Unconventional Coherence Peak in Cuprate Superconductors
Abstract
The Hebel-Slichter coherence peak, observed in the spin-lattice relaxation rate just below the critical temperature , serves as a crucial experimental validation of the Bardeen-Cooper-Schrieffer pairing symmetry in conventional superconductors. However, no coherence peak in has been observed in unconventional superconductors like cuprates. In this study, an unconventional coherence peak is identified for the first time using nuclear quadrupole resonance on YBa2Cu4O8, pointing to a distinctive pairing symmetry. The spin-lattice relaxation rate in nuclear quadrupole resonance and nuclear magnetic resonance with nuclear spin comprises the magnetic relaxation rate , which probes magnetic fluctuations, and the quadrupole relaxation rate , which probes charge fluctuations. By utilizing 63Cu and 65Cu isotopes, we successfully distinguish and of YBa2Cu4O8 and reveal the presence of the coherence peak in but not in , in contrast to conventional superconductors. Our finding demonstrates that unconventional superconductors do not exhibit a coherence peak in when the relaxation is due to fluctuations of the hyperfine field. Conversely, a coherence peak is expected when the relaxation is caused by electric field gradient fluctuations, due to the different coherence factors between charge and magnetic fluctuations. Our successful measurements of for the chains of YBa2Cu4O8 suggest that, should the conditions for predominant quadrupole relaxation be satisfied, this phenomenon could provide a novel approach to exploring the unconventional nature of the pairing mechanism in other superconductors.
I Introduction
Unconventional superconductors, such as -wave and -wave superconductors, are characterized by a varying sign of the superconducting gap function in momentum space, with their Cooper pairs widely believed to originate from electron-electron correlation rather than electron-phonon coupling[1]. Therefore, unconventional superconductors exhibit distinct physical properties, such as the absence of the Hebel-Slichter coherence peak [2, 3, 4, 5, 6, 7, 8, 9, 10]. In conventional superconductors with uniform superconducting gaps, the density of states (DOS) diverges at the gap energy, leading to a huge enhancement of spin-lattice relaxation rate just below the critical temperature in nuclear magnetic resonance (NMR) experiment, called the Hebel-Slichter coherence peak[11]. Conversely, this Hebel-Slichter coherence peak is absent in unconventional superconductors[10], such as d-wave cuprate superconductors, where the DOS divergence persists. Consequently, detecting coherence peaks poses a significant challenge for unconventional superconductors. In this work, we reveal a novel unconventional coherence peak in high-temperature cuprate superconductors using nuclear quadrupole resonance (NQR), introducing a distinctive feature of unconventional superconductivity.
The nucleus with the spin is nonspherical and possesses an electric quadrupole moment , where and are spatial directions x, y, z. This results in the nucleus having an electrostatic energy that varies depending on its orientation within an electric field gradient (EFG) generated by the local potential from its surrounding environments[12]. Hence, the electrostatic energy splitting between different nucleus spin is utilized in NQR, analogous to how NMR employs the magnetic energy splitting of each spin under a magnetic field. Furthermore, the electric quadrupole moment is a valuable tool for investigating electric field dynamics, akin to how the magnetic moment is used to study spin dynamics. These tools complement each other in exploring electromagnetic fluctuations in condensed matter[13]. A key dynamic signal is the spin-lattice relaxation rate, which describes how the nuclei arrive at their thermal equilibrium via the process of spin-lattice relaxation and is proportional to the summation of the imaginary part of the dynamical susceptibility. In the conventional -wave superconducting state, the rate can be expressed as[14]
(1) |
where and are the relaxation times in the normal state and the superconducting state, respectively. is the DOS in the normal state and for is the DOS in the superconducting state. is the Fermi distribution function and is the superconducting gap. The sign of the coherence factor is contingent on the nature of the perturbation causing the transition. Magnetic relaxations, being not time-reversal invariant, have which enhance the divergency of and the magnetic relaxation rate exhibits a Hebel-Slichter coherence peak in conventional superconductors[14]. In contrast, quadrupole relaxations, being time-reversal invariant, have which compensate for the divergency of and the quadrupole relaxation rate drops rapidly below in conventional superconductors[15, 16]. In unconventional superconductors, drops or decreases gradually below , without a coherence peak being observed[2]. These phenomena contradict the logarithmic divergence of in cuprate superconductors[17]. Additionally, the cuprate superconductor gap sign varies in momentum space, significantly impacting the coherence factor and potentially leading to a shift of the coherence peak to .
