Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2024 Jun 20;121(26):e2401840121. doi: 10.1073/pnas.2401840121

Predicted hot superconductivity in LaSc2H24 under pressure

Xin-Ling He a,b,c,1, Wenbo Zhao a,c,1, Yu Xie a,c, Andreas Hermann d, Russell J Hemley e,f,g, Hanyu Liu a,c,h,2, Yanming Ma a,c,h
PMCID: PMC11214075  PMID: 38900793

Significance

Room-temperature superconductivity has been a long-held dream of condensed matter physics for over a century. Here, we predict a class of thermodynamically stable clathrate hydrides consisting of two previously unreported H24 and H30 cages at megabar pressures. Among these hydrides, LaSc2H24 is calculated to exhibit superconductivity with Tc up to 316 K at 167 GPa even when considering the effects of anharmonicity. Such “hot” superconductivity above room temperature is attributed to an unusually large hydrogen-derived density of states at the Fermi level arising from the unusual combination of H30 and H24 cages of this structure. Further exploration of the La–Sc–H ternary system is expected to facilitate the creation of other high-temperature superconductors.

Keywords: superconductivity, high pressure, hydride

Abstract

The recent theory-driven discovery of a class of clathrate hydrides (e.g., CaH6, YH6, YH9, and LaH10) with superconducting critical temperatures (Tc) well above 200 K has opened the prospects for “hot” superconductivity above room temperature under pressure. Recent efforts focus on the search for superconductors among ternary hydrides that accommodate more diverse material types and configurations compared to binary hydrides. Through extensive computational searches, we report the prediction of a unique class of thermodynamically stable clathrate hydrides structures consisting of two previously unreported H24 and H30 hydrogen clathrate cages at megabar pressures. Among these phases, LaSc2H24 shows potential hot superconductivity at the thermodynamically stable pressure range of 167 to 300 GPa, with calculated Tcs up to 331 K at 250 GPa and 316 K at 167 GPa when the important effects of anharmonicity are included. The very high critical temperatures are attributed to an unusually large hydrogen-derived density of states at the Fermi level arising from the newly reported peculiar H30 as well as H24 cages in the structure. Our predicted introduction of Sc in the La–H system is expected to facilitate future design and realization of hot superconductors in ternary clathrate superhydrides.


The quest for materials that superconduct at and even above room temperature remains one of the most important research topics in condensed matter and materials physics. The discovery of covalent H3S (Tc = 203 K at 155 GPa) (1, 2) and a class of clathrate binary hydrides with high-Tc values in the range of 215 to 260 K at high pressures, including CaH6 (3, 4), YH6 (5, 6), YH9 (6, 7), and LaH10 (810), all of which inspired by theoretical predictions assuming conventional superconductivity (1115), has sparked great interest to search for room-temperature superconductivity among compressed hydrides. The first of these theoretically proposed superconducting clathrate hydrides CaH6 (a predicted Tc of 220 to 235 K at 150 GPa) consisting of sodalite H24 cages (13) was recently confirmed experimentally in two independent studies (3, 4). These exciting findings have encouraged the search for other superconducting hydrides with even higher critical temperatures among clathrate-like hydrogen structures.

Recently, attention has been focused on ternary hydrides, since they provide much wider range of candidate compositions and structures that may exhibit high-temperature superconductivity (1619). For example, our predicted metastable clathrate hydride Li2MgH16 with the highest Tc calculated to date (~473 to 351 K at 250 to 300 GPa) (20) was designed through a strategy of introducing additional electrons from the Li guest metal into the host H2-based MgH16 (21). More recently, guided by theoretical predictions (2224), the first stoichiometric high-Tc ternary LaBeH8 hydride superconductor with a well-defined crystal structure was synthesized with a measured Tc of 110 K at 80 GPa (25). Previous studies have also reported a series of nonstoichiometric high-Tc ternary alloy hydrides, where the random occupation of metal atoms in the lattice sites can render the crystal structure effectively identical to the parent clathrate structure of the binary systems (2634); these include theoretically predicted CaYH12 (258 K at 200 GPa) (26) and YLu3H24 (288 K at 110 GPa) (29), as well as the experimentally synthesized (La,Y)H10 (~253 K at 183 GPa) (31) and (La,Ce)H9 (148 to 178 K at 97 to 172 GPa) (32, 33). These findings suggest the prospect of finding superconductors with the Tc exceeding room temperature in ternary hydrides. While ultimately such superconductors need to be stabilized and function at much lower pressures, it remains critical to establish whether any upper Tc limits exist for hydride superconductors. Therefore, searching for thermodynamically stable superconducting ternary hydrides with higher Tc than LaH10 (250 to 260 K at ~180 GPa) is a pressing task.

