Next Article in Journal
Cellulose Acetate–PHB Biocomposite from Saccharum officinarum for Ni (II) Adsorption: Equilibrium and Kinetics
Previous Article in Journal
Investigating the Mechanical and Thermal Performance of HDPE Composites Based on Nano-Graphite Particles
Previous Article in Special Issue
The Effect of Doping with Aluminum on the Optical, Structural, and Morphological Properties of Thin Films of SnO2 Semiconductors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microscopic Characterization of Pb10−xCux(PO4)6O by 31P and 63/65Cu NMR Measurements

1
Ames National Laboratory, U.S. DOE, and Department of Physics and Astronomy, Iowa State University, Ames, IA 50011, USA
2
Key Laboratory of Quantum Materials and Devices of Ministry of Education, School of Physics, Southeast University, Nanjing 211189, China
*
Author to whom correspondence should be addressed.
J. Compos. Sci. 2025, 9(7), 377; https://doi.org/10.3390/jcs9070377
Submission received: 18 June 2025 / Revised: 11 July 2025 / Accepted: 14 July 2025 / Published: 18 July 2025
(This article belongs to the Special Issue Optical–Electric–Magnetic Multifunctional Composite Materials)

Abstract

The report of the first room-temperature, ambient-pressure superconductivity in copper-doped lead apatite Pb10−xCux(PO4)6O has attracted lots of attention. However, subsequent studies revealed the presence of numerous impurity phases in the polycrystalline sample, and the sharp superconducting-like transition is not due to a superconducting transition but most likely due to a reduction in resistivity caused by the first-order structural phase transition of Cu2S at around 385 K from the β phase at high temperature to the γ phase at low temperature. Before now, only bulk measurements have been performed on a Pb10−xCux(PO4)6O powder sample, which could be affected by the impurity phases, masking the intrinsic properties of Pb10−xCux(PO4)6O. In this study, 31P and 63/65Cu nuclear magnetic resonance (NMR) measurements have been performed on a Pb10−xCux(PO4)6O powder sample to investigate its physical properties from a microscopic point of view. Our NMR data evidence the non-magnetic insulating nature of Pb10−xCux(PO4)6O without any trace of electron correlation effects. Furthermore, the 63/65Cu NMR results suggest that no copper or very little copper is substituted for Pb in Pb10(PO4)6O prepared by sintering Pb2SO5 and Cu3P.

1. Introduction

Recently, copper-doped lead apatite Pb10−xCux(PO4)6O has been proposed as a candidate room-temperature and ambient-pressure superconductor from resistivity and levitation measurements [1,2,3], which has stimulated a lot of experimental and theoretical studies on this material. However, the resistivity is not actually zero but around 2 × 10−3 Ωcm at the temperature just below the claimed superconducting transition temperature [3], making the claim doubtful. Experimentally, impurity phases have been reported to be unavoidable from the solid-state reaction methods [2,4,5,6]. Subsequent studies indicate that the sharp superconducting-like transition is most likely due to a reduction in resistivity caused by the first-order structural phase transition of Cu2S at around 385 K, from the high temperature β phase to the γ phase at a low temperature, and Pb10−xCux(PO4)6O shows an insulating behavior [7,8]. It has also been suggested that the original reported diamagnetic behavior and levitation experiments may arise from either diamagnetism or small ferromagnetic impurities [9,10,11], but not due to superconductivity.
Even though the parent compound Pb10(PO4)6O was considered to be an insulator, theoretical studies based on density functional theory (DFT) calculations suggest that Cu doping will create flat bands and induce insulator–metal transition, leading to a metallic state in Cu-substituted Pb10(PO4)6O [12,13,14,15,16,17,18,19,20]. However, an ab initio calculation study based on strong-coupling Migdal–Eliashberg theory indicates that the electron–phonon coupling strength is insufficient to overcome the Coulomb repulsion between electron pairs and thus does not support high-temperature superconductivity in Pb9Cu(PO4)6O via the conventional electron–phonon Migdal–Eliashberg mechanism [21]. Even neglecting Coulomb repulsion of the electron pairs, a superconducting transition temperature was suggested to be less than 2 K assuming a phonon-mediated superconducting mechanism [21]. Thus, if high-temperature superconductivity were possible in this system, the mechanism of superconductivity would be unconventional beyond the BCS theory. In fact, recent self-consistent many-body perturbation theory predicts strong spin fluctuations in Pb10−xCux(PO4)6O [22]. Furthermore, antiferromagnetic fluctuations for the undoped material were suggested by a fluctuation exchange calculation study, which also suggests ferromagnetic fluctuations at a very low electron doping level [23]. On the contrary, another theoretical study suggests that orbital and spin fluctuations are weak, not enough to support superconductivity at high temperatures [24]. Experimentally, a logarithmic temperature dependence of resistivity ρ ∝ ln(1/T) in a wide temperature range of 10–300 K has been observed and was ascribed to the strong correlation effect in this system [25]. However, due to the multiphases in the compounds, the results of bulk measurements usually contain contributions from all the phases, and it may not be easy to distinguish and separate each contribution. Thus, the nature of the magnetic fluctuations of Pb10−xCux(PO4)6O is still an open question.
Nuclear magnetic resonance (NMR) is a powerful technique to investigate static spin susceptibility and low-energy spin excitations from a microscopic point of view. The temperature dependence of the nuclear spin–lattice relaxation rate (1/T1) reflects the wave vector q-summed dynamical susceptibility. On the other hand, NMR spectrum measurements, in particular, the NMR shift K, give us information on static magnetic susceptibility. Thus, from the temperature dependence of 1/T1 and K, one can microscopically obtain valuable insights into the electronic and magnetic properties. Here, we performed 31P and 63/65Cu NMR on a Pb10−xCux(PO4)6O powder sample to reveal its intrinsic physical properties. 31P NMR results indicate that Pb10−xCux(PO4)6O is a non-magnetic insulator without any obvious magnetic fluctuations. 63/65Cu NMR results reveal that no copper or very little copper substitutes for Pb in Pb10(PO4)6O prepared using the solid-state reaction of Pb2SO5 and Cu3P.

2. Materials and Methods

2.1. Material Preparation

The Pb10−xCux(PO4)6O powder sample was prepared using a solid-state reaction of Pb2SO5 and Cu3P [6]. The uniformly mixed pellets of Pb2SO5 and Cu3P were sealed in a high-vacuum quartz tube and raised to a temperature of 925 °C at a rate of 1.5 °C/min, held for 20 h, and then cooled in a furnace to obtain the target product Pb10−xCux(PO4)6O. Aside from the main phase Pb10−xCux(PO4)6O with x to be determined, copper metal and Cu2S have been found to be impurity phases from the X-ray diffraction measurements [6].

2.2. Nuclear Magnetic Resonance

NMR measurements of 31P (I = 1 2 , Υ N 2 π = 17.2356 MHz/T), 63Cu (I = 3 2 , Υ N 2 π = 11.285 MHz/T), and 65Cu (I = 3 2 , Υ N 2 π = 12.089 MHz/T) nuclei were conducted using a laboratory-built, phase-coherent spin-echo pulse spectrometer. A high-homogeneity magnetic field was generated by a custom-built magnet from Oxford Instruments. The power amplifiers from JEOL Ltd. (Tokyo, Japan) were used to generate adjustable radio frequency pulses to excite the nuclear-spin system, while the pulse generator and the NMR probe were lab-built. The 31P, 63Cu, and 65Cu NMR spectra were obtained by Fourier transform of the Hahn spin-echo signals at magnetic fields μ0H of 7.3685 T, 7.3685 T, and 6.8784 T, respectively. To avoid the 63/65Cu NMR contribution of an NMR coil, a silver coil was used. The 1/T1 was measured with a saturation recovery method. 1/T1 at each temperature was determined by fitting the nuclear magnetization M versus time t using the exponential function 1 − M(t)/M(∞) = e ( t T 1 ) β , where M(t) and M(∞) are the nuclear magnetization at time t after the saturation and the equilibrium nuclear magnetization at t → ∞, respectively. Here, β is the stretching exponent. A fit with β < 1 indicates a distribution of relaxation rates. For 31P NMR, the recovery curves were fit with a nearly temperature-independent value of β~0.65. On the other hand, single exponential behaviors of the nuclear recovery curves were found for the 63/65Cu NMR measurements. The nuclear spin–spin relaxation rate 1/T2 for 31P at each temperature was determined by fitting the decay curve of nuclear magnetization M measured at 2τ using the exponential function M(2τ) = M(0) e ( 2 τ T 2 ) β , where τ is the time interval between the first π/2 pulse and the second π pulse, M(2τ) is the nuclear magnetization for pulse separation τ, M(0) is a fitting parameter corresponding to the nuclear magnetization at time τ = 0, and β′ is the stretching exponent. β′ was found to be around 1.5 and nearly temperature-independent.

3. Results and Discussion

3.1. 31P NMR Spectra

Figure 1a shows the temperature dependence of the 31P NMR spectra of Pb10−xCux(PO4)6O measured under μ0H = 7.3685 T. The Larmor frequency f0 is also shown in the figure. The observed spectra are relatively sharp, whose full width at half maximum (FWHM) is 6.41 kHz (= 3.72 Oe) at 300 K and increases slightly with decreasing temperature, as shown in the inset in Figure 1b. The peak position of the 31P NMR spectra shifts very little as temperature decreases. The temperature dependence of NMR shift for 31P, 31K, determined by the peak position, is shown in Figure 1b, where 31K decreases slightly as the temperature lowers from −0.009(1)% at room temperature to −0.016(1)% at 4.2 K. Since the NMR shift is related to static magnetic susceptibility (χ), the small negative 31K could be due to the diamagnetism observed in the magnetic susceptibility of this material [6]. It is noted that, although the magnetic susceptibility shows an increase at low temperatures, we did not observe such a change in 31K corresponding to the increase in χ at low temperatures [6,11]. Thus, we conclude that the increase in magnetic susceptibility at low temperatures is not intrinsic but extrinsic, coming from impurity phases.
As we will show below, almost no Pb atoms are replaced by Cu atoms in our powder sample Pb10−xCux(PO4)6O (that is, x~0). The small magnitude of 31K suggests no obvious magnetism in our sample and a non-magnetic semiconducting ground state of Pb10-xCux(PO4)6O with x~0, where 31K is expected to be temperature-independent. The small temperature dependence of 31K, therefore, could originate from carriers due to thermal excitation. In semiconductors, the NMR shift is known to be proportional to T1/2*exp(−∆/2kBT), where ∆ is the magnitude of the energy gap between the top of the valence band and the bottom of the conduction band [26]. However, the temperature dependence of 31K could not be reproduced by this formula. Instead, we found that the temperature dependence of 31K can be reproduced by a simple thermal activation form of 31K = K0 + A × exp(−∆/kBT) with K0 = −0.0163%, A = 0.0105%, and ∆/kB = 110 K, as shown by the solid curve in Figure 1b. Here, K0 is the temperature-independent part of 31K, which originates from diamagnetism. At present, we do not have any clear idea of how to explain the temperature dependence of 31K. However, similar temperature dependence of the NMR shift has been observed in several semiconductors, such as PbTe [27], and topological insulators (Bi,Sb)(Te,Se,S) [28,29], where the thermal activation behavior is considered to originate from the thermally excited carriers. It is noted that the magnitude of ∆~110(5) K (=9.5 meV) is much smaller than the band gap (200–500 meV) between the valence and conduction bands reported from the temperature dependence of resistivity measurements [11,30]. Thus, the small activation energy may suggest the existence of an impurity band between the valence and conduction bands.

3.2. Nuclear Spin–Lattice Relaxation Rate 1/31T1 and Nuclear Spin–Spin Relaxation Rate 1/31T2

We also carried out measurements of the nuclear spin–lattice relaxation rate 1/T1 of 31P NMR to derive further information on the physical properties of Pb10−xCux(PO4)6O. The inset in Figure 2a shows typical recovery curves at 100 and 300 K, where the solid curves are the fits when using the stretched exponential function with 1/T1 = 0.05 (1/s) and β = 0.63(4), and 1/T1 = 0.02 (1/s) and β = 0.66(2) for T = 100 K and 300 K, respectively. The temperature dependence of the estimated 1/T1 is shown in Figure 2a, where 1/T1 decreases gradually from 300 K to ~50 K and then becomes nearly independent of temperature at low temperatures. The relatively small values of 1/T1 suggest a non-magnetic insulating ground state without any obvious magnetic fluctuations in Pb10−xCux(PO4)6O, consistent with 31K data shown above.
In semiconductors, the temperature dependence of 1/T1 is expected to be 1/T1~T2*exp(−∆/2kBT) [26]. However, the experimental data cannot be reproduced by the formula at all. Instead, similar to the case of 31K, we found that the 1/T1 data between ~60 K and 300 K can be roughly reproduced by the simple thermal activation form of 1/T1 = A′ × exp(−∆/kBT) with A′ = 0.07 (1/s) and ∆/kB = 110(20) K, as shown by the solid curve in Figure 2a. Below ~60 K, 1/31T1 deviates from the fitting; although we do not know the origin, it is most likely due to a small quantity of impurities, which may produce a small increase in FWHM at low temperatures, as observed.
As shown above, our NMR spectra and 1/T1 measurements of 31P indicate that the system has a non-magnetic insulating ground state without any phase transitions below 300 K. We did not find any trace of phase transitions either in the temperature dependence of the 31P spin–spin relaxation rate 1/31T2 shown in Figure 2b, where 1/31T2 increases slightly and smoothly with a decreasing temperature, evidencing no phase transition below 300 K.
Several theoretical studies indicate that the dominant contribution to the flat bands in Pb10−xCux(PO4)6O is from Cu d-electrons [12,13,14,15,16,17,18,19,20]. Therefore, since magnetic fluctuations, if they exist, are expected to originate from the Cu d electrons, NMR measurements on Cu sites are very important. Thus, we carried out 63/65Cu NMR measurements on the Pb10−xCux(PO4)6O powder sample. We observed very sharp 63/65Cu NMR lines, as shown in the inset of Figure 3a. We examined the observed 63/65NMR lines from the temperature dependencies of NMR shifts (Figure 3a) and 1/T1 (Figure 3b) measurements. The observed NMR shift of K~0.23% and 1/T1 data (1/T1T = 0.79 (1/sK) for 63Cu and 1/T1T = 0.91 (1/sK) for 65Cu) are the same as those in Cu metals [31]. Therefore, we conclude that the observed Cu signals are from Cu metals, not from Cu ions in Pb10-xCux(PO4)6O if any Cu atoms exist. We also tried to search for any other 63/65Cu NMR signal; however, no other NMR signal was detected. Thus, our NMR results suggest that no copper or only very little copper substitutes for Pb in our Pb10−xCux(PO4)6O powder sample, which was synthesized by sintering Pb2SO5 and Cu3P.
It is worth mentioning that we do not observe any 63/65Cu NMR signals from Cu3P in our powder sample. This is consistent with no observation of any trace of 31P NMR signals from a metallic Cu3P in which K = 0.012% and 1/T1T = 0.21 (1/sK) for 31P NMR were reported [32]. Thus, we consider that most starting materials of Cu3P reacted in forming Cu metal or Cu2S. This conclusion is supported by the fact that a lot of copper metal was observed in the sample studied here [6], and by the observation of other copper compounds in the sample prepared by similar preparation methods [33,34]. Actually, recent studies based on DFT suggest that, since the substitutional formation energy is large, Cu substitution is highly thermodynamically disfavored and the incorporation of Cu at the Pb sites proves to be extremely difficult [18]. Here, our NMR measurements revealed that almost no Pb atoms are replaced by Cu atoms in our Pb10−xCux(PO4)6O powder sample (that is, x~0), consistent with the theoretical prediction [18]. It is also noted that no 63/65Cu NMR signals from Cu2S were detected in our measurements, suggesting that not much Cu2S is present in our powder sample. Quite recently, single crystals of Pb10−xCux(PO4)6O have been successfully synthesized via the traveling solvent floating zone growth method [35]. The measurements on the single crystals confirmed the non-magnetic insulating state of Pb10−xCux(PO4)6O. In the single crystals, EDS measurements indicate that there are Cu atoms in the single crystal, but x was found to be around 0.5 on average with a very large distribution of copper content between 0.1 and ~1 throughout the sample. As described above, no observation of magnetic fluctuations has been found in our Pb10−xCux(PO4)6O powder sample with x~0, and this is most likely due to the lack of Cu substitution. Therefore, it is interesting to perform NMR measurements on single crystals to see whether magnetic fluctuations exist or not, even though x is distributed.

4. Conclusions

In summary, we have performed 31P and 63/65Cu NMR measurements on a Pb10−xCux(PO4)6O powder sample prepared by sintering Pb2SO5 and Cu3P. From the results of 63/65Cu NMR measurements, we have concluded that most starting materials of Cu3P react in forming Cu metal or Cu2S and that the value of x in our Pb10−xCux(PO4)6O is close to 0, showing that almost no Pb atoms are replaced by Cu atoms with this sample preparation method. From 31P NMR measurements, the non-magnetic insulating state has been revealed for Pb10−xCux(PO4)6O with x~0. Thus, we consider that most previous experimental studies on Pb10−xCux(PO4)6O powder samples might only have measured Pb10(PO4)6O or one with very little copper doping. Further studies are urgently required to shed light on the intrinsic physical properties of Cu-substituted Pb10−xCux(PO4)6O.

Author Contributions

Conceptualization, Q.-P.D.; formal analysis, Q.-P.D. and Y.F.; investigation, Q.-P.D., Y.S., Q.H., W.W., X.Z., X.W., Z.S. and Y.F.; data curation, Q.-P.D.; writing—original draft preparation, Q.-P.D.; writing—review and editing, Q.-P.D. and Y.F.; funding acquisition, Y.S. and Y.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the U.S. Department of Energy (DOE), Office of Basic Energy Sciences, Division of Materials Sciences and Engineering. Ames National Laboratory is operated for the U.S. DOE by Iowa State University under Contract No. DE-AC02-07CH11358. This work was also supported by the National Natural Science Foundation of China (Grants No. 12374135 and No. 12374136).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Lee, S.; Kim, J.; Im, S.; An, S.; Kwon, Y.W.; Ho, A.K. Consideration for the development of room-temperature ambient-pressure superconductor (LK-99). J. Korean Cryst. Growth Cryst. Technol. 2023, 33, 61. [Google Scholar]
  2. Lee, S.; Kim, J.H.; Kwon, Y.W. The First Room-Temperature Ambient-Pressure Superconductor. arXiv 2023, arXiv:2307.12008. [Google Scholar]
  3. Lee, S.; Kim, J.; Kim, H.T.; Im, S.; An, S.; Auh, K.H. Superconductor Pb10-xCux(PO4)6O showing levitation at room temperature and atmospheric pressure and mechanism. arXiv 2023, arXiv:2307.12037. [Google Scholar]
  4. Liu, L.; Meng, Z.; Wang, Z.X.; Chen, H.; Duan, Z.; Zhou, X.; Yan, H.; Qin, P.; Liu, Z. Semiconducting Transport in Pb10−XCux(PO4)6O Sintered from Pb2SO5 and Cu3P. Adv. Funct. Mater. 2023, 33, 2308938. [Google Scholar] [CrossRef]
  5. Kumar, K.; Karn, N.K.; Awana, V.P.S. Synthesis of possible room temperature superconductor LK-99: Pb9Cu(PO4)6O. Supercond. Sci. Technol. 2023, 36, 10LT02. [Google Scholar] [CrossRef]
  6. Hou, Q.; Wei, W.; Zhou, X.; Wang, X.; Wang, T.; Sun, Y.; Shi, Z. Synthesis, transport and magnetic properties of Cu-doped apatite Pb10−xCux(PO4)6O. Matter 2023, 6, 4408–4418. [Google Scholar] [CrossRef]
  7. Zhu, S.; Wu, W.; Li, Z.; Luo, J. First-order transition in LK-99 containing Cu2S. Matter 2023, 6, 4401–4407. [Google Scholar] [CrossRef]
  8. Jain, P.K. Superionic Phase Transition of Copper(I) Sulfide and Its Implication for Purported Superconductivity of LK-99. J. Phys. Chem. C 2023, 127, 18253–18255. [Google Scholar] [CrossRef]
  9. Guo, K.; Li, Y.; Jia, S. Ferromagnetic half levitation of LK-99-like synthetic samples. Sci. China Phys. Mech. Astron. 2023, 66, 107411. [Google Scholar] [CrossRef]
  10. Wu, H.; Yang, L.; Xiao, B.; Chang, H. Successful growth and room temperature ambient-pressure magnetic levitation of LK-99. arXiv 2023, arXiv:2308.01516. [Google Scholar]
  11. Wang, P.; Liu, X.; Ge, J.; Ji, C.; Ji, H.; Liu, Y.; Ai, Y.; Ma, G.; Qi, S.; Wang, J. Ferromagnetic and insulating behavior in both half magnetic levitation and non-levitation LK-99 like samples. Quantum Front. 2023, 2, 10. [Google Scholar] [CrossRef]
  12. Lai, J.; Li, J.; Liu, P.; Sun, Y.; Chen, X.Q. First-principles study on the electronic structure of Pb10−xCux(PO4)6O (x = 0, 1). J. Mater. Sci. Technol. 2024, 171, 66. [Google Scholar] [CrossRef]
  13. Griffin, S.M. Origin of correlated isolated flat bands in copper-substituted lead phosphate apatite. arXiv 2023, arXiv:2307.16892. [Google Scholar]
  14. Si, L.; Held, K. Electronic structure of the putative room-temperature superconductor Pb9Cu(PO4)6O. Phys. Rev. B 2023, 108, L121110. [Google Scholar] [CrossRef]
  15. Kurleto, R.; Lany, S.; Pashov, D.; Acharya, S.; Schilfgaarde, M.V.; Dessau, D.S. Pb-apatite framework as a generator of novel flat-band CuO based physics. arXiv 2023, arXiv:2308.00698. [Google Scholar]
  16. Cabezas-Escares, J.; Barrera, N.F.; Cardenas, C.; Munoz, F. Electronic structure and vibrational stability of copper-substituted lead apatite LK-99. Phys. Rev. B 2024, 109, 144515. [Google Scholar] [CrossRef]
  17. Tao, K.; Chen, R.; Yang, L.; Gao, J.; Xue, D.; Jia, C. The 1/4 occupied O atoms induced ultra-flatband and the one-dimensional channels in the Pb10−xCux(PO4)6O4 (x = 0, 0.5) crystal. APL Mater. 2024, 12, 021120. [Google Scholar] [CrossRef]
  18. Jiang, Y.; Lee, S.B.; Herzog-Arbeitman, J.; Yu, J.; Feng, X.; Hu, H.; Călugăru, D.; Brodale, P.S.; Gormley, E.L.; Vergniory, M.G.; et al. Pb9Cu(PO4)6(OH)2: Phonon bands, localized flat-band magnetism, models, and chemical analysis. Phys. Rev. B 2023, 108, 235127. [Google Scholar] [CrossRef]
  19. Liu, R.; Guo, T.; Lu, J.; Ren, J.; Ma, T. Different phase leads to different transport behavior in Pb9Cu(PO4)6O compounds. Int. J. Mod. Phys. C 2025, 36, 7. [Google Scholar] [CrossRef]
  20. Yang, S.; Liu, G.; Zhong, Y. Ab initio investigations on the electronic properties and stability of Cu-substituted lead apatite (LK-99) family with different doping concentrations (x = 0, 1, 2). Mater. Today Commun. 2023, 37, 107379. [Google Scholar] [CrossRef]
  21. Paudyal, H.; Flatt’e, M.E.; Paudyal, D. Implications of the electron-phonon coupling in CuPb9(PO4)6O for superconductivity: An ab initio study. Phys. Rev. Mater. 2024, 8, L011801. [Google Scholar] [CrossRef]
  22. Pashov, D.; Acharya, A.; Lany, S.; Dessau, D.S.; Schilfgaarde, M.V. Multiple Slater determinants and strong spin-fluctuations as key ingredients of the electronic structure of electron- and hole-doped Pb10−xCux(PO4)6O. arXiv 2023, arXiv:2308.09900. [Google Scholar]
  23. Shimizu, M.; Otsuki, J.; Jeschke, H.O. Magnetic fluctuations in Pb9Cu(PO4)6O. Phys. Rev. B 2023, 108, L201105. [Google Scholar] [CrossRef]
  24. Witt, N.; Si, L.; Tomczak, J.M.; Held, K.; Wehling, T.O. No superconductivity in Pb9Cu1(PO4)6O found in orbital and spin fluctuation exchange calculations. SciPost Phys. 2023, 15, 197. [Google Scholar] [CrossRef]
  25. Zhang, Y.; Liu, C.; Zhu, X.; Wen, H.-H. Ferromagnetism and insulating behavior with a logarithmic temperature dependence of resistivity in Pb10−xCux(PO4)6O. Sci. China Phys. Mech. Astron. 2024, 67, 217413. [Google Scholar] [CrossRef]
  26. Wolf, D. Spin Temperature and Nuclear-Spin Relaxation in Matter, 1st ed.; Oxford University Press: Oxford, UK, 1979. [Google Scholar]
  27. Taylor, R.E.; Alkan, F.; Koumoulis, D.; Lake, M.P.; King, D.; Dyowski, C.; Bouchard, L.-S. A Combined NMR and DFT Study of Narrow Gap Semiconductors: The Case of PbTe. J. Phys. Chem. C 2013, 117, 8959–8967. [Google Scholar] [CrossRef]
  28. Koumoulis, D.; Leung, B.; Chasapis, T.C.; Taylor, R.; King, D.; Kanatzidis, M.G.; Bouchard, L.-S. Understanding Bulk Defects in Topological Insulators from Nuclear-Spin Interactions. Adv. Funct. Mater. 2014, 24, 1519–1528. [Google Scholar] [CrossRef]
  29. Park, J.K.; Kang, D.H.; Park, S.K.; Lee, J.S. NMR probe of suppressed bulk conductivity in the topological insulator Bi0.5Sb1.5Te3. RSC Adv. 2022, 12, 2531. [Google Scholar] [CrossRef]
  30. Singh, M.; Saha, P.; Kumar, K.; Takhar, D.; Birajdar, B.; Awana, V.P.S.; Patnaik, S. Electromagnetic properties of copper doped lead apatite Pb10−xCux(PO4)6O. J. Mater. Sci. 2024, 59, 1464–1471. [Google Scholar] [CrossRef]
  31. Carter, G.C.; Bemmett, L.H.; Karha, D.J. Metallic Shifts in NMR: A Review of the Theory and Comprehensive Critical Data Compilation of Metallic Materials; Pergamon Press: Oxford, UK; New York, NY, USA; Toronto, ON, Canada; Sidney, Australia; Paris, France; Frankfurt, Germany, 1977. [Google Scholar]
  32. Furo, I.; Bakonyi, I.; Tompa, K.; Zsoldos, E.; Heinmaa, I.; Alla, M.; Lippmma, E. 31P nuclear magnetic resonance Knight shift and linewidth in Ni3P and Cu3P: A magic-angle spinning study. J. Phys. Condens. Matter 1990, 2, 4217. [Google Scholar] [CrossRef]
  33. Liu, C.; Cheng, W.; Zhang, X.; Xu, J.; Li, J.; Shi, Q.; Yuan, C.; Xu, L.; Zhou, H.; Zhu, S.; et al. Phases and magnetism at microscale in compounds containing nominal Pb10−xCux(PO4)6O. Phys. Rev. Mater. 2023, 7, 084804. [Google Scholar] [CrossRef]
  34. Kumar, K.; Karn, N.K.; Kumar, Y.; Awana, V.P.S. Absence of Superconductivity in LK-99 at Ambient Conditions. ACS Omega 2023, 8, 41737–41743. [Google Scholar] [CrossRef]
  35. Puphal, P.; Akbar, M.Y.P.; Hepting, M.; Goering, E.; Isobe, M.; Nugroho, A.A.; Keimer, B. Single crystal synthesis, structure, and magnetism of Pb10−xCux(PO4)6O. APL Mater. 2023, 11, 101128. [Google Scholar] [CrossRef]
Figure 1. (a) Temperature dependence of the 31P NMR spectra and (b) 31K of Pb10−xCux(PO4)6O. Inset in (b): Temperature dependence of the FWHM of the NMR spectra.
Figure 1. (a) Temperature dependence of the 31P NMR spectra and (b) 31K of Pb10−xCux(PO4)6O. Inset in (b): Temperature dependence of the FWHM of the NMR spectra.
Jcs 09 00377 g001
Figure 2. (a) Temperature dependence of 31T1−1 of Pb10−xCux(PO4)6O. The solid line is the fitting result with a thermal activation form 1/T1 = A′ × exp(−∆/kBT), with A′ = 0.07 s−1 and ∆/kB = 110 K. Inset: Recovery curves at T = 100 K and 300 K. Solid lines are fit by the relaxation function described in the text. (b) Temperature dependence of 31T2−1 of Pb10−xCux(PO4)6O. Inset: Relaxation curve at T = 10 K. The solid line is fit by the relaxation function described in the text.
Figure 2. (a) Temperature dependence of 31T1−1 of Pb10−xCux(PO4)6O. The solid line is the fitting result with a thermal activation form 1/T1 = A′ × exp(−∆/kBT), with A′ = 0.07 s−1 and ∆/kB = 110 K. Inset: Recovery curves at T = 100 K and 300 K. Solid lines are fit by the relaxation function described in the text. (b) Temperature dependence of 31T2−1 of Pb10−xCux(PO4)6O. Inset: Relaxation curve at T = 10 K. The solid line is fit by the relaxation function described in the text.
Jcs 09 00377 g002
Figure 3. 63/65K and 1/63/65T1 in Pb10−xCux(PO4)6O. (a) Temperature dependencies of NMR shifts 63/65K. Inset: 63/65Cu NMR spectra measured at a frequency of 83.34389 MHz and T = 4.3 K. (b) Temperature dependence of 1/63/65T1. The solid lines are linear fits with (1/T1 = 0.79 T(1/sK) for 63Cu and 1/T1 = 0.91 T(1/sK) for 65Cu). Inset: Recovery curves for 63Cu and 65Cu at T = 4.3 K. Solid lines are fits by the relaxation function described in the text.
Figure 3. 63/65K and 1/63/65T1 in Pb10−xCux(PO4)6O. (a) Temperature dependencies of NMR shifts 63/65K. Inset: 63/65Cu NMR spectra measured at a frequency of 83.34389 MHz and T = 4.3 K. (b) Temperature dependence of 1/63/65T1. The solid lines are linear fits with (1/T1 = 0.79 T(1/sK) for 63Cu and 1/T1 = 0.91 T(1/sK) for 65Cu). Inset: Recovery curves for 63Cu and 65Cu at T = 4.3 K. Solid lines are fits by the relaxation function described in the text.
Jcs 09 00377 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ding, Q.-P.; Sun, Y.; Hou, Q.; Wei, W.; Zhou, X.; Wang, X.; Shi, Z.; Furukawa, Y. Microscopic Characterization of Pb10−xCux(PO4)6O by 31P and 63/65Cu NMR Measurements. J. Compos. Sci. 2025, 9, 377. https://doi.org/10.3390/jcs9070377

AMA Style

Ding Q-P, Sun Y, Hou Q, Wei W, Zhou X, Wang X, Shi Z, Furukawa Y. Microscopic Characterization of Pb10−xCux(PO4)6O by 31P and 63/65Cu NMR Measurements. Journal of Composites Science. 2025; 9(7):377. https://doi.org/10.3390/jcs9070377

Chicago/Turabian Style

Ding, Qing-Ping, Yue Sun, Qiang Hou, Wei Wei, Xin Zhou, Xinyue Wang, Zhixiang Shi, and Yuji Furukawa. 2025. "Microscopic Characterization of Pb10−xCux(PO4)6O by 31P and 63/65Cu NMR Measurements" Journal of Composites Science 9, no. 7: 377. https://doi.org/10.3390/jcs9070377

APA Style

Ding, Q.-P., Sun, Y., Hou, Q., Wei, W., Zhou, X., Wang, X., Shi, Z., & Furukawa, Y. (2025). Microscopic Characterization of Pb10−xCux(PO4)6O by 31P and 63/65Cu NMR Measurements. Journal of Composites Science, 9(7), 377. https://doi.org/10.3390/jcs9070377

Article Metrics

Back to TopTop