Skip to main content
National Science Review logoLink to National Science Review
. 2023 Feb 14;10(5):nwad034. doi: 10.1093/nsr/nwad034

Pressure-induced superconductivity at 32 K in MoB2

Cuiying Pei 1,d, Jianfeng Zhang 2,d, Qi Wang 3,4,5,d, Yi Zhao 6, Lingling Gao 7, Chunsheng Gong 8, Shangjie Tian 9, Ruitao Luo 10, Mingtao Li 11, Wenge Yang 12, Zhong-Yi Lu 13, Hechang Lei 14,, Kai Liu 15,, Yanpeng Qi 16,17,18,
PMCID: PMC10228782  PMID: 37260928

ABSTRACT

Since the discovery of superconductivity in MgB2 (Tc ∼ 39 K), the search for superconductivity in related materials with similar structures or ingredients has never stopped. Although about 100 binary borides have been explored, only a few of them show superconductivity with relatively low Tc. In this work, we report the discovery of superconductivity up to 32 K, which is the highest Tc in transition-metal borides, in MoB2 under pressure. The Tc of MoB2 in the α phase can be well explained by theoretical calculations in the framework of electron-phonon coupling. Furthermore, the coupling between the d electrons of Mo and the out-of-plane Mo-phonon modes are the main driving force of the 32 K superconductivity of MoB2. Our study sheds light on the exploration of high-Tc superconductors in transition metal borides.

Keywords: superconductivity, high pressure, borides


Discover the superconductivity up to 32 K in MoB2 under pressure and highlight roles of transitional metals in the high-Tc superconductors.

INTRODUCTION

Superconductors with high transition temperature (high Tc) are long-sought targets in the condensed matter physics and materials science communities. Materials with light elements [1–5] are especially favorable as they can provide high Debye frequency, which is proportional to the superconducting Tc according to Bardeen-Cooper-Schrieffer (BCS) theory [6]. In various light-element materials, metal borides have attracted much attention due to their unique crystal structure and rich physical phenomena. The discovery of superconductivity in MgB2 at 39 K in 2001 has reinforced the scientific importance of metal diborides (MB2) [1]. Nevertheless, after nearly two decades of exploration of superconductivity in the rich family of metal diborides, only a few MB2 show superconducting behavior. Furthermore, the Tc of these superconducting MB2 compounds are much lower than that of MgB2 (Table S1 within the online supplementary material).

Molybdenum diboride (MoB2) is unique among the MB2 family since it is the only material that has two structural forms: the α-MoB2 phase (AlB2 type, space group P6/mmm) [7–9] and β-MoB2 phase (CaSi2 type, space group Inline graphic) [10–12]. Although both phases of MoB2 have similar triangular Mo layers, their B atom arrangements are quite different. The B atoms in α-MoB2 constitute AA-staking two-dimensional (2D) graphitic boron layers. In contrast, there are two different kinds of B layers in β-MoB2 (Fig. 1(a)): one forms a nearly planar quasi-2D honeycomb lattice similar to those in α-MoB2 and MgB2; the other builds a buckled honeycomb network. As a result, the Mo atoms in β-MoB2 locate above the centers of hexatomic boron rings in the planar boron layers, but occupy the top sites right above B atoms in the buckled boron layers (Fig. 1(a)). MoB2 may thus provide a good material platform to make a comparative study with MgB2.

Figure 1.

Figure 1.

Crystal structure of MoB2 and its evolution with pressure. (a) Crystal structures of α-MoB2 (space group P6/mmm) and β-MoB2 (space group Inline graphic). (b) XRD patterns of MoB2 measured at room temperature with an increase of external pressure up to 90 GPa. The XRD wavelength λ is 0.6199 Å. Red and black curves are used to distinguish the structure transformation at 65.2 GPa.