II Result
Whether a coherence peak is present in motivated us to perform NQR on two isotopes 63Cu and 65Cu in oxygen-stoichiometric and underdoped YBa2Cu4O8. This compound is an unconventional superconductor with a transition temperature K[18], as confirmed by the temperature-dependent magnetization measurement in Fig. 6. As previously mentioned, NQR is particularly sensitive to the local chemical environment and can distinguish between Cu atoms situated in distinct positions. YBa2Cu4O8 contains two Cu sites, namely, the chain Cu(1) and the planar Cu(2) (depicted in the upper inset in Fig. 1), offering a unique avenue to probe and and explore the coherence peak. The Cu(1) and Cu(2) sites have different EFG strengths, resulting in different resonance frequencies as shown in the spectra inserted in Fig. 1. Additionally, the distinct EFG tensors of Cu(1) and Cu(2) influence their electric dynamic behavior, which is discussed later. Two isotopes 63Cu and 65Cu lead to two NQR resonance peaks at each site, for a total of four peaks at both sites, as shown in the lower inset in Fig. 1 for the spectrum at .
The four NQR spin-lattice relaxation rates corresponding to the four peaks are measured and plotted in Fig. 1. For the planar Cu(2), the of 63Cu(2) and 65Cu(2) (green and purple lines, respectively) drop due to the reduction in DOS caused by the superconducting gap below , with noticeable kink behaviors observed at . The of Cu(2) above keeps increasing and saturates around 200 K[19]. These data are consistent with previous reports in the whole temperature range[4, 6] and similar to YBa2Cu3O7[5]. On the other hand, the of 63Cu(1) and 65Cu(1) (cyan and red lines, respectively) decrease with temperature decreasing above , resembling the behavior of a conventional metal. Notably, the of Cu(1) exhibits a slight increase just below , followed by a decrease at lower temperatures. These distinctive characteristics are highly unusual, suggesting a coherence peak akin to that observed in conventional superconductors.
Let us focus our attention on the superconducting transition and eliminate the linear temperature term in the metal by drawing /T around in Figs. 2(a) and 2(b). Both /T of 63Cu and 65Cu at the Cu(1) site exhibit a coherence peak just below , whereas /T of the Cu(2) site commences a quite steep descent below . The distinct temperature-dependence behaviors of Cu(1) and Cu(2) are attributed to their different EFG. can be expressed as a sum of magnetic relaxation rate and quadrupole relaxation rate , for I = 3/2 NQR experiment[20]. Theoretically, the nuclear spin Hamiltonian of the interaction between quadrupole moment Q and EFG can be written as[12]
(2) |
where is the quadrupole resonance frequency along the principal axis, . is an asymmetry parameter of the EFG, where , and are the EFGs along the x, y, and z directions, respectively. Only the terms with and can flip spins and contribute to , necessitating a sufficiently large to measure . The planar Cu(2) is isotropic with a negligible [4], resulting in of Cu(2) being purely affected by magnetic relaxation and unable to detect EFG fluctuations[2]. Therefore, although there are charge fluctuations in the Cu-O plane detected by planar 17O with [21], they cannot be detected by Cu(2)[22]. On the other hand, the chain Cu(1) is anisotropic with = 0.85, causing of Cu(1) to be affected by both magnetic and quadrupole relaxation. If the coherence peak is due to magnetic relaxation, it should be observed at both Cu(1) and Cu(2) sites. The presence of the coherence peak exclusively at the Cu(1) site suggests its origin from quadrupole relaxation.
and can be decomposed by utilizing two isotopes 63,65Cu, which possess the same spin I=3/2 but different gyromagnetic ratios and quadrupole moments[2]. Using isotopes can avoid the influence from form factors and principle axes directions[23], since they occupy the same atom site. The decomposition equations are and , where the quotients and are known[24]. Before decomposing, the ratio of is calculated to assess the weight of and , shown in Figs. 2(c) and 2(d). at the Cu(2) site remains a constant value around the magnetic quotient for various temperatures, indicating negligible and the presence of only . Conversely, at the Cu(1) site deviates from around , suggesting a detectable component.
