almost completely cancelled in ΔFns(T)\Delta F_{\textrm{ns}}(T). The superconducting gap function Δ(T)\Delta(T) may be calculated from these components of the free energy using Tallon2

ζN(0)Δ(T)2\displaystyle\zeta N(0)\Delta(T)^{2} =2ΔFns(T)+TΔSns(T)\displaystyle=2\Delta F_{\textrm{ns}}(T)+T\Delta S_{\textrm{ns}}(T)
2ΔUns(T)TΔSns(T)\displaystyle\equiv 2\Delta U_{\textrm{ns}}(T)-T\Delta S_{\textrm{ns}}(T) (7)

where ζ\zeta = 1 for ss-wave and 1/2 for dd-wave. The calculated Δ(T)\Delta(T) values are plotted in Fig. 5(b). The gap is the magnitude of the order parameter, Δ\Delta^{\prime}, rather than the spectral gap, Δ\Delta, which is higher Loram1 . Roughly speaking, Δ0=Δ02E2\Delta^{\prime}_{0}=\sqrt{\Delta_{0}^{2}-E^{*2}} Loram1 . From the impurity suppression of superfluid density Tallon7 we estimate Δ023.4\Delta_{0}\approx 23.4 meV while from the high-TT entropy suppression we determine E19.1E^{*}\approx 19.1 meV. The quadratic relation above then implies Δ0=13.5\Delta_{0}^{\prime}=13.5 meV, very consistent with the values in Fig. 5(b). The values of Δ(T)\Delta^{\prime}(T) are seen to descend towards zero at TcT_{\textrm{c}} then persist with a more slow decline above TcT_{\textrm{c}}. Here the pairing is incoherent Storey1 ; Kondo and is a feature of the strong superconducting fluctuations above TcT_{\textrm{c}}.

Comparison of Y124 with Y123

It is highly instructive to compare the measured data for Y124 with that for Y123. This is shown for γ(T)\gamma(T) in Fig. 6(a) and for S(T)S(T) in Fig. 6(b). The Y123 data is taken from ref. Loram_IRC which, in contrast to earlier reports Loram1 ; Cooper1 , includes a small correction for the background DOS in the undoped state. We also show the entropy-conserving normal-state functions as well. Two features are prominent. (i) the size of the specific heat jump for Y124 is much smaller than for Y123 due to the presence of the pseudogap in the former. (ii) while the γ(T)\gamma(T) curves converge above TcT_{\textrm{c}} the S(T)S(T) curves remain separated and parallel to the highest temperature. This is clear evidence that the pseudogap remains present in Y124 to the highest temperatures measured - here 400 K.

Refer to caption
Figure 6: Comparison of (a) γ(T)\gamma(T) and (b) S(T)S(T) for Y124 with nearly fully-oxygenated Y123. At highest oxygenation, x=0.97x=0.97, a double transition (see side-hump) is observed while a small depletion, x=0.92x=0.92, sees these merge into a single transition with a larger peak size. Above TcT_{\textrm{c}} the γ(T)\gamma(T) curves merge while for S(T)S(T) they remain separated. The inset in (a) shows the double-peaked Fermi window, /T\partial\mathcal{F}/\partial T, for calculating γ(T)\gamma(T) using Eq. 2, where (E,T)=[fln(f)+(1f)ln(1f)]\mathcal{F}(E,T)=[f\ln(f)+(1-f)\ln(1-f)]. The window is centred on a normal-state triangular gap in the DOS (red lines) located at EFE_{\textrm{F}}. The blue, purple and olive-green curves show /T\partial\mathcal{F}/\partial T for temperatures kBT=0.1Ek_{\textrm{B}}T=0.1E^{*}, 0.3E0.3E^{*} and 1.0E1.0E^{*}, respectively. The inset in (b) shows the single-peaked Fermi window, (E,T)\mathcal{F}(E,T), for calculating entropy, again centred on a triangular gap in the DOS.