Pressure, as a conventional thermodynamic parameter, is a clean and powerful tool to tune the electronic properties of materials. It is possible to trigger structural and electronic transitions, subsequently inducing novel quantum phenomena [13]. Here we report on the structure and transport properties of MoB2 under various pressures. We find that MoB2 exhibits superconductivity under high pressure, reaching a Tc of 32 K around 100 GPa, which is the second highest Tc among all known boride superconductors. Synchrotron X-ray diffraction (XRD) measurements indicate that β-MoB2 transforms to α-MoB2 around 65 GPa. Although the compressed MoB2 has the same structure and comparable Tc with MgB2, the superconducting mechanism of the former is completely different from that of the latter. Theoretical calculation suggests that the d electrons and phonon modes of transition metal Mo atoms play utterly different roles in the emergence of superconductivity in contrast to the dominance of p electrons and phonon modes of B atoms in the superconductivity of MgB2.

RESULTS AND DISCUSSION

We performed in situ XRD measurements on the structural evolution of MoB2 under various pressures. As shown in Fig. 1(b), all the diffraction peaks in the low-pressure range can be indexed well to a rhombohedral primitive cell of β-MoB2 (space group Inline graphic), and both the a-axis and c-axis lattice constants decrease with increasing pressure (Figs. S2 and S3 and Table S2 within the online supplementary material). The structure of β-MoB2 is robust until 65 GPa. Beyond ∼65 GPa, additional diffraction peaks emerge, indicating the occurrence of a structural phase transition. Meanwhile, we carried out a global minimization of the enthalpy of MoB2 under high pressure by combining ab initio total energy calculations and the Calypso technique on structure predictions [14]. At 90 GPa, we theoretically predict two structural forms (space groups P6/mmm and I41/amd) of MoB2 with lower enthalpies than the Inline graphic structure (Fig. S4a within the online supplementary material). No imaginary frequency is found in the phonon dispersions of these two predicted structures, suggesting their dynamical stability (Fig. S4c and S4d within the online supplementary material). We find that the XRD pattern at 90 GPa can be well refined by using the hexagonal α-MoB2 structure (space group P6/mmm; Figs S4b and S5 within the online supplementary material). These experimental and theoretical results suggest that, under high pressure, there is a structural phase transition from β-MoB2 to α-MoB2 with a critical pressure Pc ∼ 70 GPa. With the increasing pressure, the content of α-MoB2 increases and almost reaches 100% at 90 GPa (Fig. S2a within the online supplementary material).

Since MoB2 under high pressure possesses the same crystal structure as MgB2, a question arises naturally: is it possible to achieve superconductivity in MoB2 under high pressure? Hence, we measured the electrical resistivity ρ(T) of a β-MoB2 single crystal at various pressures. Figure 2(a) shows the typical ρ(T) curves for pressure up to 109.7 GPa. The ρ(T) curves display a metallic-like behavior in the whole pressure range. When the pressure increases to 21.7 GPa, a small drop of ρ is observed at the lowest measuring temperature (Tmin = 1.8 K), as shown in Fig. 2(b). With further increasing pressure, zero resistivity is achieved at low temperature for P > 38.5 GPa, indicating the emergence of superconductivity. The superconducting Tc increases dramatically with pressure. Beyond the critical pressure (Pc = 70 GPa), where the structural phase transition happens, the growth of Tc slows down and the maximum Tc of 32.4 K is attained at P = 109.7 GPa, which is the highest pressure we can exert on the sample. At this pressure, the superconducting transition width ΔTc [Inline graphic] is about 4.2 K (Fig. 2(b)), which is rather sharp when compared with the large value of Tc. Figure 2(c) demonstrates that at 109.7 GPa the resistivity drop is continuously suppressed with increasing magnetic field and it shifts to about 5 K at 9 T. Such behavior further confirms that the sharp decrease of ρ(T) should originate from a superconducting transition. The derived upper critical field μ0Hc2(T) as a function of temperature T can be fitted well using the empirical Ginzburg-Landau formula (Fig. 2(d)) μ0Hc2(T) = μ0Hc2(0)(1 − t2)/(1 + t2), where t = T/Tc is the reduced temperature with zero-field superconducting Tc. The fitted zero-temperature upper critical field μ0Hc2(0) of MoB2 from the 90% ρn criterion can reach 9.4(1) T with Tc = 31.7(2) K, which is consistent with the measured value of 32.4 K.