The magnetic and quadrupole relaxations of Cu(1) are decomposed using the relations mentioned above, and the results are illustrated in Figs. 3(a) and 3(b), respectively. The depicted in Fig. 3(a) drops below due to the loss of DOS, aligning with the behavior of the at the Cu(2) site where solely magnetic relaxation occurs and . In contrast, the displays a prominent coherence peak just below , in sharp contrast to conventional superconductors where the coherence peak is evident in the and absent in . This discrepancy elucidates the absence of a coherence peak at the Cu(2) site, as it is unaffected by quadrupole relaxation. It is noteworthy that the is smaller than , underscoring the higher sensitivity of most nuclei, such as Cu, to magnetic fluctuations in dipolar interaction than to EFG fluctuations in quadrupole interaction. Moreover, in the superconducting state, on lowering the temperature, the quadrupolar relaxation diminishes faster than the magnetic one[22]. Consequently, the coherence peaks in Fig. 2 (a) have historically been inconspicuous and challenging to detect.
There is one thing we need to emphasize: The superfluid density of the system around is dominated by the CuO2 plane, since the bare transition temperature of the chain is much smaller than the plane owing to the strong fluctuations in one dimension [25, 26, 27, 28]. Therefore, although is measured at the chains, the peak from directly probes the charge dynamics of the CuO2 plane. To simplify our discussion and arrive at a qualitative understanding, we focus on only the CuO2 plane and leave the general discussion to the appendix. The newly observed peak is proportional to a -summed charge susceptibility of cuprate superconductivity, underscoring the significance of both long and short wavelengths to . Theoretically, can be written as
(3) |
where the is the structure factor of the quadrupole interaction. A key aspect of this equation is the coherence factor . For a d-wave superconductor, a prominent momentum q is around . This excitation, along with the corresponding coherence factor, has previously resulted in a distinct spin resonance peak in neutron scattering of cuprate superconductors[29]. Similarly, this coherence factor around gives rise to a sign reversal , which cannot counterbalance the divergence from DOS in the superconducting state. Hence, this may lead to this unconventional coherence peak in upon entering the superconducting transition.
III Discussion
The unconventional coherence peak identified in the nuclear quadrupole relaxation rate of YBa2Cu4O8 complements the relaxation rate pattern depicted in Fig. 4. In conventional superconductors, a Hebel-Slichter coherence peak is typically observed in the magnetic relaxation rate, as illustrated in Fig. 4 (a), whereas the quadrupole relaxation rate drops below due to coherence factors, as illustrated in Fig. 4 (b). It reflects that the gap is isotropic without a sign reversal in conventional superconductors, as shown in Fig. 4 (c). Conversely, unconventional superconducting gaps exhibit a sign reversal attributed to electron-electron correlation. In such cases, the magnetic relaxation rate does not display a coherence peak around experimentally, and the quadrupole relaxation rate acquires a coherence peak, as illustrated in Figs. 4(d) and 4(e). Historically, the absence of the Hebel-Slichter coherence peak in was widely discussed [8, 9, 7, 30, 31, 32, 33, 34]. One of the prevailing views is that the sign-changing gap is the reason for this absence, which eliminates the coherence factor contribution [31, 32, 30, 17]. Meanwhile, it will lead to a phenomenon that the be enhanced by the coherence factors just below due to the gap sign changes at various positions around Fermi surfaces, with scattering between different signs becoming dominant, as shown in Fig. 4 (f).