To see this, consider the Fermi window for the entropy given in Eq. 2. This is the single-peaked function shown in the inset to Fig. 6(b). Three different temperatures of 0.1, 0.3 and 1.0 E/kBE^{*}/k_{\textrm{B}} are displayed, where EE^{*} is the magnitude of the triangular gap shown in the figure as a representation of the pseudogap. Provided that the pseudogap remains open, this Fermi window always sees the gap and at higher temperatures loses a fixed fraction of states so that S(T)S(T) is displaced down in parallel fashion relative to that for Y123 - as evidenced in Fig. 6(b), and again in Fig. 7(b). In contrast, because γS/T\gamma\equiv\partial S/\partial T, the Fermi window for γ(T)\gamma(T) is the TT-derivative of that for S(T)S(T). This is a double-peaked function as shown in the inset to Fig 6(a). It can be seen that at high enough temperature this Fermi window falls outside of the gap. Therefore γ(T)\gamma(T) recovers its full ungapped magnitude at high TT, despite the presence of the gap. The data in Fig. 6(a) and (b) are entirely consistent with this picture. The two systems have essentially the same background DOS. In the normal-state Y123 is ungapped while Y124 is gapped to the highest temperature as evidenced by the parallel suppression of S(T)S(T). If the gap were to close with increasing TT then S(T)S(T) for Y124 would recover to that for Y123. It does not. By extrapolating S(T)S(T) back to the ordinate axis one can read off the normal-state gap magnitude. For a triangular gap, as in the insets to Fig. 6, but with a finite DOS of N1N_{1} at EFE_{\textrm{F}} and a constant DOS of N0N_{0} above EE^{*} this negative intercept is 2ln2kB(N0N1)VME2\ln 2\,k_{\textrm{B}}(N_{0}-N_{1})V_{\textrm{M}}E^{*}, where VMV_{\textrm{M}} is the molar volume. The finite value of γn\gamma_{\textrm{n}} at T=0T=0 is given by γn0=4ln2×2.374kB2N1VM\gamma_{\textrm{n}}^{0}=4\ln 2\times 2.374k_{\textrm{B}}^{2}N_{1}V_{\textrm{M}}, while at high TT we have γn=(2/3)π2kB2N0VM\gamma_{\textrm{n}}=(2/3)\pi^{2}k_{\textrm{B}}^{2}N_{0}V_{\textrm{M}}. From these we obtain E=19.1E^{*}=19.1 meV and N0=6.14×103N_{0}=6.14\times 10^{-3} states/meV/cell.

Refer to caption
Figure 7: (a) 89Y NMR Knight shift for YBa2Cu3O6+x as reported by Alloul et al. Alloul with xx values annotated. Also shown is the MAS 89Y NMR Knight shift and the scaled TT variation of 1/T11/T_{1} for YBa2Cu4O8 as reported by Williams et al. Williams2 . (b) Y123: Entropy SS (blue solid curve Loram_IRC ), aW89χsTa_{\textrm{W}}\,^{89}\chi_{\textrm{s}}T (purple squares Alloul ) and bulk susceptibility aWχsTa_{\textrm{W}}\chi_{\textrm{s}}T (purple dashed curve Loram_IRC ) for Y123 at nearly full oxygenation. Y124: Entropy SS (red solid curve - this work), aW89χsTa_{W}\,^{89}\chi_{\textrm{s}}T (red spheres Williams ) and T/63T1T/^{63}T_{1} (open black triangles Raffa ; Williams2 ) for Y124. Then, as a proxy for Y124: aW89χsTa_{W}\,^{89}\chi_{\textrm{s}}T for Y123 with x=0.75x=0.75 (solid green triangles - Alloul Alloul ); and aWχsTa_{W}\chi_{\textrm{s}}T for Y123 with x=0.73x=0.73 (dashed green curve - Loram Loram_IRC ).

High-temperature susceptibility and entropy

We now extend this comparison to 400 K, a range never previously achieved in differential measurements. We combine this data with 89Y Knight shift data, bulk susceptibility measurements and 1/63T11/^{63}T_{1} data in such a way that demonstrates, individually and collectively, the persistence of the pseudogap to 400 K and beyond. 89K(T)s{}_{\textrm{s}}(T) probes the spin susceptibility of the CuO2 planes that sandwich the Y atom in Y123 and Y124 Alloul . The 89Y nucleus has the additional benefit of having no quadrupole moment so there is no quadrupole splitting of the resonance arising from electric field gradients. Further, the use of magic-angle spinning (MAS) enables extremely narrow line widths as will be used below Williams . The 1/63T11/^{63}T_{1} relaxation rate is a weighted sum over q of the imaginary part of the spin susceptibility, χ′′(q,ω)\chi^{\prime\prime}(\textrm{\bf q},\,\omega), where, for the 63Cu nucleus, the weighting form factor is strongly enhanced near the antiferromagnetic wave vector, q=(π,π)\textrm{\bf q}=(\pi,\pi) Crossover and hence 1/63T11/^{63}T_{1} is dominated by the antinodal pseudogap.