Figure 2.

Figure 2.

Transport properties of MoB2 as functions of pressure and magnetic fields in run IV. (a) Electrical resistivity ρ(T) of MoB2 as a function of temperature at different pressures. (b) Enlarged ρ(T) curves in the vicinity of the superconducting transition. (c) Electrical resistivity ρ(T) under various magnetic fields at 109.7 GPa. (d) Temperature dependence of the upper critical field μ0Hc2(T) at 109.7 GPa. Here the Tc are determined at 10%, 50% and 90% of the normal state resistivity just above the onset superconducting transition temperature. The solid lines represent the fits using the Ginzburg-Landau formula.

The measurements on different samples of MoB2 for five independent runs provide consistent and reproducible results (Figs S6 and S7 within the online supplementary material), confirming this intrinsic superconductivity under pressure. The superconducting phase diagram of MoB2 as a function of the pressure is summarized in Fig. 3. It can be seen that the superconducting state emerges around 20 GPa, and then the Tc increases further with applied pressure. The Tc raises dramatically at a rate of 0.7 K/GPa in the range of 40 to 70 GPa, and beyond the structure-transition pressure (Pc ∼ 70 GPa) the growth of Tc slows down (0.1 K/GPa). The Tc of MoB2 rises to as high as ∼32 K at a pressure of 109.7 GPa and still does not exhibit the trend of saturation.

Figure 3.

Figure 3.

Phase diagram of MoB2. Pressure dependence of the superconducting transition temperature Tc for MoB2 up to 109.7 GPa in different runs. The values of Tc are determined from the high-pressure resistivity (90% ρn criterion).

To explore the origin of the relatively high-Tc superconductivity in MoB2 under pressure, we performed density functional theory (DFT) [15,16] and density functional perturbation theory (DFPT) [17,18] calculations on MoB2 in the α phase at 90 GPa, for which the experimental Tc can reach ∼31.5 K at 91.4 GPa. The calculated band structure of α-MoB2 at 90 GPa along the high-symmetry paths of the Brillouin zone is shown in Fig. 4(a). There are several bands crossing the Fermi level (EF), indicating its metallic character. Based on the analysis of the total and partial density of states (Fig. 4(b)), Mo d orbitals (especially the dz2 orbital) have larger contributions than B p orbitals around EF. The phonon dispersion and the phonon density of states F(ω) for α-MoB2 at 90 GPa are displayed in Figs. 4(c) and 4(d), respectively. Because of the large mismatch between the atomic masses of Mo and B, there is a gap between the low-frequency acoustic branch contributed by Mo atoms and the high-frequency optical branch contributed by B atoms. Obviously, the acoustic modes along the A-L path and around the L and H points make great contributions to the electron-phonon coupling (EPC, represented by the red dots in Fig. 4(c)), which is also manifested in the Eliashberg spectral function α2F(ω) (red line in Fig. 4(d)). Figure 4(e) demonstrates a representative phonon mode of α-MoB2 at the H point at 90 GPa. For clarity, a Inline graphic supercell is employed and the arrows on the B atoms are magnified two times. It can be seen that this phonon mode consists of the relative vibrations of Mo atoms perpendicular to the B-B honeycomb plane and the related in-plane breath-like vibrations of B atoms. The calculated total EPC constant λ is 1.60. Based on the McMillan-Allen-Dynes formula [19,20], we can then calculate the superconducting Tc with an effective screened Coulomb repulsion constant μ*. By adopting several values of μ* in a commonly used empirical range of 0.08 to 0.15 [19,20], we obtained Tc between 27.2 and 33.3 K (see Fig. S8 within the online supplementary material), showing good order of magnitude with the measured one (∼32 K). These theoretical calculations confirm that the observed superconductivity in MoB2 at high pressure should belong to the conventional BCS type.