The sharp contrasts observed in and from conventional superconductors provide a novel method to explore the unconventional nature of the pairing mechanism in unconventional superconductors. The spin-singlet Cooper pairs are widely believed to be formed above owing to the small superfluid density in unconventional superconductors[35]. The unconventional peak observed in leads to a hallmark of forming phase coherent unconventional superconducting condensate below . This peak can be used to diagnose unconventional (sign-changing gap) superconductivity with significant quadrupole relaxation.
On the other hand, measurement conditions for are quite stringent, necessitating substantial EFG fluctuations, significant and the presence of two or more isotopes. Cuprates have charge-density wave (CDW) fluctuations in the phase diagram[36, 37], which provide opportunity to detect . Although CDW fluctuations contribute to , the peak found here cannot be due to CDW which competes with superconductivity and CDW order appears only when superconductivity is killed. Some cuprates contain chain Cu with large and two isotopes 63,65Cu. Underdoped YBa2Cu3O7-δ sharing similar electronic band structure with YBa2Cu4O8 can also be utilized to identify the unconventional coherence peak[38]. However, measuring in other unconventional superconductors poses challenges, such as in iron arsenide where only one isotope of 75As is present[39]. Theoretically, can be estimated by comparing the relaxation rate measured by NMR and NQR. However, the different principle axes and the external field make it hard to compare NMR and NQR directly[23]. Another method is to compare the relaxation rate of satellite peaks with the central peak. Nevertheless, achieving this requires exceedingly precise data that surpass current experimental precision levels[20]. As measurement precision improves, there may be opportunities to explore a wider range of systems. We hope our study will inspire extensive future experimental and theoretical investigations to elucidate whether the unconventional coherence peak is a universal characteristic of unconventional superconductivity and delve into its underlying mechanisms.
Acknowledgements.
We thank Professor Chengtian Lin of Max-Planck-Institute für Festkörperforschung for his help in sample synthesis. This work was supported by the National Key Research and Development Program of China (Grants No. 2022YFA1403903, No. 2022YFA1602800, and No. 2022YFA1403901), the National Natural Science Foundation of China (Grants No. 12134018, No. 12174428, No. 11888101, and No. 12488201), the Strategic Priority Research Program and Key Research Program of Frontier Sciences of the Chinese Academy of Sciences (Grant No. XDB33010100), the Chinese Academy of Sciences Project for Young Scientists in Basic Research (2022YSBR-048), the New Cornerstone Investigator Program, and the Synergetic Extreme Condition User Facility (SECUF).Appendix A METHODS
A.1 Sample growth and characterization
YBa2Cu4O8 single crystals are synthesized with YBa2Cu3O7-δ powders and the same molar ratio CuO (99.9%). The precursor YBa2Cu3O7-δ is prepared by solid-state reaction. Stoichiometric proportions of Y2O3(99.99%), BaCO3(99.99%), and CuO(99.9%) are thoroughly mixed and grounded and then calcined in air at ∘C for 24 h. The products are ground and then pressed into a pellet and calcined at 890 ∘C for 48 h in Ar(95%)-O2(5%). The prepared YBa2Cu3O7-δ powders and CuO are mixed uniformly and then put into the Al2O3 crucible with wt% KOH as flux. After keeping at 700 ∘C for 4 h, samples are cooled to 500 ∘C at a speed of 8 ∘C/h and fast cooled to room temperature finally. By soaking them in ethanol to remove the flux, we get small single crystals with a typical size of 0.1 mm. The x-ray diffraction pattern demonstrates the samples are YBa2Cu4O8, as shown in Fig. 5. The magnetic susceptibility measured with a magnetic property measurement system (MPMS-III) exhibits perfect diamagnetism, as shown in Fig. 6.