Fig. 7(a) shows the 89Y Knight shift 89K(T)s{}_{\textrm{s}}(T) for YBa2Cu3O6+x as reported by Alloul et al. Alloul , with values of xx annotated. Also plotted is MAS 89Ks{}_{\textrm{s}} for samples of Y124 from our laboratory Williams2 (red spheres), along with the 1/63T11/^{63}T_{1} data from Raffa et al. Raffa (up triangles) which we showed Williams2 scales precisely with 89Ks{}_{\textrm{s}}, as can be seen. (The conversion scale is 89K=s48.8×1/63T1164.5{}_{\textrm{s}}=48.8\times 1/^{63}T_{1}-164.5.) Notably, there is an excellent match over the entire temperature range with Alloul’s data for x=0.75x=0.75 (green up-triangles) where the doping state (0.13) and TcT_{\textrm{c}} for Y123 are much the same as those of Y124. As in Fig. 2, we convert 89Ks{}_{\textrm{s}} to spin susceptibility using the calibration of Alloul et al. Alloul (see Methods for more detail) and multiply by aWa_{\textrm{W}} to express in S/TS/T units. We find an excellent agreement between Alloul’s 89Ks{}_{\textrm{s}} and the measured entropy for Y123 across a wide range of doping and temperature (see Fig. S4 in Supplementary Information, SI).

In Fig. 7(b) we assemble four distinct data sets (SS, 89Ks{}_{\textrm{s}}, 1/63T11/^{63}T_{1}, and χs\chi_{\textrm{s}} from the bulk susceptibility) for Y123 at near full oxygenation (x=0.97x=0.97), and for Y124. All are expressed in entropy units, in this case using the factor aWTa_{W}T to convert susceptibilities (including the 1/63T11/^{63}T_{1} data expressed as a susceptibility). The χs\chi_{\textrm{s}} data is shown for Y123 with x=0.97x=0.97 and x=0.73x=0.73 (purple and green dashed curves, respectively, the latter as a proxy for Y124) and is taken from Loram et al. Loram_IRC . Evidently SS (blue solid curve), aW89χsTa_{W}\,^{89}\chi_{\textrm{s}}T (purple squares) and aWχsTa_{W}\chi_{\textrm{s}}T (purple dashed curve) for Y123 at full oxygenation all track linearly to the origin indicating the absence of the pseudogap. For Y124 on the other hand, SS (red solid curve - this work), aW89χsTa_{W}\,^{89}\chi_{\textrm{s}}T (red spheres) and T/63T1T/^{63}T_{1} (open black triangles), together with Alloul’s aW89χsTa_{W}\,^{89}\chi_{\textrm{s}}T (solid green triangles) for Y123 with x=0.75x=0.75 and aWχsTa_{W}\chi_{\textrm{s}}T (dashed green curve) for Y123 with x=0.73x=0.73, both as proxies for Y124, all reveal a linear high-temperature behaviour that extrapolates to a negative intercept on the yy-axis indicating the presence of a gap - the pseudogap. For Y124 the entropy curve almost completely overlays the bulk susceptibility green-dashed curve which is barely visible, so the figure is reproduced in the SI with the green-dashed curve overlaying the entropy. Evidently, the agreement over the full temperature range is excellent. Importantly, our entropy data extends to 400 K as does the χs\chi_{\textrm{s}} data, and the 89Ks{}_{\textrm{s}} data extends to 370 K. Fig. 7(b) represents the central result of this work. There is no indication, at any temperature, of the entropy recovering to the gap-less curve observed for fully-oxygenated Y123 that would signify the closing of the pseudogap at, or around, some TT^{*} value. The small upturn in S(T)S(T) near 400 K simply represents the limitations of the present differential technique at such a high temperature and is not seen in the χs\chi_{\textrm{s}} data. We conclude that the pseudogap does not close at some postulated TT^{*} in the range 150 to 200 K but remains open to the highest temperature investigated - 400 K. A similar conclusion has recently been drawn from 1/17T11/^{17}T_{1} planar oxygen NMR relaxation data for a number of cuprates Haase . We showed the same long ago Tstar for the in-plane resistivity and similarly for the cc-axis resistivity entrant ; Bernhard1 .