Figure 4.

Figure 4.

Calculated electronic structure and atomic displacements for the typical phonon modes of α-MoB2 under 90 GPa. (a) Electronic band structure. (b) Density of states N(E). (c) Phonon dispersion. The sizes of the red dots schematically denote the electron-phonon coupling strength λqv. (d) Phonon density of states F(ω) (black line) and Eliashberg spectral function α2F(ω) (red line). (e) Atomic displacements of the lowest acoustic mode at the H point (Fig. 4(c)), which has a large EPC. A Inline graphic supercell is employed. The arrows on Mo/B atoms indicate vibrational directions and magnitudes, where the arrows on the B atoms are magnified two times for clarity. (f) The in-plane B-B stretching mode for MgB2 at ambient pressure. The Mo, Mg and B atoms are represented by purple, red and yellow balls, respectively.

Although α-MoB2 at high pressure shares the same crystal structure as MgB2 and owns a comparable superconducting Tc with that of MgB2 at ambient pressure [1], their electronic structures and superconducting ingredients are distinctly different. Firstly, for MgB2, the p orbitals of B atoms in the B-B honeycomb lattice contribute most around the Fermi level [21,22], while for α-MoB2 under high pressure, the states near EF are dominated by the d electrons of Mo atoms (Fig. 4(b)) [23]. As a result, MoB2 has very large band dispersion along the kz direction (such as the Γ-A path; Fig. 4(a)) and its Fermi surface shows a three-dimensional characteristic, i.e. without two-dimensional Fermi surfaces as in MgB2 [22,24] (Fig. S9 within the online supplementary material). Secondly, the main contribution to the EPC of α-MoB2 under high pressure derives from the low-frequency phonon branch (Fig. 4(c) and 4(d)); instead, the high-frequency branch plays a major role in MgB2 [25–28]. To be specific, for high-pressure α-MoB2, the out-of-plane phonon mode of Mo atoms couples strongly with Mo d electrons near EF (Fig. 4(b) and 4(e)). In comparison, it is the in-plane B-B stretching mode in MgB2 that interacts intensively with the σ bond in the boron honeycomb lattice around EF (Fig. 4(f)) [21]. Last but not least, previous theoretical studies suggest that the anisotropy in the EPC and the anharmonicity in the stretching phonon mode are very crucial for the high Tc of MgB2 [27–29]. Here, for α-MoB2 under high pressure, our calculations seem to coincide with the observed Tc without invoking the above two factors. These results reveal that the superconducting mechanism in high-pressure α-MoB2 is distinct from that in MgB2, suggesting the possibility of exploring new phonon-mediated high-Tc superconductors in transition metal borides.

In addition to the α phase of MoB2, we have also calculated the superconducting Tc of β-MoB2 using the same method. Nevertheless, the calculated Tc of β-MoB2 is always lower than 5 K and shows a decreasing tendency with pressure (see Fig. S10 within the online supplementary material), which are inconsistent with our experimental observations (Fig. 3). The small calculated Tc of β-MoB2 is related to its low electronic density of states N(EF) at the Fermi level (Fig. S10), which originates from the strong bonding between the Mo atoms and the B atoms in buckled boron layers. In the real β-MoB2 sample, the Mo atoms around the stacking fault or grain boundary can easily slide under pressure, which may drive some buckled boron layers to flat ones. As a result, the N(EF) could increase along with more planar boron layers (Fig. S10), hence facilitating the enhanced superconductivity. On the other hand, the pressure evolution of the strength of EPC in the β-MoB2 sample can be extracted from the fitting of the Raman spectrum by the Fano function [30–32], as shown in Fig. S11 within the online supplementary material. The Raman response to EPC manifests an asymmetric profile of the spectral line shape of Raman-active phonons. Explicitly, the inverse of the Fano asymmetric parameter |1/q| is proportional to the joint density of states, hence counting the strength of EPC. The representative fitting at 13.0 GPa is plotted in the inset of Fig. S11d. Combined with our first-principles calculations, we find that there are two Raman-active phonon modes that can provide strong EPC: one is the double-degenerated Eg mode around 200 cm−1 and the other is a Ag mode around 300 cm−1. Similar to the case in α-MoB2, both these two strong EPC modes come from the relative vibrations of Mo atoms, as shown in Fig. S11c. From the pressure dependent |1/q|, we note that the one from the Eg mode is nearly independent of pressure. However, the other one extracted from the Ag mode increases dramatically with increasing pressure, which presents compelling evidence of the enhanced EPC in MoB2 induced by increasing pressure. In addition, the Ag mode involves the vertical vibrations of Mo atoms relative to the B-B plane, which is similar to the corresponding phonon mode in α-MoB2 at high pressure (Fig. 4(e)), suggesting that more Mo atoms get rid of bonding with the B atoms in buckled layers with increasing pressure. The increased frequency and the enhanced EPC of this Ag mode as well as the boosted N(EF) mentioned above probably have a close connection with the increase of Tc in β-MoB2 in the phonon-mediated superconductivity regime.