A.2 NQR measurements
The skin effect of the metallic state and the penetration depth of the superconducting state can shield the detection signal into the sample, so to achieve a large detectable volume we ground the samples and sieve with 300 mesh standard sieves to ensure a uniform powder particle size ( m). NQR measurements are carried out using a commercial NMR spectrometer from Thamway Co. Ltd. The NQR spectra are acquired by integrating the intensity of spin echo at each frequency, as shown in Fig. 7. The quadrupole resonance frequencies are summarized in Fig. 8. We notice that the of Cu(1) shows a small kink at and decreases below in Fig. 8. It is natural to ask whether this kink can influence the spin-lattice relaxation rate . First of all, the change in is equal to the static EFG change crossing . There are two possible origins for this static EFG change: (a) the electronic structure changes; (b) the coupling between electron and nuclear changes. Then, we can separate change into two cases:
-
•
If the static EFG change comes from the electrons, the only electronic changes that occur at are from the superconductivity. So, this case is equal to saying that the superconducting ordering induces both the change and peak crossing . And the change indicates that superconducting does influence the Cu(1) chain. In this case, is decided by superconducting transition but not .
-
•
If the static EFG change comes from the coupling, we can estimate this change to . As discussed in the theoretical analysis subsection, the spin relaxation rate is proportional to the relaxation matrix between a state and by . Hence, the static EFG influence to the by the . In other words, if we imagine the change of comes from a static coupling change instead of the electron fluctuations, the by the static EFG is changed by .
We know that just below is less than in Fig. 8. This means change of static EFG will change by in case (b). However, the in Fig. 2 (a) increases about just below which is much larger than this small change. Moreover, decreases monotonically with decreasing temperature below , while peaks just below and decreases at lower temperatures. So the peak of cannot come from change. The change of reflects the change of DOS and the appearance of the coherence factor in the superconducting state.
Spin-lattice relaxation rate for every nucleus is measured using a comb-shaped-pulse recovery method with the recovery function , where error bars are the standard error of the least square fit. is the nuclear magnetization at time t after saturation pulses. is the value of in an equilibrium state and , where is the initial value of after the saturation pulses. Both and are fitting parameters. in the full temperature zone is shown in Fig. 9. can be written as a sum of two contributions of and , for 63,65Cu NQR experiment. is proportional to the square of the gyromagnetic ratio, , so . is proportional to the square of the quadruple moment, , so . Based on these relationships, and can be distinguished.
We want to add a note here. The EFG principal axis of Cu(1) is along the axis and the principal axis of Cu(2) is along the axis[4]. In a conventional superconductor, the Hebel-Slichter coherence peak appears in all directions. The coherence peak reflects the divergence of DOS and coherence between electrons. Whether there is a coherence peak does not depend on the direction of the principal axes of EFG. However, if one wants to separate and by comparing NMR and NQR, or by comparing atoms at different sites, the principal axis is important. is determined by the fluctuations perpendicular to the applied magnetic field and is determined by the fluctuations perpendicular to the principal axes when . The mismatch between the applied magnetic field and the principle axes leads to different values of NMR from NQR. When , even if the applied field is along the principal axes, NMR and NQR cannot get the same value[23]. Moreover, the principal axes of Cu(1) and Cu(2) are different, so their from NQR cannot be compared directly[4]. We use the method of comparing of 63Cu and 65Cu at the same site, which have the same EFG and form factor, to avoid this problem.
A.3 Theoretical analysis
The Hamiltonian of quadrupole interaction can be compactly written as
(4) |
The EFG tensor can be expressed as where is the spatial function linking the electron density to nucleus quadrupole . Thus, the charge fluctuation in NQR is mainly determined by the density fluctuation of electrons and quadrupole of the nuclear. The Fourier transformed Hamiltonian in the lattice can be expressed as , where is the structure factor determined by and A contains the quadrupole moment and spin of nuclear. The relaxation of the nuclear spin and lattice toward the thermodynamic equilibrium can be described by the master equation , where is the population vector of the different energy levels [20]. The relaxation matrix is given by through second-order perturbation theory. Here is the density-density correlation function. For superconductors, can be calculated by the Green’s function as
(5) | |||||
where is the Pauli matrix and is the Matsubara Green’s function for SC. Then we can get
(6) |
As we discuss in the main text, the coherence peak is from the CuO2 plane. By focusing on the plane, Fig. 10 shows the quadrupole relaxation rate in d-wave superconductors calculated by summing the points near .