LaH10 was the first metal superhydride (XHn, n > 6) to be predicted (14, 15), synthesized (35), and shown experimentally to be a near room-temperature superconductor (8), a result that has now been reproduced in experiments on the end-member and LaH10-based alloys by at least three other groups (9, 10, 31, 34). Additional studies indicated that higher Tc values found for LaH10 can arise from different degrees of nitrogen doping (36) and a decrease in Tc can arise with the addition of magnetic impurities (34). It is of great interest to examine the extent to which superconductivity could be further optimized by introducing other elements to form ternary hydrides with higher critical temperatures. An example is a substitutional alloy of (La,Ce)H9 with a higher Tc than pure CeH9 (32, 33), which is attributed to the element of La having a lower mass than Ce that thereby enhances the Debye temperature of the system. Conversely, note that Sc has a similar valence electronic structure and a lower mass than La, allowing us to expect to tune the superconductivity in the La–H system by the introduction of Sc. To this end, using the swarm-intelligence-based CALYPSO structure prediction method (37, 38) in combination with first-principles calculations, we systematically investigated the crystal structure and superconductivity of the La–Sc–H system under pressure. A remarkable clathrate structure of LaSc2H24 with two unusual H24 and H30 cages was found and shown to be thermodynamically stable above 167 GPa with considering anharmonic effects. Further electron–phonon coupling (EPC) simulations that include anharmonic effects reveal that the predicted phase is a promising hot superconductor with an estimated Tc of 316 K at 167 GPa (peaking at 331 K at 250 GPa). The results indicate the structure has the lowest stability pressure of any hot superconductor studied to date.

Results and Discussion

Based on the results of variable-composition structure searches, we constructed the ternary phase diagram of the La–Sc–H system at 300 GPa. A unique hydrogen-rich compound with the formula LaSc2H24 and space group P6/mmm is identified as thermodynamically stable at 300 GPa (Fig. 1A and SI Appendix, Fig. S1A). As is well known, the contribution of zero-point energy (ZPE) can influence the stability of hydrogen-rich compounds due to the high vibrational frequency associated with the low mass of the H atoms. Including the effect of ZPE, we found that the unique LaSc2H24 compound remains thermodynamically stable at 300 GPa (Fig. 1B and SI Appendix, Fig. S1B). We also investigated the stability of LaSc2H24 at lower pressure by calculating the enthalpy relative to several possible decomposition routes that include the formation of solid hydrogen (39, 40) and binary hydrides (14, 15, 41, 42). Taking into account the effects of both anharmonicity and ZPE, the results indicate that this ternary hydride is thermodynamically stable at pressures as low as 167 GPa (SI Appendix, Fig. S2B).

Fig. 1.

Fig. 1.

Calculated ternary phase diagram of the La–Sc–H system relative to elemental La (4345), Sc (46), and H (39, 40), as well as binary La-H (14, 15, 41), Sc-H (42), and La-Sc at 300 GPa (A) without and (B) with inclusion of ZPE within the harmonic approximation. Red solid circles and colored squares indicate thermodynamically stable and unstable phases, respectively.