Refer to caption
Figure 8: Black curve: The as-measured field-induced differential specific heat coefficient, Δγ(13,T)=γ(13,T)γ(0,T)\Delta\gamma(13,T)=\gamma(13,T)-\gamma(0,T). Red curve: resonance γ(T)\gamma(T) as calculated using Eq. 8. The blue curve is the difference, which, below 2.8 K, is extrapolated to a finite value as T0T\rightarrow 0 to ensure overall entropy conservation.

Scattering resonance

Finally, we wish to discuss the upturn in the raw γ(T)\gamma(T) data at low TT seen in Fig. 1(a). This is in fact a peak rather than an upturn and may be identified with an impurity resonance Pan . Scanning tunneling spectroscopy (STS) measurements in lightly Zn-doped Bi2Sr2CaCu2O8+δ reveal resonance spots in spatial maps at low energy and low temperature Pan . Away from these spots, tunneling spectra reveal a well-formed dd-wave superconducting gap with sharp coherence peaks. Tunneling spectra collected on the spots (the location of individual Zn atoms) show a nearly full suppression of both the gap and the coherence peaks with, instead, a sharp resonance appearing at εr=1.5\varepsilon_{\textrm{r}}=-1.5 meV. To calculate the entropy contribution arising from this resonance we replace the DOS in Eq. 2 by a delta function, N(E)=Nrδ(εεr)N(E)=N_{\textrm{r}}\delta(\varepsilon-\varepsilon_{\textrm{r}}). The equation integrates to give:

Δγres=Ares(εr2kBT)3cosh2(εr2kBT)\Delta\gamma_{\textrm{res}}=A_{\textrm{res}}\left(\frac{\varepsilon_{\textrm{r}}}{2k_{\textrm{B}}T}\right)^{3}\cosh^{-2}\left(\frac{\varepsilon_{\textrm{r}}}{2k_{\textrm{B}}T}\right) (8)

where the amplitude Ares=(4kB2/εr)NrA_{\textrm{res}}=\left(4k_{\textrm{B}}^{2}/\varepsilon_{\textrm{r}}\right)N_{\textrm{r}}.

In view of the relationship between entropy and spin susceptibility discussed above, it is highly instructive to contrast this resonance component of γ(T)\gamma(T) with that of the susceptibility. Again, replacing the DOS in Eq. 3 by N(E)=Nrδ(εεr)N(E)=N_{\textrm{r}}\delta(\varepsilon-\varepsilon_{\textrm{r}}) we find:

χs,res=Bres(εr2kBT)cosh2(εr2kBT)\chi_{\textrm{s,res}}=B_{\textrm{res}}\left(\frac{\varepsilon_{\textrm{r}}}{2k_{\textrm{B}}T}\right)\cosh^{-2}\left(\frac{\varepsilon_{\textrm{r}}}{2k_{\textrm{B}}T}\right) (9)

where the amplitude Bres=(μB2/2εr)NrB_{\textrm{res}}=\left(\mu_{\textrm{B}}^{2}/2\varepsilon_{\textrm{r}}\right)N_{\textrm{r}}. The interesting point in relation to Eqs. 8 and  9 is that Δγres\Delta\gamma_{\textrm{res}} falls off rapidly as T3T^{-3} while χs,res\chi_{\textrm{s,res}} falls off more slowly as T1T^{-1}. This is borne out by our experimental data. Experimental evidence from the magnetic susceptibility for a Zn-induced resonance within the pseudogap was discussed previously in relation to Y124 Williams1 and La2-xSrxCuO4 Islam .

Fig. 8 shows the as-measured field-induced change in specific heat coefficient, Δγ(13,T)=γ(13,T)γ(0,T)\Delta\gamma(13,T)=\gamma(13,T)-\gamma(0,T). Recall that this difference contains no correction for the residual phonon contribution so is free of any imposed model. As well as the suppression of fluctuations around the specific heat anomaly near TcT_{\textrm{c}} the low-temperature resonance is evident. We fit this using Eq. 8 with the parameters Ares=0.61A_{\textrm{res}}=0.61 mJ/g.at.K2 and εr=1.55\varepsilon_{\textrm{r}}=-1.55 meV, the latter value being nicely consistent with the STS result for Bi2Sr2CaCu2O8+δ Pan . This fit is the red curve in Fig. 8 and the difference is shown by the blue curve. The calculated resonance response is an excellent fit and shows the rapid decay at higher temperatures associated with the T3T^{-3} tail. The difference is close to entropy conserving and requires a straightforward extrapolation below 2.8 K of its trend above 2.8 K to achieve exact entropy conservation. This rapid decay of the resonance in γ(T)\gamma(T) contrasts the predicted much slower T1T^{-1} decay in the resonance part of the spin susceptibility. We have previously investigated the 89Y NMR Knight shift in Zn-doped Y124 Williams1 . The Zn resonance contribution to the spin susceptibility is seen in a satellite peak which has a slowly-decaying Curie temperature dependence observable all the way up to 300 K, thus nicely confirming the behaviour predicted by Eq. 9.

Conclusions

In conclusion, we have measured the electronic specific heat of YBa2Cu4-xZnxO8 using a precision differential technique that allows separation of the electronic term from the lattice term up to an unprecedented 400 K. The pure sample reveals the expected Volovik effect which is fully suppressed in the Zn-doped samples. We show that the pseudogap, characteristic of underdoped cuprates, always remains open to above 400 K, far above the nominal pseudogap onset temperature usually proposed, TT^{*}\approx 150 - 200 K. Weak thermodynamic transitions reported at TT^{*}, for example the onset of susceptibility nematicity, occur within the already fully established pseudogap and are not transitions into the pseudogap state as widely claimed. The spin susceptibility, derived from the 17O and 89Y Knight shift and expressed in entropy units, is numerically the same as the electronic entropy divided by temperature, indicating that the near-nodal states are those of weakly-interacting Fermions. We derive the field-dependent condensation energy and superconducting energy gap. These expose the presence of strong pairing fluctuations extending well above TcT_{\textrm{c}} as well as canonical impurity scattering behaviour, including the expected low-temperature resonance response which in the entropy channel decays rapidly as T3T^{-3}, but in the spin channel decays slowly as T1T^{-1}. The measurements and analysis reveal the remarkable utility of the differential specific heat technique in exposing the rich physics of strongly-correlated electronic materials.

.1 Methods

Materials. The samples were prepared from stoichiometric proportions of high purity Y2O3, dried Ba(NO3)2, ZnO and CuO, pressed as pellets and reacted for 16 hours at 935C under an oxygen pressure of 60 bar. The samples were ground finely and the process repeated three more times. X-ray diffraction revealed single-phase YBa2Cu4O8 (see Fig. S1) with impurity less than 2% (only CuO identified). Lattice parameters were found to be a=0.3842a=0.3842 nm, b=0.3870b=0.3870 nm, and c=2.7235c=2.7235 nm under Ammm symmetry with a preferred alignment of the cc-axis normal to the plane of the pellets. The TcT_{\textrm{c}} values, determined from sharp diamagnetic onset were 81.2 K (0% Zn), 51.8 K (2% Zn) and 29.2 K (4% Zn). The quoted Zn concentrations are those referred to the CuO2 plane as essentially no Zn resides on the chains. Thus 2% Zn refers to the composition YBa2Cu3.96Zn0.04O8 and 4% refers to YBa2Cu3.92Zn0.08O8. Thermoelectric power measurements at 290 K give 6.95 μ\muV/K (0% Zn), 6.89 μ\muV/K (2% Zn) and 6.93 μ\muV/K (4% Zn). From the correlation of thermoelectric power with doping Obertelli this amounts to essentially identical doping states of 0.130 holes/Cu for each.

The suppression of TcT_{\textrm{c}} with Zn substitution is very much in line with that for underdoped Y123 at the same doping state. Fig. S2 shows TcT_{\textrm{c}} as a function of doping for 0, 2, 4 and 6% planar Zn substitution for Y0.8Ca0.2Ba2Cu3O7-δ while the red stars show the data for Zn-substituted Y124. They are very consistent. The rapid suppression of TcT_{\textrm{c}} for underdoped samples (open symbols) compared with overdoped (filled symbols) is a signature of the pseudogap which lowers the DOS at the Fermi level and hence raises the scattering rate Tallon_scattering . Fig. S3 shows TcT_{\textrm{c}} as a function of planar Zn concentration for Y0.8Ca0.2Ba2Cu3O7-δ at various doping states while red stars show the same data for Y124. Again they are very consistent. Typically, for higher Zn concentrations, the TcT_{\textrm{c}} value sits higher than that expected from Abrikosov-Gorkov pairbreaking due to the statistical overlap of nearest neighbours to the Zn substituent Tallon_scattering .

Spin susceptibility and Knight shift. The spin susceptibility, χs\chi_{\textrm{s}}, is related to the bulk magnetic susceptibility, χm\chi_{\textrm{m}}, by χm=χs+χ0\chi_{\textrm{m}}=\chi_{\textrm{s}}+\chi_{0} where the constant χ0\chi_{0} comprises diamagnetic and van Vleck terms and is evaluated for Y123 by Alloul et al. Alloul . The measured Knight shift, 89Ks{}_{\textrm{s}}, is given by 89Ks{}_{\textrm{s}} = σ0(x)+a(x)89χs\sigma_{0}(x)+a(x)\,^{89}\chi_{\textrm{s}}, where σ0(x)\sigma_{0}(x) is the TT-independent chemical shift and a(x)a(x) is the relevant hyperfine coupling constant and, as indicated, both change only with oxygen content, xx. For each xx therefore, χs\chi_{\textrm{s}}, χm\chi_{\textrm{m}} and 89Ks{}_{\textrm{s}} are linearly related to within an additive constant. We used values of a(x)a(x) reported by Alloul Alloul , while for each xx the additive constant was determined by matching the χs89{}^{89}\chi_{\textrm{s}} data to our bulk susceptibility data, χs\chi_{\textrm{s}} Loram_IRC . This fixed the value of the constant σ0(x)\sigma_{0}(x) which differed somewhat from those of Alloul but other literature values also reflect those differences Williams2 . The overall TT-dependence (independent of σ0(x)\sigma_{0}(x)) was an excellent match.

Acknowledgements

We are grateful to Dr. J. R. Cooper for helpful comments on the Knight shift and spin susceptibility. We also thank Dr Martin Ryan for assistance with the x-ray diffraction analysis.

Author contributions

JLT synthesized and characterized the samples, JWL carried out the specific heat measurements and the initial analysis to extract the electronic specific heat. JLT analyzed the data and wrote the paper.

deceased November 2017.

References

  • (1) Norman, M. R., Pines, D. & Kallin, C. The pseudogap: friend or foe of high-TcT_{\textrm{c}}. Adv. Phys. 54, 715–733 (2005).
  • (2) Timusk, T. & Statt, B. The pseudogap in high-temperature superconductors: an experimental survey. Rep. Progr. Phys. 62, 61-122 (1999).
  • (3) Tallon, J. L. & Loram, J. W. The doping dependence of TT^{*} – what is the real high-TcT_{\textrm{c}} phase diagram? Physica C 349, 53 (2001).
  • (4) Tallon, J. L., Storey, J. G., Cooper, J. R. & Loram, J. W. Locating the pseudogap closing point in cuprate superconductors: absence of entrant or reentrant behavior. Phys. Rev. B 101, 174512 (2020).
  • (5) Fauqué, B., Sidis, Y., Hinkov, V., Pailhès, S. Lin, C. T., Chaud, X. & Bourges, P. Magnetic order in the pseudogap phase of high-TcT_{\textrm{c}} superconductors. Phys. Rev. Lett. 96, 197001 (2006).
  • (6) Xia, J. et al. Polar Kerr-effect measurements of the high-temperature YBa2Cu3O6+x superconductor: evidence for broken symmetry near the pseudogap temperature. Phys. Rev. Lett. 100, 127002 (2008).
  • (7) Hashimoto, M. et al., Particle–hole symmetry breaking in the pseudogap state of Bi2201. Nature Phys. 6, 414-418 (2010).
  • (8) He, R.-H. et al. From a single-band metal to a high-temperature superconductor via two thermal phase transitions. Science 331, 15791583 (2011).
  • (9) Shekhter, A. et al. Bounding the pseudogap with a line of phase transitions in YBa2Cu3O6+δ. Nature 498, 75-77 (2013).
  • (10) Sato, Y. et al. Thermodynamic evidence for a nematic phase transition at the onset of the pseudogap in YBa2Cu3Oy. Nature Phys. 13, 1074-1078 (2017).
  • (11) Loram, J. W., Mirza, K. A., Cooper, J. R. & Liang, W. Y. Electronic specific heat of YBa2Cu3O6+x from 1.8 to 300 K. Phys. Rev. Lett. 71, 1740-1743 (1993).
  • (12) Loram, J. W., Mirza, K. A., Cooper, J. R., Liang, W. Y. & Wade, J. M. Electronic specific heat of YBa2Cu3O6+x from 1.8 to 300 K. J. Supercon. 7, 243-249 (1994).
  • (13) Loram, J. W., Luo, J. L., Cooper, J. R., Liang, W. Y. & Tallon, J. L. The condensation energy and pseudogap energy scale of Bi:2212 from the electronic specific heat. Physica C 341-348, 831 (2000).
  • (14) Williams, G. V. M., Talion, J. L., Meinhold, R. & Jánossy, A. 89Y NMR study of the effect of Zn substitution on the spin dynamics of YBa2Cu408. Phys. Rev. B 51 16503-16506 (1995).
  • (15) Tallon, J. L., Bernhard, C., Williams, G. V. M. & Loram, J. W. Zn-induced TcT_{c} Reduction in High-TcT_{c} Superconductors: Scattering in the Presence of a Pseudogap. Phys. Rev. Lett. 79, 5294-5297 (1997).
  • (16) Volovik, G. E. JETP Lett. 58, 469-473 (1993).
  • (17) Chen, S.-D., Hashimoto, M., He, Y., Song, D., Xu, K.-J., He, J.-F., Devereaux, T. P., Eisaki, H., Lu, D.-H., Zaanen, J. & Shen, Z.-X. Incoherent strange metal sharply bounded by a critical doping in Bi2212, Science 366, 1099-1102 (2019).
  • (18) Storey, J.G. Hall effect and Fermi surface reconstruction via electron pockets in the high-TcT_{\textrm{c}} cuprates, Europhys. Lett. 113, 27003 (2016).
  • (19) Kunisada, S. et al. Observation of small Fermi pockets protected by clean CuO2 sheets of a high-TcT_{\textrm{c}} superconductor. Science 369, 833-838 (2020).
  • (20) Bucher, B., Steiner, P., Karpinski, J., Kaldis, E. & Wachter, P. Influence of the Spin Gap on the Normal State Transport in YBa2Cu408. Phys. Rev. Lett. 70, 2012-2015 (1993).
  • (21) Raffa, F., Ohno, T., Mali, M., Roos, J., Brinkmann, D., Conder, K. & Eremin, M. Isotope Dependence of the Spin Gap in YBa2Cu408 as Determined by Cu NQR Relaxation. Phys. Rev. Lett. 81, 5912-5915 (1998).
  • (22) Vishik, I. M. et al. Phase competition in trisected superconducting dome. Proc. Nat. Acad, Sci. 109, 18332-18337 (2012).
  • (23) Padamsee, H., Neighbor, J. E. & Shiffman, C. A. Quasiparticle Phenomenology for Thermodynamics of Strong-Coupling Superconductors. J. Low Temp. Phys. 12, 387-411 (1973).
  • (24) Loram, J. W., Mirza, K. A. & Cooper, J. R. Properties of the superconducting condensate and normal-state pseudogap in high-TcT_{\textrm{c}} superconductors derived from the electronic specific heat. IRC in Superconductivity Research Review Cambridge University (1998).
  • (25) Loram, J. W., Mirza, K. A., Wade, J. M., Cooper, J. R. & Liang, W. Y. The Electronic Specific Heat of Cuprate Superconductors. Physica C 235-240, 134-137 (1994).
  • (26) Tomeno, I., Machi, T., Tai, K., Koshizuka, N., Karnbe, S., Hayashi, A., Ueda, Y. & Yasuoka, H. NMR study of spin dynamics at planar oxygen and copper sites in YBa2Cu408. Phys. Rev. B 49, 15327-15334 (1994).
  • (27) Loram, J. W., Luo, J. L., Cooper, J. R., Liang, W. Y. & Tallon, J. L. Evidence on the pseudogap and condensate from the electronic specific heat. J. Phys. Chem. Solids 62, 59-64 (2001).
  • (28) Wang, Y., Revaz, B., Erb, A. & Junod, A. Direct observation and anisotropy of the contribution of gap nodes in the low-temperature specific heat of YBa2Cu3O7. Phys. Rev. B 63, 094508 (2001).
  • (29) Wang, Y., Yan, J., Shan, L., Wen, H.-H., Tanabe, Y., Adachi, T. & Koike, Y. Weak-coupling dd-wave BCS superconductivity and unpaired electrons in overdoped La2-xSrxCuO4 single crystals. Phys. Rev. B 76, 064512 (2007)
  • (30) Benseman, T. M., Cooper, J. R. & Balakrishnan, G. Temperature and field dependence of the intrinsic tunnelling structure in overdoped Bi2Sr2CaCu2O8+δ, Phys. Rev. B 98, 014507 (2018).
  • (31) Sebastian, S. E., Harrison, N., Liang, R., Bonn, D. A., Hardy, W. N., Mielke, C. H. & Lonzarich, G. G. Quantum Oscillations from Nodal Bilayer Magnetic Breakdown in the Underdoped High Temperature Superconductor YBa2Cu3O6+x. Phys. Rev. Lett. 108, 196403 (2012).
  • (32) Tallon, J. L., Barber, F., Storey, J. G. & Loram, J. W. Coexistence of the superconducting energy gap and pseudogap above and below the transition temperature of cuprate superconductors. Phys. Rev. B 87, 140508(R) (2013).
  • (33) Storey, J. G. Incoherent superconductivity well above TcT_{\textrm{c}} in high-TcT_{\textrm{c}} cuprates – harmonizing the spectroscopic and thermodynamic data. New J. Phys. 19, 073026 (2017).
  • (34) Kondo, T., Malaeb, W., Ishida, Y., Sasagawa, T., Sakamoto, H., Takeuchi, T., Tohyama, T. & Shin, S. Point nodes persisting far beyond TcT_{\textrm{c}} in Bi2212. Nature Commun. 6, 7699 (2015).
  • (35) Cooper, J. R. & Loram, J. W. Some Correlations Between the Thermodynamic and Transport Properties of High TcT_{\textrm{c}} Oxides in the Normal State. J. Phys. I France 6, 2237-2263 (1996).
  • (36) Alloul, H., Ohno, T. & Mendels, P. 89Y NMR evidence for a fermi-liquid behavior in YBa2Cu3O6+x. Phys. Rev. Lett. 63, 1700-1703 (1989).
  • (37) Williams, G. V. M., Tallon, J. L., Quilty, J. W., Trodahl, H. J. & Flower, N. E. Absence of an Isotope Effect in the Pseudogap in YBa2Cu4O8 as Determined by High-Resolution 89Y NMR. Phys. Rev. Lett. 80, 377-380 (1998).
  • (38) Williams, G. V. M., Tallon, J. L. & Loram, J. W. Crossover temperatures in the normal-state phase diagram of high-TcT_{\textrm{c}} superconductors. Phys. Rev. B 58, 15053-15061 (1998).
  • (39) Williams, G. V. M., Pringle, D. J. & Tallon, J. L. Contrasting oxygen and copper isotope effects in YBa2Cu4O8 superconducting and normal states. Phys. Rev. B 61, 9257-9260(R) (2000).
  • (40) Tallon, J. L., Bernhard, C., Binninger, U., Hofer, A., Williams, G. V. M., Ansaldo, E. J., Budnick, J. I. & Niedermayer, Ch. In-plane anisotropy of the penetration depth due to superconductivity on the CuO chains in YBa2Cu3O7-δ, Y2Ba4Cu7O15-δ and YBa2Cu4O8. Phys. Rev. Lett. 74, 1008-1011 (1995).
  • (41) Nachtigal, J., Avramovska, M., Erb, A., Pavicevic, D., Guehne, R. & Haase, J. Temperature independent cuprate pseudogap from planar oxygen NMR. Condens. Matter 5, 66 (2020).
  • (42) Yu, Li, Munzar, D., Boris, A. V., Yordanov, P., Chaloupka, J., Wolf, Th., Lin, C. T., Keimer, B. & Bernhard, C. Evidence for two separate energy gaps in underdoped high-temperature cuprate superconductors from broadband infrared ellipsometry. Phys. Rev. Lett. 100, 177004 (2008).
  • (43) Pan, S. H., Hudson, E. W., Lang, K. M., Eisaki, H., Uchida, S. & Davis, J. C. Imaging the effects of individual zinc impurity atoms on superconductivity in Bi2Sr2CaCu2O8+δ. Nature 403, 746-750 (2003).
  • (44) Islam, R. S., Cooper, J. R., Loram, J. W. & Naqib, S. H. Pseudogap and doping-dependent magnetic properties of La2-xSrxCu1-yZnyO4. Phys. Rev. B 81 054511-8 (2010).
  • (45) Obertelli, S. D., Cooper, J. R. & Tallon, J. L. Systematics in the Thermoelectric power of High-TcT_{\textrm{c}} Oxides. Phys. Rev. B 46, 14928-14931 (1992).