CONCLUSION

In summary, we found superconductivity up to 32 K in MoB2 under pressure. Although compressed α-MoB2 and MgB2 both possess the AlB2-type structures and comparable Tc, the features of their electron-phonon couplings that mediate the superconductivity are quite different. In both MgB2 and α-MoB2, the active phonon modes directly modulate the active electronic bands, causing the strong electron-phonon coupling [33]. However, it is the out-of-plane Mo phonon that directly modulates the d electrons of Mo in α-MoB2, in comparison with the stretching B-phonon mode modulating the p orbitals of B in MgB2. Thus, not only do the electrons in light-element atom networks need to be considered, but the relatively localized electrons of metal atoms may also be worthy of attention in the search for high-Tc superconductors.

MATERIALS AND METHODS

High-quality single crystals of MoB2 were grown by Al flux (see the online supplementary material). In situ high-pressure XRD and resistivity measurements were performed in a diamond anvil cell. We employed the swarm-intelligence-based CALYPSO structure prediction method [14] to find the energetically stable structures of MoB2 under high pressure. The electronic structure, phonon spectrum and EPC of AlB2-type α-MoB2 at 90 GPa were studied based on DFT [15,16] and DFPT [17,18] calculations. The details of the experiment are given in the online supplementary material.

Note added

After we submitted this paper, we learned that pressure-induced superconductivity was observed in WB2 by another group [34]. Most of the data in that paper are consistent with our results [35]. We also observed superconductivity in itinerant antiferromagnet CrB2 via the application of external pressure [36].

Supplementary Material

nwad034_Supplemental_File

Contributor Information

Cuiying Pei, School of Physical Science and Technology, ShanghaiTech University, Shanghai 201210, China.

Jianfeng Zhang, Department of Physics and Beijing Key Laboratory of Opto-electronic Functional Materials & Micro-nano Devices, Renmin University of China, Beijing 100872, China.

Qi Wang, School of Physical Science and Technology, ShanghaiTech University, Shanghai 201210, China; Department of Physics and Beijing Key Laboratory of Opto-electronic Functional Materials & Micro-nano Devices, Renmin University of China, Beijing 100872, China; ShanghaiTech Laboratory for Topological Physics, ShanghaiTech University, Shanghai 201210, China.

Yi Zhao, School of Physical Science and Technology, ShanghaiTech University, Shanghai 201210, China.

Lingling Gao, School of Physical Science and Technology, ShanghaiTech University, Shanghai 201210, China.

Chunsheng Gong, Department of Physics and Beijing Key Laboratory of Opto-electronic Functional Materials & Micro-nano Devices, Renmin University of China, Beijing 100872, China.

Shangjie Tian, Department of Physics and Beijing Key Laboratory of Opto-electronic Functional Materials & Micro-nano Devices, Renmin University of China, Beijing 100872, China.

Ruitao Luo, Department of Physics and Beijing Key Laboratory of Opto-electronic Functional Materials & Micro-nano Devices, Renmin University of China, Beijing 100872, China.

Mingtao Li, Center for High Pressure Science and Technology Advanced Research, Shanghai 201203, China.

Wenge Yang, Center for High Pressure Science and Technology Advanced Research, Shanghai 201203, China.

Zhong-Yi Lu, Department of Physics and Beijing Key Laboratory of Opto-electronic Functional Materials & Micro-nano Devices, Renmin University of China, Beijing 100872, China.

Hechang Lei, Department of Physics and Beijing Key Laboratory of Opto-electronic Functional Materials & Micro-nano Devices, Renmin University of China, Beijing 100872, China.

Kai Liu, Department of Physics and Beijing Key Laboratory of Opto-electronic Functional Materials & Micro-nano Devices, Renmin University of China, Beijing 100872, China.

Yanpeng Qi, School of Physical Science and Technology, ShanghaiTech University, Shanghai 201210, China; ShanghaiTech Laboratory for Topological Physics, ShanghaiTech University, Shanghai 201210, China; Shanghai Key Laboratory of High-resolution Electron Microscopy, ShanghaiTech University, Shanghai 201210, China.

ACKNOWLEDGEMENTS

Computational resources were provided by the Physical Laboratory of High Performance Computing at Renmin University of China. The authors acknowledge support from the Analytical Instrumentation Center (♯ SPST-AIC10112914), SPST, ShanghaiTech University. The authors thank BL15U1 staff at Shanghai Synchrotron Radiation Facility for assistance during the data collection.

FUNDING

This work was supported by the National Key R&D Program of China (2018YFA0704300, 2018YFE0202600 and 2017YFA0302903), the National Natural Science Foundation of China (52272265, U1932217, 11774424, 11974246, 12004252, 12174443, 12274459 and 11804011), the Beijing Natural Science Foundation (Z200005), the CAS Interdisciplinary Innovation Team, the Fundamental Research Funds for the Central Universities, and the Research Funds of Renmin University of China (18XNLG14, 19XNLG13 and 19XNLG17).

AUTHOR CONTRIBUTIONS

Y.-P.Q., H.-C.L. and K.L. proposed and supervised the project. Q.W., C.-S.G., S.-J.T., R.-T.L. and H.-C.L. carried out the synthesis. C.-Y.P., Y.Z., L.-L.G., M.-T.L., W.-G.Y. and Y.-P.Q. performed the high-pressure experiments. J.-F.Z., Z.-Y.L. and K.L. carried out calculations. All authors discussed the results and participated in analyzing the experimental results.

Conflict of interest statement. None declared.

REFERENCES

  • 1. Nagamatsu J, Nakagawa N, Muranaka Tet al. Superconductivity at 39 K in magnesium diboride. Nature 2001; 410: 63–4. 10.1038/35065039 [DOI] [PubMed] [Google Scholar]
  • 2. Tobin JR. Superconductivity Research Developments. New York: Nova Science Publishers, 2008. [Google Scholar]
  • 3. Drozdov AP, Eremets MI, Troyan IAet al. Conventional superconductivity at 203 kelvin at high pressures in the sulfur hydride system. Nature 2015; 525: 73–6. 10.1038/nature14964 [DOI] [PubMed] [Google Scholar]
  • 4. Drozdov AP, Kong PP, Minkov VSet al. Superconductivity at 250 K in lanthanum hydride under high pressures. Nature 2019; 569: 528–31. 10.1038/s41586-019-1201-8 [DOI] [PubMed] [Google Scholar]
  • 5. Snider E, Dasenbrock-Gammon N, McBride Ret al. Room-temperature superconductivity in a carbonaceous sulfur hydride. Nature 2020; 586: 373–7. 10.1038/s41586-020-2801-z [DOI] [PubMed] [Google Scholar]
  • 6. Bardeen J, Cooper LN, Schrieffer JR. Theory of superconductivity. Phys Rev 1957; 108: 1175–204. 10.1103/PhysRev.108.1175 [DOI] [Google Scholar]
  • 7. Kudaka K, Iizumi K, Sasaki Tet al. Mechanochemical synthesis of MoB2 and Mo2B5. J Alloys Compd 2001; 315: 104–7. 10.1016/S0925-8388(00)01082-3 [DOI] [Google Scholar]
  • 8. Steinitz R, Binder I, Moskowitz D. System molybdenum-boron and some properties of the molybdenum-borides. JOM 1952; 4: 983–7. 10.1007/BF03397758 [DOI] [Google Scholar]
  • 9. Ding LP, Shao P, Zhang FHet al. Crystal structures, stabilities, electronic properties, and hardness of MoB2: first-principles calculations. Inorg Chem 2016; 55: 7033–40. 10.1021/acs.inorgchem.6b00899 [DOI] [PubMed] [Google Scholar]
  • 10. Okada S, Atoda T, Higashi Iet al. Preparation of single crystals of MoB2 by the aluminium-flux technique and some of their properties. J Mater Sci 1987; 22: 2993–9. 10.1007/BF01086503 [DOI] [Google Scholar]
  • 11. Frotscher M, Klein W, Bauer Jet al. M2B5 or M2B4? A reinvestigation of the Mo/B and W/B system. Z Anorg Allg Chem 2007; 633: 2626–30. 10.1002/zaac.200700376 [DOI] [Google Scholar]
  • 12. Zhang TT, Miao H, Wang Qet al. Phononic helical nodal lines with PT protection in MoB2. Phys Rev Lett 2019; 123: 245302. 10.1103/PhysRevLett.123.245302 [DOI] [PubMed] [Google Scholar]
  • 13. Qi Y, Naumov PG, Ali MNet al. Superconductivity in Weyl semimetal candidate MoTe2. Nat Commun 2016; 7: 11038. 10.1038/ncomms11038 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14. Wang Y, Lv J, Zhu Let al. CALYPSO: a method for crystal structure prediction. Comput Phys Commun 2012; 183: 2063–70. 10.1016/j.cpc.2012.05.008 [DOI] [Google Scholar]
  • 15. Hohenberg P, Kohn W. Inhomogeneous electron gas. Phys Rev 1964; 136: B864–71. 10.1103/PhysRev.136.B864 [DOI] [Google Scholar]
  • 16. Kohn W, Sham LJ. Self-consistent equations including exchange and correlation effects. Phys Rev 1965; 140: A1133–8. 10.1103/PhysRev.140.A1133 [DOI] [Google Scholar]
  • 17. Giustino F. Electron-phonon interactions from first principles. Rev Mod Phys 2017; 89: 015003. 10.1103/RevModPhys.89.015003 [DOI] [Google Scholar]
  • 18. Baroni S, de Gironcoli S, Dal Corso Aet al. Phonons and related crystal properties from density-functional perturbation theory. Rev Mod Phys 2001; 73: 515–62. 10.1103/RevModPhys.73.515 [DOI] [Google Scholar]
  • 19. Lee KH, Chang KJ, Cohen ML. First-principles calculations of the Coulomb pseudopotential μ*: application to Al. Phys Rev B 1995; 52: 1425–8. 10.1103/PhysRevB.52.1425 [DOI] [PubMed] [Google Scholar]
  • 20. Richardson CF, Ashcroft NW. High temperature superconductivity in metallic hydrogen: electron-electron enhancements. Phys Rev Lett 1997; 78: 118–21. 10.1103/PhysRevLett.78.118 [DOI] [Google Scholar]
  • 21. An JM, Pickett WE. Superconductivity of MgB2: covalent bonds driven metallic. Phys Rev Lett 2001; 86: 4366–9. 10.1103/PhysRevLett.86.4366 [DOI] [PubMed] [Google Scholar]
  • 22. Kortus J Mazin II, Belashchenko KDet al. Superconductivity of metallic boron in MgB2. Phys Rev Lett 2001; 86: 4656–9. 10.1103/PhysRevLett.86.4656 [DOI] [PubMed] [Google Scholar]
  • 23. Quan Y, Lee KW, Pickett WE. MoB2 under pressure: superconducting Mo enhanced by boron. Phys Rev B 2021; 104: 224504. 10.1103/PhysRevB.104.224504 [DOI] [Google Scholar]
  • 24. Choi HJ, Roundy D, Sun Het al. The origin of the anomalous superconducting properties of MgB2. Nature 2002; 418: 758–60. 10.1038/nature00898 [DOI] [PubMed] [Google Scholar]
  • 25. Bohnen KP, Heid R, Renker B. Phonon dispersion and electron-phonon coupling in MgB2 and AlB2. Phys Rev Lett 2001; 86: 5771–4. 10.1103/PhysRevLett.86.5771 [DOI] [PubMed] [Google Scholar]
  • 26. Kong Y, Dolgov OV, Jepsen Oet al. Electron-phonon interaction in the normal and superconducting states of MgB2. Phys Rev B 2001; 64: 020501. 10.1103/PhysRevB.64.020501 [DOI] [Google Scholar]
  • 27. Liu AY Mazin II, Kortus J. Beyond Eliashberg superconductivity in MgB2: anharmonicity, two-phonon scattering, and multiple gaps. Phys Rev Lett 2001; 87: 087005. 10.1103/PhysRevLett.87.087005 [DOI] [PubMed] [Google Scholar]
  • 28. Yildirim T, Gulseren O, Lynn JWet al. Giant anharmonicity and nonlinear electron-phonon coupling in MgB2: a combined first-principles calculation and neutron scattering study. Phys Rev Lett 2001; 87: 037001. 10.1103/PhysRevLett.87.037001 [DOI] [PubMed] [Google Scholar]
  • 29. Choi HJ, Roundy D, Sun Het al. First-principles calculation of the superconducting transition in MgB2 within the anisotropic Eliashberg formalism. Phys Rev B 2002; 66: 020513. 10.1103/PhysRevB.66.020513 [DOI] [Google Scholar]
  • 30. Cardona M, Abstreiter G, Pinczuk A. Light Scattering in Solids IV. Berlin: Springer, 1984. [Google Scholar]
  • 31. Hadjiev VG, Zhou X, Strohm Tet al. Strong superconductivity-induced phonon self-energy effects in HgBa2Ca3Cu4O10 + δ. Phys Rev B 1998; 58: 1043–50. 10.1103/PhysRevB.58.1043 [DOI] [Google Scholar]
  • 32. Friedl B, Thomsen C, Cardona M. Determination of the superconducting gap in YBa2Cu3O7-δ. Phys Rev Lett 1990; 65: 915–8. 10.1103/PhysRevB.62.146 [DOI] [PubMed] [Google Scholar]
  • 33. Gao M, Lu ZY, Xiang T. Prediction of phonon-mediated high-temperature superconductivity in Li3B4C2. Phys Rev B 2015; 91: 045132. 10.1103/PhysRevB.91.045132 [DOI] [Google Scholar]
  • 34. Lim J, Hire AC, Quan Yet al. Creating superconductivity in WB2 through pressure-induced metastable planar defects. Nat Commun 2022; 13: 7901. 10.1103/s41467-022-351 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35. Pei C, Zhang J, Gong Cet al. Distinct superconducting behaviors of pressurized WB2 and ReB2 with different local B layers. Sci China Phys Mech Astron 2022; 65: 287412. 10.1007/s11433-022-1911-x [DOI] [Google Scholar]
  • 36. Pei C, Yang P, Gong Cet al. Pressure-induced superconductivity in itinerant antiferromagnet CrB2. arXiv: 2109.15213.

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

nwad034_Supplemental_File

Articles from National Science Review are provided here courtesy of Oxford University Press

RESOURCES