Moreover, the ternary LaSc2H24 compound crystallizes in a different superhydride clathrate structure that consists of La-centered H30 cages and Sc-centered H24 cages with the metal atoms arranged in the AlB2 structure type (47) (Fig. 2A). The H cages themselves have unusual structures, not found in known clathrate (1315, 19, 20, 48) or zeolite networks (49): six rhombuses, 12 pentagons, and two hexagons comprise each H30 cage, while each H24 cage consists of six pentagons and three hexagons. Notably, the H–H bond lengths of 1.08, 1.12, 1.16, and 1.20 Å at 300 GPa are slightly longer than the H–H distance (0.98 Å) predicted for monatomic solid hydrogen at 500 GPa (50, 51). On the other hand, these H–H distances are similar to those found in superconducting hydrogen clathrate structures, such as CaH6 (1.24 Å at 150 GPa) (13) and LaH10 (1.07 and 1.16 Å at 300 GPa) (14, 15); the results contrast with the more strongly covalent molecular H–H bond lengths (e.g., <0.8 Å) found in other identified hydrogen-rich compounds that are either low Tc superconductors or even nonmetallic (21, 52, 53), although there are exceptions (54, 55). The H atoms in LaSc2H24 form a weak covalently bonded nonmolecular hydrogen sublattice typical of those found in other clathrate hydride structures and associated with their very high-Tc superconductivity.

Fig. 2.

Fig. 2.

(A) The crystal structure of LaSc2H24 at 300 GPa, consisting of La-centered H30 cages and Sc-centered H24 cages. Each H30 cage contains six rhombuses, 12 pentagons, and two hexagons, while each H24 cage consists of six pentagons and three hexagons. (B) Electronic band structure and (C) PDOS of LaSc2H24 at 300 GPa.

Previous studies have established that the size of the metal atom is crucial to the stability of these hydrogen-rich clathrate structures (e.g., refs. 14 and 15). Subsequent investigations of the relationship between the size of the metal atom M and the structure of the hydrogenic lattice using the Density Functional Theory (DFT)-Chemical Pressure method also showed that La and Ca atoms comfortably fit the H32 cage in MH10 and H24 cage in MH6, respectively (56). Since the atomic radii of Sc and Ca are close and the configurations of H clathrate structures in LaSc2H24 are similar to the H32 cage in MH10 and the H24 cage in MH6, the coordination environment of La and Sc in the H30 and H24 cages, respectively, of the LaSc2H24 structure is consistent with established systematics.

To examine the electronic properties, we calculated the electronic band structure and projected electronic density of states (PDOS) of LaSc2H24 at 300 GPa. As shown in Fig. 2B, several electronic bands cross the Fermi level, indicating the metallic character of LaSc2H24. The PDOS of LaSc2H24 (Fig. 2C) also shows a large H-derived DOS (0.11 states/spin/Ry/Å3) at the Fermi level which is slightly higher than that in LaH10 (0.09 states/spin/Ry/Å3) at 300 GPa (14, 15), indicating that LaSc2H24 is expected to exhibit H-dominated high-temperature superconductivity. Full phonon and EPC calculations were thus carried out within the harmonic approximation. The phonon dispersion calculations revealed an absence of imaginary frequencies and thus dynamical stability of the structure at 300 GPa (Fig. 3A). The phonon density of states (phDOS) shows that the phonons below 16 THz are associated with vibrations of the La and Sc atoms, whereas modes from 16 to 65 THz mainly involve the H atoms (Fig. 3B).

Fig. 3.

Fig. 3.

Lattice dynamics and superconducting properties of LaSc2H24 within the harmonic approximation at 300 GPa. (A) Phonon spectra, where orange solid circles show the phonon linewidth with a radius proportional to its strength, (B) projected phDOS, (C) Eliashberg spectral function α2F(ω) as well as the electron–phonon integral λ(ω), and (D) superconducting gap.

The calculated total EPC constant (λ) obtained by integrating over all available phonon frequencies was found to be 3.34 at 300 GPa. The EPC of H modes dominates the contribution to this large λ (i.e., 74%; Fig. 3C), especially the soft phonons between 16 and 29 THz, which have a large phonon linewidth (Fig. 3A). For strong EPC with λ > 1.5, direct numerical solution of the Eliashberg equations give an accurate description of the superconducting properties (57, 58). The results reveal a temperature-dependent superconducting energy gap with a predicted Tc of 303 to 329 K at 300 GPa, assuming the Coulomb pseudopotential μ* of 0.16 to 0.1 (Fig. 3D). In addition, our harmonic phonon calculations indicate that LaSc2H24 is dynamically stable down to 290 GPa (SI Appendix, Fig. S3A), developing a harmonic instability at lower pressures (SI Appendix, Fig. S4). We further explored the superconductivity in the 290 to 350 GPa pressure range within the harmonic approximation. As can be seen from Table 1, the λ and Tc of LaSc2H24 decrease with increasing pressure, whereas ωlog increases, consistent with hardening on compression of the lower frequency phonons with their large EPC (SI Appendix, Fig. S3). Similar results were found for the LaH10 and YH10 (14, 15).

Table 1.

The calculated EPC constant λ, phonon frequency logarithmic average ωlog (K), electronic density of states at the Fermi level N(Ef) (states/spin/Ry/f.u.), and superconducting critical temperature Tc (K) with μ∗ = 0.16 to 0.1 by numerically solving the isotropic Eliashberg equations for LaSc2H24 within the harmonic and anharmonic approximation at different pressures (GPa).

Calculation Pressure λ ω log N(Ef) T c
Harmonic 290 3.56 865 18.69 311 to 339
300 3.34 968 17.27 303 to 329
350 2.20 1,294 15.64 266 to 294
Anharmonic 150 3.94 745 18.44 281 to 311
167 3.73 794 18.76 285 to 316
200 3.21 965 19.31 291 to 325
250 2.89 1,075 19.63 296 to 331
300 2.61 1,165 19.61 293 to 330

It is known that anharmonic effects of quantum nuclear and/or thermal origin can affect both the dynamic stability and superconductivity in many compressed hydrides (5961). We thus performed anharmonic phonon calculations for LaSc2H24 under pressure using the stochastic self-consistent harmonic approximation (SSCHA) approach (59, 62). SSCHA structural relaxation calculations show that anharmonic effects slightly modify the crystal structure of LaSc2H24 (SI Appendix, Table S2). For example, the calculated structural parameters and H–H distances at 300 GPa increase slightly by 1.7% and 2.8%, respectively, which induces a change in electronic properties, specifically a higher electronic density of states at the Fermi level N(Ef) (Table 1).

To further explore how anharmonic effects affect the superconductivity, we carried out the SSCHA calculations for lattice phonons of LaSc2H24. The renormalized phonon spectra at the anharmonic level indicate that LaSc2H24 is dynamically stable down to 150 GPa (SI Appendix, Fig. S4), which is significantly lower than the harmonic approximation result of 290 GPa. Simulations of the superconductivity that include anharmonicity via renormalized phonon frequencies predict that the critical temperature remains very high, e.g., with Tc = 331 K at 250 GPa (μ*of 0.1; Table 1). The calculated Tc is larger than that originally calculated for LaH10 using similar theoretical techniques (14, 15). Given the close agreement between theory and experiment for the maximum Tc of LaH10 [~280 K (14, 15) and 260 K (810), respectively, near ~200 GPa], the result indicates the potentially important role of incorporating Sc in the La–H system for realizing superconductivity above room temperature. Moreover, the Tc is predicted to remain very high (316 K with μ* of 0.1) at a much more accessible pressure of 167 GPa.

Furthermore, we find that the inclusion of self-consistent anharmonic and nuclear quantum effects significantly expands the stability range of the compound, but this has relatively little effect on the calculated Tc (~330 K at 300 GPa). In addition, the ab initio molecular dynamics (AIMD) simulations at 150 GPa and 300 K show that all atoms remained on the P6/mmm lattice sites (SI Appendix, Fig. S5), which indicates that LaSc2H24 is robust against thermal fluctuations. Finally, with its critical temperature of 316 K at 167 GPa, the LaSc2H24 phase found here is thermodynamically stable at the lowest pressure compared to superconductors reported to date with a predicted Tc above room temperature (Fig. 4).

Fig. 4.

Fig. 4.

Pressure dependence of Tc values for typical superconductors. The green filled circles represent the Tc values of well-known superconductors from experiments (110, 1618, 25, 3133). The blue half-filled (unfilled) rhombuses represent the Tc values of the predicted thermodynamically stable (metastable) hydrides (14, 15, 19, 20, 24, 29, 30, 48). The red half-filled (unfilled) stars mark the Tc values (with μ* of 0.1) of the thermodynamically stable (metastable) LaSc2H24 by considering anharmonic effects at a pressure range of 150 to 300 GPa.

Conclusions

The high-pressure phase diagram of the La–Sc–H system examined using the swarm-intelligence-based CALYPSO structure prediction method in combination with first-principles simulations reveals a unique hexagonal and thermodynamically stable LaSc2H24 phase at megabar pressures. The LaSc2H24 structure consists of a unique three-dimensional H clathrate network containing an unusual combination of H24 and H30 cages. Most strikingly, self-consistent phonon calculations that take into account anharmonic and nuclear quantum effects predict a superconducting Tc well above room temperature in the 167 to 300 GPa pressure range (up to 331 K at 250 GPa). The results expand our perspective on how ternary hydrides offer a platform to engineer phases with a greater diversity of hydrogen frameworks. More than introducing additional elements (e.g., doping) to shift the Fermi level of the parent binary hydride, unique ternary structure types may form with electronic structures that lead to strong EPC and therefore enhanced superconducting properties. Our results further suggest that hot superconducting hydrides with Tcs exceeding 300 K can exist in the experimentally accessible pressure range at 167 GPa and add to the diversity of clathrate hydride structures predicted by theory and confirmed in the laboratory.

Methods

To explore the thermodynamic stability of La–Sc–H compounds, we have utilized the swarm intelligence structure prediction method as implemented in the CALYPSO code (37, 38), conducting variable-composition searches with up to 32 atoms per primitive cell at 300 GPa. For compositions with enthalpies above the convex hull within 50 meV/atom, further fixed composition structural predictions, with system sizes containing up to 2 formula units per simulation cell, were carried out until they reached convergence after generating about 2,000 structures for each stoichiometry.

The following structural optimization and electronic property calculations were performed using DFT within the Perdew–Burke–Ernzerhof parametrization of the generalized gradient approximation (63), as implemented in the Vienna ab initio simulation package code (64). The all-electron projector augmented wave (PAW) method (65) is employed with valence electrons of 5s25p65d16s2, 3s23p63d15s2, and 1s1 for La, Sc, and H, respectively. A plane-wave basis set with an energy cutoff of 850 eV and uniform Γ-centered k-point grids with a resolution of 2π × 0.03 Å−1 were used in the electronic self-consistent calculations. The structures were fully optimized until the maximum energy and force were less than 10−8 eV and 1 meV/Å, respectively.

The harmonic phonon frequencies and EPC were calculated by the Quantum-ESPRESSO package within the density functional perturbation theory (66), where the PAW pseudopotentials (67) were used with a kinetic energy cutoff of 90 Ry and a charge density cutoff of 720 Ry. Self-consistent electron density and EPC were calculated by employing 24 × 24 × 24 k-point meshes and 6 × 6 × 6 q-point meshes for LaSc2H24. The anharmonic effects were estimated using the SSCHA code (59, 62). The SSCHA calculations were performed using a 2 × 2 × 2 supercell containing 216 atoms at 0 K, which yielded anharmonic dynamical matrices on a commensurate 2 × 2 × 2 q-point grid.

An 80 Ry energy cutoff and a 3 × 3 × 3 k-point mesh for the Brillouin zone integrations were sufficient in the supercell to converge the SSCHA gradient. The anharmonic phonon frequencies were obtained from the positional free energy Hessian including the third-order terms. Harmonic phonon frequencies and electron–phonon matrix elements were calculated on a grid of 6 × 6 × 6 points. The difference between the harmonic and anharmonic dynamical matrices in the 2 × 2 × 2 phonon-momentum grid was interpolated to a 6 × 6 × 6 grid. Adding the harmonic 6 × 6 × 6 grid dynamical matrices to the result, the anharmonic 6 × 6 × 6 q-grid dynamical matrices were obtained. The ZPE is calculated within the harmonic approximation at 300 GPa. To evaluate the lowest pressure required for the thermodynamic stability of LaSc2H24, we recalculated the ZPE of the ternary hydride and selected decomposition routes at the anharmonic level by using the SSCHA within a pressure range of 150 to 300 GPa. In addition, the superconducting gap and Tc values at the harmonic and anharmonic levels were calculated by numerically solving the isotropic Eliashberg equations (57, 58). The Matsubara frequency cutoff is taken to be about 10 times the highest phonon frequency.

The AIMD simulations were carried out using a 2 × 2 × 3 supercell with 324 atoms for LaSc2H24, with Γ point for k-mesh sampling. A Nosé–Hoover thermostat was adopted to perform the NVT simulations with a temperature of 300 K and a pressure of 150 GPa lasting for 20 ps with a time step of 0.5 fs. The statistical information of mean square displacements and atomic trajectories was extracted from the last 15 ps.

Supplementary Material

Appendix 01 (PDF)

Acknowledgments

This work was supported by the National Key Research and Development Program of China (Grant No. 2022YFA1402304), National Natural Science Foundation of China (Grant No. 12074138, 52288102, 52090024, and 12374007), Program for Jilin University Science and Technology Innovative Research Team (2021TD-05), the Program for Jilin University Computational Interdisciplinary Innovative Platform, the Strategic Priority Research Program of Chinese Academy of Sciences (Grant No. XDB33000000), the Royal Society International Exchange Scheme (Grant No. IEC\NSFC\201359), US NSF (DMR-2104881), and the Fundamental Research Funds for the Central Universities and computing facilities at the High-Performance Computing Centre of Jilin University.

Author contributions

H.L. designed research; X.-L.H. and W.Z. performed research; X.-L.H., W.Z., Y.X., A.H., R.J.H., H.L., and Y.M. analyzed data; and X.-L.H., W.Z., Y.X., A.H., R.J.H., H.L., and Y.M. wrote the paper.

Competing interests

The authors declare no competing interest.

Footnotes

This article is a PNAS Direct Submission.

Data, Materials, and Software Availability

All study data are included in this article and/or SI Appendix.

Supporting Information

References

  • 1.Drozdov A. P., Eremets M. I., Troyan I. A., Ksenofontov V., Shylin S. I., Conventional superconductivity at 203 kelvin at high pressures in the sulfur hydride system. Nature 525, 73–76 (2015). [DOI] [PubMed] [Google Scholar]
  • 2.Einaga M., et al. , Crystal structure of the superconducting phase of sulfur hydride. Nat. Phys. 12, 835–838 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Ma L., et al. , High-temperature superconducting phase in clathrate calcium hydride CaH6 up to 215 K at a pressure of 172 GPa. Phys. Rev. Lett. 128, 167001 (2022). [DOI] [PubMed] [Google Scholar]
  • 4.Li Z., et al. , Superconductivity above 200 K discovered in superhydrides of calcium. Nat. Commun. 13, 2863 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Troyan I. A., et al. , Anomalous high-temperature superconductivity in YH6. Adv. Mater. 33, 2006832 (2021). [DOI] [PubMed] [Google Scholar]
  • 6.Kong P., et al. , Superconductivity up to 243 K in the yttrium-hydrogen system under high pressure. Nat. Commun. 12, 5075 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Snider E., et al. , Synthesis of yttrium superhydride superconductor with a transition temperature up to 262 K by catalytic hydrogenation at high pressures. Phys. Rev. Lett. 126, 117003 (2021). [DOI] [PubMed] [Google Scholar]
  • 8.Somayazulu M., et al. , Evidence for superconductivity above 260 K in lanthanum superhydride at megabar pressures. Phys. Rev. Lett. 122, 027001 (2019). [DOI] [PubMed] [Google Scholar]
  • 9.Drozdov A. P., et al. , Superconductivity at 250 K in lanthanum hydride under high pressures. Nature 569, 528–531 (2019). [DOI] [PubMed] [Google Scholar]
  • 10.Hong F., et al. , Superconductivity of lanthanum superhydride investigated using the standard four-probe configuration under high pressures. Chin. Phys. Lett. 37, 107401 (2020). [Google Scholar]
  • 11.Li Y., Hao J., Liu H., Li Y., Ma Y., The metallization and superconductivity of dense hydrogen sulfide. J. Chem. Phys. 140, 174712 (2014). [DOI] [PubMed] [Google Scholar]
  • 12.Duan D., et al. , Pressure-induced metallization of dense (H2S)2H2 with high-Tc superconductivity. Sci. Rep. 4, 6968 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Wang H., Tse J. S., Tanaka K., Iitaka T., Ma Y., Superconductive sodalite-like clathrate calcium hydride at high pressures. Proc. Natl. Acad. Sci. U.S.A. 109, 6463–6466 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Peng F., et al. , Hydrogen clathrate structures in rare earth hydrides at high pressures: Possible route to room-temperature superconductivity. Phys. Rev. Lett. 119, 107001 (2017). [DOI] [PubMed] [Google Scholar]
  • 15.Liu H., Naumov I. I., Hoffmann R., Ashcroft N. W., Hemley R. J., Potential high-Tc superconducting lanthanum and yttrium hydrides at high pressure. Proc. Natl. Acad. Sci. USA 114, 6990–6995 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Flores-Livas J. A., et al. , A perspective on conventional high-temperature superconductors at high pressure: Methods and materials. Phys. Rep. 856, 1–78 (2020). [Google Scholar]
  • 17.Boeri L., et al. , The 2021 room-temperature superconductivity roadmap. J. Phys.: Condens. Matter 34, 183002 (2022). [DOI] [PubMed] [Google Scholar]
  • 18.Hilleke K. P., Zurek E., Tuning chemical precompression: Theoretical design and crystal chemistry of novel hydrides in the quest for warm and light superconductivity at ambient pressures. J. Appl. Phys. 131, 070901 (2022). [Google Scholar]
  • 19.An D., et al. , Thermodynamically stable room-temperature superconductors in Li-Na hydrides under high pressures. arXiv [Preprint] (2023). 10.48550/arXiv.2303.09805. Accessed 17 March 2023. [DOI]
  • 20.Sun Y., Lv J., Xie Y., Liu H., Ma Y., Route to a superconducting phase above room temperature in electron-doped hydride compounds under high pressure. Phys. Rev. Lett. 123, 097001 (2019). [DOI] [PubMed] [Google Scholar]
  • 21.Lonie D. C., Hooper J., Altintas B., Zurek E., Metallization of magnesium polyhydrides under pressure. Phys. Rev. B 87, 054107 (2013). [Google Scholar]
  • 22.Di Cataldo S., Heil C., von der Linden W., Boeri L., LaBH8: Towards high-Tc low-pressure superconductivity in ternary superhydrides. Phys. Rev. B 104, L020511 (2021). [Google Scholar]
  • 23.Liang X., et al. , Prediction of high-Tc superconductivity in ternary lanthanum borohydrides. Phys. Rev. B 104, 134501 (2021). [Google Scholar]
  • 24.Zhang Z., et al. , Design principles for high-temperature superconductors with a hydrogen-based alloy backbone at moderate pressure. Phys. Rev. Lett. 128, 047001 (2022). [DOI] [PubMed] [Google Scholar]
  • 25.Song Y., et al. , Stoichiometric ternary superhydride LaBeH8 as a new template for high-temperature superconductivity at 110 K under 80 GPa. Phys. