The electronic band structure and features of the charge density distribution in lead oxide phosphate Pb10(PO4)6O and the copper-doped compound Pb9Cu(PO4)6O have been studied by the density functional theory. Despite the differences in chemical compositions and crystal structures, the type of chemical bonding in the Pb10(PO4)6O and Pb9Cu(PO4)6O compounds was found to be similar to the type of chemical bonding that we previously revealed in high-temperature superconductors and in parent compounds for their production—monoclinic α-Bi2O3 and orthorhombic La2CuO4. Although the lack of experimental data on the electronic band structure and physical properties of the Pb10(PO4)6O and Pb10−xCux(PO4)6O compounds does not currently allow us to conclude that superconductivity could exist at room temperature and atmospheric pressure in the Pb10−xCux(PO4)6O compound, further studies of the properties of lead oxide phosphate and other minerals of the apatite supergroup might be useful for identifying new types of promising materials for the production of high-temperature superconductors.
I. INTRODUCTION
Recently, in three articles,1–3 it has been claimed that the compound Pb10−xCux(PO4)6O (0.9 < x < 1.1) is a superconductor at room temperature and atmospheric pressure. The parent compound for the above-mentioned superconductor is lead oxide phosphate,4 Pb10(PO4)6O. These publications were received with enthusiasm by a wide audience due to their potential applications. On the contrary, experts met them with a degree of skepticism due to the lack of convincing experimental results presented.5 As follows from the articles,1–3 the replacement of one Pb atom located in the 4f position in the crystal structure described by the space group P63/m6 with a Cu atom leads to the insulator–metal transition and the appearance of superconductivity at room temperature and atmospheric pressure.
Shortly after the publications of Refs. 1–3, a number of reports appeared on the arXiv site cond-mat.supr-con; the authors, in most cases, could not reproduce the results of Korean scientists (see, for example, Ref. 7). The only paper8 known to us so far reported the discovery of superconductivity at 110 K in a Pb10−xCux(PO4)6O sample. The first theoretical studies using the density functional theory (DFT) found flat bands and high density of electronic states (DOS) on the Fermi surface in the Pb9Cu(PO4)6O compound (see, for example, Refs. 9 and 10), which was regarded by the authors as evidence of the possibility of high-temperature superconductivity (HTSC) in compounds with a similar electronic band structure.
Due to the great attention to the potential perspectives of the Pb10−xCux(PO4)6O compound in the field of HTSC, it was interesting to compare the type of chemical bonding in the parent Pb10(PO4)6O and copper-doped Pb9Cu(PO4)6O compounds with the type of chemical bonding that had been previously found by us, both in monoclinic α-Bi2O3 and orthorhombic La2CuO4—the parent compounds for obtaining cuprate HTSCs—and in HTSCs themselves.11,12
II. METHODS OF CALCULATIONS
To achieve this goal, we applied the approach that we had used earlier to calculate the electronic band structure and analyze the features of the charge density distribution in bismuth and antimony chalcogenides, iron pnictides, cuprates, and cuprate-based HTSCs,11–14 as well as in trigonal Se and Te15 and in molecular crystals of chalcogens.16 The WIEN2k program17,18 based on the “augmented plane waves plus local orbitals” method of DFT was used in this approach for calculations of the electronic band structure and for finding the charge density distribution in crystal lattices. The main results of this article were obtained using the Perdew–Burke–Ernzerhof (PBE)19 exchange-correlation functional, which is a variant of the Generalized Gradient Approximation (GGA). For comparison, the calculations were also performed using the Becke and Johnson exchange potential modified by Tran and Blaha20 (mBJ) and the correlations in the Local Density Approximation (LDA).21
For valence states, relativistic effects have been taken into account using the second variational method including spin–orbital coupling. The core electron states were calculated fully relativistically.17 To achieve the required accuracy, the following parameters were used: the expansion in l-orbital momentum inside atomic spheres carried out up to lmax = 10, the largest values of the wave vector kmax used for expansion in plane waves found using the product Rmt · kmax = 5.5, where Rmt is the smallest of all atomic sphere radii in the unit cell. The total number of k-points in the full Brillouin zone was taken to be equal to 800. The convergence criterion of the iterative calculation procedure for the total energy was 10−4 Ry, and for the charge, the difference between the charge density integrated over the unit cell for two successive iterations of the self-consistency cycle was taken to be equal to 10−4 e (e is the positive value of the elementary charge).
The charge density distribution ρ(r) in crystal lattices of the Pb10(PO4)6O and Pb9Cu(PO4)6O compounds was taken from the results of calculations of the electronic band structure. To study the peculiarities of ρ(r), we used the program CRITIC2,22 developed for the topological analysis of scalar fields in periodic structures based on the “Quantum Theory of Atoms in Molecules and Crystals” (QTAIM) method.23–25 According to Bader’s QTAIM, the points rc, at which the charge density gradient is equal to zero ∇ρ(rc) = 0, are called critical points. The types of critical points in the charge density distribution are characterized by the rank ω [the number of nonzero eigenvalues λi—the main values of curvature of the Hesse matrix of the second derivatives of the charge density over coordinates (∂2ρ/∂xi∂xj)] and signature ϭ—the algebraic sum of λi signs: (ω, ϭ). There are four types of stable critical points:23–25 (3, −3) – nucleus (local maximum), (3, +3) – cage (local minimum), (3, +1) – ring (the first saddle critical point), and (3, −1) – bond (the second saddle critical point). Critical points of the bond type [bond critical points (BCPs)] play an essential role in the classification of the chemical bonding types. Covalent bonding is characterized by the following Bader’s parameters for BCPs:23–25 the negative sign of the charge density Laplacian ∇2ρb < 0, two negative values of curvature λ1,2 < 0, which are large in absolute values , and large values of the charge density ρb. While for the ionic bonding, ∇2ρb > 0, , ρb is small, and the charge density is concentrated mainly on atoms.
In our paper,11 we have revealed a correlation between the critical temperature Tc of HTSC and the value of Laplacian ∇2ρb at the BCP with the highest charge density ρb. In addition, it has been shown11,12 that in the parent compounds of cuprate HTSCs, monoclinic α-Bi2O3 and orthorhombic La2CuO4, as well as in HTSCs themselves, there is a specific mechanism of chemical bonding. This mechanism is characterized by positive values of the charge density Laplacian ∇2ρb for all BCPs and by the number of BCPs per atom, Nat, exceeding the number of valence electrons in atoms. This mechanism can cause charge density fluctuations in crystals. Recently, short-range dynamic charge density fluctuations have been detected in YBa2Cu3O7−δ and Nd1+xBa2−xCu3O7−δ using Resonance Inelastic X-ray Scattering (RIXS).26,27 It was noted in Refs. 26 and 27 that these fluctuations have a characteristic energy of several millielectronvolts, arise at temperatures above the temperature T*, at which the pseudogap begins to appear, and are observed in a wide range of hole concentrations p in the electron subsystem on the T-p phase diagram, almost for all cuprate HTSCs.
It is shown below that, despite the differences in chemical compositions and crystal structures, the parameters of BCPs in lead oxide phosphate Pb10(PO4)6O and in the doped compound Pb9Cu(PO4)6O are similar to the parameters of BCPs of HTSCs and their parent compounds.
Recently, one of the authors of the article4 (Krivovichev) has revised28 the previously published data on the crystal structure of the Pb10(PO4)6O compound after x-ray diffraction experiments on a single crystal fabricated by Merker and Wondratschek.29 The main differences between the crystal structure presented in Ref. 28 and the results of the paper4 are as follows: the symmetry of the crystal structure of the Pb10(PO4)6O compound is described by trigonal space group (N 147)6 with a double value of the lattice parameter c and the changed environment of the oxygen atom O4, which is not part of the PO4 tetrahedra. These data are essential for the calculations of the electronic bands’ structure, especially for the question of the presence of flat bands near the Fermi surface of Pb9Cu(PO4)6O, but they are not fundamentally important for the main purpose of this work, which is the classification of the type of chemical bonding in Pb10(PO4)6O and in the doped compound Pb9Cu(PO4)6O. Due to the fact that studies of the crystal structure of Pb10(PO4)6O are still ongoing, we performed calculations with published values of the crystal lattice parameters: a = b = 9.865 Å, c = 7.4306 Å for Pb10(PO4)6O,4 and a = b = 9.843 Å, c = 7.428 Å for Pb9Cu(PO4)6O1–3 without optimizing their values. Our studies of the electronic structure of trigonal Se and Te have shown15 that optimization of crystal lattice parameters does not change the type of chemical bonding in the crystal. For the same reason, we did not optimize the values of the parameters of the positions occupied by atoms in the unit cell of the crystal in order to reduce the forces acting on the atoms. The experimental positions of the atoms and their coordinates in the unit cell of Pb10(PO4)6O4 (176 space group P63/m6) are given in Table I.