The above analysis provides a qualitative understanding of the coherence peak in -wave superconductors. On the other hand, since NQR relaxation rate can be observed only on the Cu(1) sites rather than on the Cu(2) sites due to a nonzero on CuO chains, we need consider a multiband model including both CuO planes and CuO chains to more convincingly describe the experimental phenomena. The Hamiltonian of the two-band model can be written as
(7) | |||||
The dispersions of the CuO chain () and CuO plane () are and respectively. The tunneling term between the two layers has the form . All the parameters in the tight-binding model can be obtained from the density-functional theory calculations. is the mean-field order parameter of the chain and plane bands. To our knowledge, how the chain becomes superconducting remains controversial. Since the failure of considering only the single-particle tunneling model has been highlighted in Ref. [25], we take the pair tunneling interaction, specifically Josephson coupling. The singlet pair tunneling term has the form
(8) |
where is the Josephson coupling strength. Because of symmetry requirements, Josephson coupling imposes the same pairing symmetry on both the CuO chains and planes. Thus, we set . Then can be determined by self-consistent equations
(9) | |||||
where are strengths of intralayer pairing potentials. Figure 11(a) shows the amplitude of order parameters on the CuO chain and plane. Because of Josephson coupling, the order parameter from the CuO planes penetrates into the chains. As a result, the phase coherence properties of the planes can also be detected in the chains. Note that if , the self-consistent calculation will yield a very small value for . Furthermore, we calculate the superfluid density along and directions shown in Fig.11(b). The quadrupole relaxation rate projected on the chain and plane bands can be calculated using this two-band model. Figure 11(c) and 11(d) show the results. The quadrupole relaxation rate on the CuO chain qualitatively captures the main features in the experiment: The coherence peak occurs just below . We emphasize that the conclusions do not depend on the specific form of interlayer coupling, as we also achieve the similar results using the single-particle tunneling model.
References
- Taillefer [2010] L. Taillefer, Scattering and pairing in cuprate superconductors, in Annual Review of Condensed Matter Physics, Vol 1, Annual Review of Condensed Matter Physics, Vol. 1, edited by J. S. Langer (2010) pp. 51–70.
- Imai et al. [1988a] T. Imai, T. Shimizu, T. Tsuda, H. Yasuoka, T. Takabatake, Y. Nakazawa, and M. Ishikawa, Nuclear Spin-Lattice Relaxation of 63,65Cu at the Cu(2) Sites of the High Superconductor YBa2Cu3O7-δ, Journal of the Physical Society of Japan 57, 1771 (1988a).
- Imai et al. [1988b] T. Imai, T. Shimizu, H. Yasuoka, Y. Ueda, and K. Kosuge, Anomalous Temperature Dependence of Cu Nuclear Spin-Lattice Relaxation in YBa2Cu3O6.91, Journal of the Physical Society of Japan 57, 2280 (1988b).
- Zimmermann et al. [1989] H. Zimmermann, M. Mali, D. Brinkmann, J. Karpinski, E. Kaldis, and S. Rusiecki, Copper NQR and NMR in the superconductor YBa2Cu4O8+x, Physica C: Superconductivity 159, 681 (1989).
- Warren and Walstedt [1990] W. W. Warren and R. E. Walstedt, NQR and NMR Studies of Spin Dynamics in High Superconducting Cuprates, Zeitschrift für Naturforschung A 45, 385 (1990).
- Machi et al. [1991] T. Machi, I. Tomeno, T. Miyatake, N. Koshizuka, S. Tanaka, T. Imai, and H. Yasuoka, Nuclear spin-lattice relaxation and Knight shift in YBa2Cu4O8, Physica C: Superconductivity 173, 32 (1991).
- Asayama et al. [1996] K. Asayama, Y. Kitaoka, G.-q. Zheng, and K. Ishida, NMR studies of high Tc superconductors, Progress in Nuclear Magnetic Resonance Spectroscopy 28, 221 (1996).
- Imai et al. [1995] T. Imai, C. P. Slichter, J. L. Cobb, and J. T. Markert, Superconductivity and spin fluctuations in the electron-doped infinitely-layered high Tc superconductor Sr0.9La0.1CuO2 ( K), Journal of Physics and Chemistry of Solids 56, 1921 (1995).
- Rigamonti et al. [1998] A. Rigamonti, F. Borsa, and P. Carretta, Basic aspects and main results of nmr-nqr spectroscopies in high-temperature superconductors, Reports on Progress in Physics 61, 1367 (1998).
- Jurkutat et al. [2019] M. Jurkutat, M. Avramovska, G. V. M. Williams, D. Dernbach, D. Pavicevic, and J. Haase, Phenomenology of 63Cu Nuclear Relaxation in Cuprate Superconductors, Journal of Superconductivity and Novel Magnetism 32, 3369 (2019).
- Hebel [1959] L. C. Hebel, Theory of Nuclear Spin Relaxation in Superconductors, Physical Review 116, 79 (1959).
- Slichter [1990] C. P. Slichter, Principles of Magnetic Resonance, Springer Series in Solid-State Sciences (Springer Berlin Heidelberg, 1990).
- Slichter [2007] C. P. Slichter, Magnetic resonance studies of high temperature superconductors, in Handbook of High-Temperature Superconductivity: Theory and Experiment, edited by J. R. Schrieffer and J. S. Brooks (Springer New York, New York, NY, 2007) pp. 215–256.
- MacLaughlin [1976] D. E. MacLaughlin, Magnetic resonance in the superconducting state, in Solid State Physics, Vol. 31, edited by F. S. Henry Ehrenreich and T. David (Academic Press, 1976) pp. 1–69.
- Wada and Asayama [1973] S. Wada and K. Asayama, Nuclear Quadrupole Spin-Lattice Relaxation of Ta181 in Type II Superconducting Ta3Sn, Journal of the Physical Society of Japan 34, 1168 (1973).
- Li et al. [2016] Z. Li, W. H. Jiao, G. H. Cao, and G.-q. Zheng, Charge fluctuations and nodeless superconductivity in quasi-one-dimensional Ta4Pd3Te16 revealed by 125Te-NMR and 181Ta-NQR, Phys. Rev. B 94, 174511 (2016).
- Xiang and Wu [2022] T. Xiang and C. Wu, D-wave Superconductivity (Cambridge University Press, Cambridge, 2022).
- Karpinski et al. [1988] J. Karpinski, E. Kaldis, E. Jilek, S. Rusiecki, and B. Bucher, Bulk synthesis of the 81-K superconductor YBa2Cu4O8 at high oxygen pressure, Nature 336, 660 (1988).
- Raffa et al. [1998] F. Raffa, T. Ohno, M. Mali, J. Roos, D. Brinkmann, K. Conder, and M. Eremin, Isotope Dependence of the Spin Gap in YBa2Cu4O8 as Determined by Cu NQR Relaxation, Physical Review Letters 81, 5912 (1998).
- Suter et al. [1998] A. Suter, M. Mali, J. Roos, and D. Brinkmann, Mixed magnetic and quadrupolar relaxation in the presence of a dominant static Zeeman Hamiltonian, Journal of Physics-Condensed Matter 10, 5977 (1998).
- Mangelschots et al. [1992] I. Mangelschots, M. Mali, J. Roos, D. Brinkmann, S. Rusiecki, J. Karpinski, and E. Kaldis, 17O NMR study in aligned YBa2Cu4O8 powder, Physica C: Superconductivity 194, 277 (1992).
- Suter et al. [2000] A. Suter, M. Mali, J. Roos, and D. Brinkmann, Charge degree of freedom and the single-spin fluid model in YBa2Cu4O8, Physical Review Letters 84, 4938 (2000).
- Goto et al. [1998] A. Goto, T. Shimizu, H. Aoki, M. Kato, K. Yoshimura, K. Kosuge, T. Matsumoto, and Y. Yamada, Anisotropy Study of the Spin-Lattice Relaxation Rates at the Cu(1) Chain Sites of YBa2Cu3O7 and YBa2Cu4O8, Journal of the Physical Society of Japan 67, 759 (1998).
- Raffa et al. [1999] F. Raffa, M. Mali, A. Suter, A. Y. Zavidonov, J. Roos, D. Brinkmann, and K. Conder, Spin and charge dynamics in the Cu-O chains of , Physical Review B 60, 3636 (1999).
- Xiang and Wheatley [1996] T. Xiang and J. M. Wheatley, Superfluid Anisotropy in YBCO: Evidence for Pair Tunneling Superconductivity, Physical Review Letters 76, 134 (1996).
- Serafin et al. [2010] A. Serafin, J. D. Fletcher, S. Adachi, N. E. Hussey, and A. Carrington, Destruction of chain superconductivity in in a weak magnetic field, Physical Review B 82, 140506 (2010).
- Atkinson and Carbotte [1995] W. A. Atkinson and J. P. Carbotte, Effect of proximity coupling of chains and planes on the penetration-depth anisotropy in , Physical Review B 52, 10601 (1995).
- Gagnon et al. [1997] R. Gagnon, S. Pu, B. Ellman, and L. Taillefer, Anisotropy of Heat Conduction in : A Probe of Chain Superconductivity, Physical Review Letters 78, 1976 (1997).
- Fong et al. [1995] H. F. Fong, B. Keimer, P. W. Anderson, D. Reznik, F. Doğan, and I. A. Aksay, Phonon and Magnetic Neutron Scattering at 41 meV in YBa2Cu3O7, Physical Review Letters 75, 316 (1995).
- Koyama and Tachiki [1989] T. Koyama and M. Tachiki, Theory of nuclear relaxation in superconducting high- oxides, Phys. Rev. B 39, 2279 (1989).
- Monien and Pines [1990] H. Monien and D. Pines, Spin excitations and pairing gaps in the superconducting state of YBa2Cu3O7-δ, Phys. Rev. B 41, 6297 (1990).
- Thelen et al. [1993] D. Thelen, D. Pines, and J. P. Lu, Evidence for pairing from nuclear-magnetic-resonance experiments in the superconducting state of , Phys. Rev. B 47, 9151 (1993).
- Zoli [1991] M. Zoli, Smearing of the Hebel-Slichter Peak by 2D Fluctuationsin Layered Cuprate Superconductors, Journal of the Physical Society of Japan 60, 3837 (1991).
- Statt [1990] B. W. Statt, Anisotropic gap and quasiparticle-damping effects on NMR measurements of high-temperature superconductors, Physical Review B 42, 6805 (1990).
- Emery and Kivelson [1995] V. J. Emery and S. A. Kivelson, Importance of phase fluctuations in superconductors with small superfluid density, Nature 374, 434 (1995).
- Keimer et al. [2015] B. Keimer, S. A. Kivelson, M. R. Norman, S. Uchida, and J. Zaanen, From quantum matter to high-temperature superconductivity in copper oxides, Nature 518, 179 (2015).
- Wu et al. [2011] T. Wu, H. Mayaffre, S. Kramer, M. Horvatic, C. Berthier, W. N. Hardy, R. X. Liang, D. A. Bonn, and M. H. Julien, Magnetic-field-induced charge-stripe order in the high-temperature superconductor YBa2Cu3Oy, Nature 477, 191 (2011).
- Oguchi et al. [1990] T. Oguchi, T. Sasaki, and K. Terakura, Electronic band structure of YBa2Cu4O8, Physica C: Superconductivity 172, 277 (1990).
- Li et al. [2011] Z. Li, D. L. Sun, C. T. Lin, Y. H. Su, J. P. Hu, and G. Q. Zheng, Nodeless energy gaps of single-crystalline Ba0.68K0.32Fe2As2 as seen via 75As NMR, Physical Review B 83, 140506 (2011).