Experimental positions of the atoms in the crystal lattice of Pb10(PO4)6O.
Atom . | Wyckoff symbol . | Atom coordinates x, y, z . |
---|---|---|
Pb1 | 6h | 0.000 56, 0.753 51, 1/4 |
Pb2 | 4f | 1/3, 2/3, 0.003 53 |
P | 6h | 0.6274, 0.5981, 1/4 |
O1 | 6h | 0.5134, 0.6639, 1/4 |
O2 | 12i | 0.7358, 0.6528, 0.0835 |
O3 | 6h | 0.5306, 0.4147, 1/4 |
O4 | 4e | 0, 0, 0.1342 |
Atom . | Wyckoff symbol . | Atom coordinates x, y, z . |
---|---|---|
Pb1 | 6h | 0.000 56, 0.753 51, 1/4 |
Pb2 | 4f | 1/3, 2/3, 0.003 53 |
P | 6h | 0.6274, 0.5981, 1/4 |
O1 | 6h | 0.5134, 0.6639, 1/4 |
O2 | 12i | 0.7358, 0.6528, 0.0835 |
O3 | 6h | 0.5306, 0.4147, 1/4 |
O4 | 4e | 0, 0, 0.1342 |
In Table I, the coordinates of one position of each type are given as fractions of the corresponding lattice parameters. Due to the fact that only one of the 4e positions is occupied by the O4 atom and the need to replace one of the four Pb2 atoms with a Cu atom in the Pb9Cu(PO4)6O compound, it turned out that it is more convenient to carry out the calculations using the specification of the positions of another trigonal space group,6, P3 (143), allowed for minerals of the apatite supergroup.30 The space group P3 has four types of positions6 with coordinates: 1a (0, 0, z), 1b (1/3, 2/3, z), 1c (2/3, 1/3, z), and 3d (x, y, z), (−y, x − y, z), (−x + y, −x, z). Obviously, the 1a position can be used for the O4 atom, the six atoms occupying the 6h positions (see Table I) can be placed in two 3d positions, 12 O2 atoms from the 12i position occupy four 3d positions, and 4 Pb2 atoms in the 4f positions can be represented by the two 1b and the two 1c positions.
III. RESULTS AND DISCUSSION
To classify the types of chemical bonding in the Pb10(PO4)6O and Pb9Cu(PO4)6O crystals, we will use Bader’s parameters of BCPs (the eigenvalues λi—main curvature values of the Hesse matrix, sign and magnitude of the charge density Laplacian ∇2ρb, and value of the charge density ρb). In addition, we will also use the dimensionless parameter flatness f (the ratio of the minimum charge density at the cage-type critical points to the maximum charge density at BCPs ), which characterizes the uniformity of the charge density distribution in the crystal. Table II shows BCPs parameters of the charge density distribution in the Pb10(PO4)6O and Pb9Cu(PO4)6O compounds, calculated using the CRITIC2 program,22 based on the results of calculations of the electronic band structure performed using the WIEN2k program17,18 and the exchange-correlation functional PBE.19 In particular, Table II shows the following data: a compound, bonded atoms, a number of BCPs of each type for non-equivalent atoms (N BCPs), interatomic distances d for the atoms involved in bonding, calculated Bader’s characteristics of BCPs (ratio of main eigenvalues of the Hesse matrix |λ1,2|/λ3, sign and magnitude of Laplacian ∇2ρb, charge density at BCPs ρb), total number of BCPs per atom Nat, and flatness f.
Characteristics of BCPs in Pb10(PO4)6O and Pb9Cu(PO4)6O crystals.
Compound . | Bonding atoms . | N BCPs . | d (Å) . | |λ1,2|/λ3 . | ∇2ρb (e/Å5) . | ρb (e/Å3) . | Nat . | f (%) . |
---|---|---|---|---|---|---|---|---|
Pb10(PO4)6O | P–O2 | 2 | 1.545 | 0.26 | 17.05 | 1.45 | NP = 4 | 0.3 |
P–O1 | 1 | 1.554 | 0.26 | 16.35 | 1.41 | NPb1_1 = 7 | ||
P–O3 | 1 | 1.568 | 0.27 | 14.74 | 1.38 | NPb1_2 = 7 | ||
Pb1–O3 | 1 | 2.440 | 0.22 | 3.29 | 0.34 | NPb2 = 9 | ||
Pb1_1–O4 | 1 | 2.582 | 0.22 | 2.29 | 0.28 | NO1 = 4 | ||
Pb2–O1 | 3 | 2.561 | 0.21 | 2.67 | 0.26 | NO2 = 4 | ||
Pb1–O2 | 2 | 2.597 | 0.20 | 2.48 | 0.25 | NO3 = 4 | ||
Pb1–O2 | 2 | 2.641 | 0.20 | 2.31 | 0.22 | NO4 = 6 | ||
Pb2–O3 | 3 | 2.659 | 0.20 | 2.26 | 0.22 | |||
Pb2–O2 | 3 | 2.944 | 0.16 | 1.45 | 0.13 | |||
Pb1–O1 | 1 | 2.953 | 0.17 | 1.41 | 0.13 | |||
Pb1_2–O4 | 1 | 3.752 | 0.14 | 0.24 | 0.03 | |||
Pb9Cu(PO4)6O | P–O2 | 2 | 1.544 | 0.26 | 17.26 | 1.45 | NP = 4 | 0.3 |
P–O1 | 1 | 1.551 | 0.26 | 16.64 | 1.42 | NPb1_1 = 7 | ||
P–O3 | 1 | 1.564 | 0.27 | 15.06 | 1.39 | NPb1_2 = 7 | ||
Pb1–O3 | 1 | 2.435 | 0.22 | 3.25 | 0.34 | NPb2 = 9 | ||
Pb1_1–O4 | 1 | 2.577 | 0.22 | 2.30 | 0.28 | NCu = 9 | ||
Pb2–O1 | 3 | 2.558 | 0.21 | 2.69 | 0.26 | NO1 = 4 | ||
Pb1–O2 | 2 | 2.592 | 0.21 | 2.49 | 0.25 | NO2 = 4 | ||
Pb1–O2 | 2 | 2.640 | 0.20 | 2.30 | 0.22 | NO3 = 4 | ||
Pb2–O3 | 3 | 2.656 | 0.20 | 2.28 | 0.22 | NO4 = 6 | ||
Cu–O1 | 3 | 2.558 | 0.17 | 1.94 | 0.15 | |||
Cu–O3 | 3 | 2.656 | 0.17 | 1.55 | 0.13 | |||
Pb2–O2 | 3 | 2.938 | 0.16 | 1.48 | 0.13 | |||
Pb1–O1 | 1 | 2.946 | 0.17 | 1.41 | 0.13 | |||
Cu–O2 | 3 | 2.938 | 0.12 | 0.88 | 0.08 | |||
Pb1_2–O4 | 1 | 3.748 | 0.13 | 0.24 | 0.03 |
Compound . | Bonding atoms . | N BCPs . | d (Å) . | |λ1,2|/λ3 . | ∇2ρb (e/Å5) . | ρb (e/Å3) . | Nat . | f (%) . |
---|---|---|---|---|---|---|---|---|
Pb10(PO4)6O | P–O2 | 2 | 1.545 | 0.26 | 17.05 | 1.45 | NP = 4 | 0.3 |
P–O1 | 1 | 1.554 | 0.26 | 16.35 | 1.41 | NPb1_1 = 7 | ||
P–O3 | 1 | 1.568 | 0.27 | 14.74 | 1.38 | NPb1_2 = 7 | ||
Pb1–O3 | 1 | 2.440 | 0.22 | 3.29 | 0.34 | NPb2 = 9 | ||
Pb1_1–O4 | 1 | 2.582 | 0.22 | 2.29 | 0.28 | NO1 = 4 | ||
Pb2–O1 | 3 | 2.561 | 0.21 | 2.67 | 0.26 | NO2 = 4 | ||
Pb1–O2 | 2 | 2.597 | 0.20 | 2.48 | 0.25 | NO3 = 4 | ||
Pb1–O2 | 2 | 2.641 | 0.20 | 2.31 | 0.22 | NO4 = 6 | ||
Pb2–O3 | 3 | 2.659 | 0.20 | 2.26 | 0.22 | |||
Pb2–O2 | 3 | 2.944 | 0.16 | 1.45 | 0.13 | |||
Pb1–O1 | 1 | 2.953 | 0.17 | 1.41 | 0.13 | |||
Pb1_2–O4 | 1 | 3.752 | 0.14 | 0.24 | 0.03 | |||
Pb9Cu(PO4)6O | P–O2 | 2 | 1.544 | 0.26 | 17.26 | 1.45 | NP = 4 | 0.3 |
P–O1 | 1 | 1.551 | 0.26 | 16.64 | 1.42 | NPb1_1 = 7 | ||
P–O3 | 1 | 1.564 | 0.27 | 15.06 | 1.39 | NPb1_2 = 7 | ||
Pb1–O3 | 1 | 2.435 | 0.22 | 3.25 | 0.34 | NPb2 = 9 | ||
Pb1_1–O4 | 1 | 2.577 | 0.22 | 2.30 | 0.28 | NCu = 9 | ||
Pb2–O1 | 3 | 2.558 | 0.21 | 2.69 | 0.26 | NO1 = 4 | ||
Pb1–O2 | 2 | 2.592 | 0.21 | 2.49 | 0.25 | NO2 = 4 | ||
Pb1–O2 | 2 | 2.640 | 0.20 | 2.30 | 0.22 | NO3 = 4 | ||
Pb2–O3 | 3 | 2.656 | 0.20 | 2.28 | 0.22 | NO4 = 6 | ||
Cu–O1 | 3 | 2.558 | 0.17 | 1.94 | 0.15 | |||
Cu–O3 | 3 | 2.656 | 0.17 | 1.55 | 0.13 | |||
Pb2–O2 | 3 | 2.938 | 0.16 | 1.48 | 0.13 | |||
Pb1–O1 | 1 | 2.946 | 0.17 | 1.41 | 0.13 | |||
Cu–O2 | 3 | 2.938 | 0.12 | 0.88 | 0.08 | |||
Pb1_2–O4 | 1 | 3.748 | 0.13 | 0.24 | 0.03 |
Let us note the most significant features of the characteristics of BCPs in Pb10(PO4)6O and Pb9Cu(PO4)6O crystals presented in Table II. First of all, the type of chemical bonding in both crystals is similar to that found by us11,12 in the parent compounds of cuprate HTSCs, monoclinic α-Bi2O3, and orthorhombic La2CuO4, as well as in HTSCs themselves. Moreover, the values of the charge density Laplacian ∇2ρb and the charge density ρb for the first three BCPs in Pb10(PO4)6O are significantly larger than in α-Bi2O3, La2CuO4, and cuprate HTSCs.11,12 This allows us to consider lead oxide phosphate, Pb10(PO4)6O, as a potential parent compound for obtaining superconductors with high values of the critical superconducting transition temperature Tc. Large positive values of the charge density Laplacian ∇2ρb indicate that the charge density is pushed out of the regions near BCPs.23–25 It means that PO4 tetrahedra create a high pressure in the electronic subsystem of the crystal, having a significant impact on the transport properties of the doped compound. The large numbers of BCPs per atom Nat shown in Table II suggest the presence of charge density fluctuations in the parent and doped compounds. It is also important that the doping of lead oxide phosphate with Cu atoms only slightly changes the values of some BCP parameters without affecting the type of chemical bonding inherent in the parent compound, Pb10(PO4)6O.
The total density of electronic states (DOSs) calculated by us for the Pb10(PO4)6O and Pb9Cu(PO4)6O compounds using the WIEN2k program17,18 and the exchange-correlation functional PBE19 are shown in Figs. 1(a) and 1(b).
The total DOS for the compounds: (a) Pb10(PO4)6O, and (b) Pb9Cu(PO4)6O.
As follows from Fig. 1(a), the bandgap for the Pb10(PO4)6O compound calculated using the PBE exchange-correlation functional is equal to 1 eV. It is well known that PBE functional usually results in underestimated bandgaps. Our calculations using the mBJ20 exchange potential and LDA21 correlations gave a value of 1.8 eV for the bandgap. The doped by Cu Pb9Cu(PO4)6O compound proved to be a metal. The total DOS for the Pb9Cu(PO4)6O compound shown in Fig. 1(b) is similar to that given in the paper.9 In our opinion, flat electron bands and very large DOS values on the Fermi surface in Pb9Cu(PO4)6O can be due to the features of the experimental crystal structure1–3 used in the theoretical calculations. It is quite possible that revised crystallographic data28 of the parent Pb10(PO4)6O compound may change the band structure and DOS near the Fermi surface for both compounds. Due to the lack of experimental data on the magnetic state of Cu atoms in the Pb9Cu(PO4)6O compound, the calculations of its electronic band structure were carried out without taking into account spin polarization.
Basing on the results of this article, it is possible to conclude that the Pb10(PO4)6O compound is a potentially promising parent compound for producing HTSCs. However, a number of experiments are required to elucidate its physical properties and find similarities with the properties of the parent cuprate HTSC compounds—orthorhombic La2CuO4 and monoclinic α-Bi2O3. First of all, accurate x-ray and neutron diffraction experiments must be carried out over a wide temperature range to determine the crystal structure of Pb10(PO4)6O reliably. The electronic band structure of Pb10(PO4)6O should also be studied by measuring the x-ray photoemission spectroscopy (XPS) and ultraviolet photoemission spectroscopy (UPS) spectra, which provide information on valence electrons. Data on unoccupied states in the conduction band can be obtained by experimental studies of optical spectra (electron energy loss function, real part of optical conductivity, and reflection coefficient) that are sensitive to interband transitions. In our recent paper,12 a good agreement was found between the experimental literature data and the results of our calculations of the electronic band structure of orthorhombic antiferromagnetic La2CuO4. RIXS experiments can answer the question about the presence of charge density fluctuations in Pb10(PO4)6O. It is also desirable to measure magnetization M(H) and M(T) for Pb10(PO4)6O at low temperatures T and in weak external magnetic fields H using a SQUID magnetometer. Previously, in the paper,32 it was found that, despite the absence of magnetic atoms in the composition of α-Bi2O3, its magnetization in fields H < 4 kOe at 4.2 K was paramagnetic, anisotropic, and significantly dependent on the magnetic prehistory of the single crystal. Moreover, a linear magnetoelectric effect at 4.2 K was found in Ref. 32—the electric polarization P appeared in α-Bi2O3 due to the application of an external magnetic field H. Both the paramagnetism and the magnetoelectric effect in α-Bi2O3 were explained in Ref. 32 by the presence of holes in the electronic structure of oxygen atoms, which served as paramagnetic centers. The discovery of the paramagnetic properties of the parent compound Pb10(PO4)6O will allow us to understand its electronic structure better.
For the Pb10−xCux(PO4)6O compound, first of all, it is desirable to clarify the type of positions occupied by Cu atoms in the crystal structure during doping of the parent compound and to vary the concentration of Cu atoms in order to find the concentration at which the insulator–metal transition occurs. The magnetic properties of Cu atoms in Pb10−xCux(PO4)6O are also of great interest. In our recent paper,12 spin-polarized calculations of the electronic band structure of orthorhombic La2CuO4 revealed the presence of an antiferromagnetic ground state with a bandgap Eg = 2 eV in the case of using the exchange potential mBJ20 and LDA correlations.21
Some other methods of doping the parent compound Pb10(PO4)6O can also be used. Apparently, the simplest variant of doping consists of adding oxygen atoms to the 4e positions, which are only a quarter filled. It is well known that the compound La2CuO4+δ becomes a superconductor at small values of δ (see Ref. 33 and references therein). Although the valence of lead in Pb10(PO4)6O is currently unknown, some of the Pb atoms can be replaced by divalent Sr or Ba atoms.30
IV. CONCLUSION
The lack of experimental data on the electronic band structure and other physical properties of Pb10(PO4)6O and Pb10−xCux(PO4)6O compounds does not currently allow us to conclude that superconductivity could exist at room temperature and atmospheric pressure in the Pb10−xCux(PO4)6O compound. However, if we assume that the correlation between the critical temperature Tc of the HTSC and the value of Laplacian ∇2ρb at the BCP with the highest charge density ρb we found earlier11 is valid for the Pb10(PO4)6O compound, then we can expect Tc of about 150–160 K for the superconductor Pb9Cu(PO4)6O (see Fig. 2).
Predicted superconducting transition temperature Tc for the Pb9Cu(PO4)6O compound.
Predicted superconducting transition temperature Tc for the Pb9Cu(PO4)6O compound.
Although this temperature range is far from room temperature, nevertheless, it is also of great interest, both from scientific and practical points of view. The study of the physical properties of lead oxide phosphate can contribute to the understanding of the HTSC mechanism. Other minerals in the apatite supergroup30 may also be of interest as parent compounds for the production of HTSC.
ACKNOWLEDGMENTS
This work has been carried out using the computing resources of the federal collective usage center Complex for Simulation and Data Processing for Mega-science Facilities at the NRC “Kurchatov Institute.”34
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
V. G. Orlov: Conceptualization (lead); Writing – original draft (equal); Writing – review & editing (equal). G. S. Sergeev: Conceptualization (supporting); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal).
DATA AVAILABILITY
The data that support the findings of this study are available within the article.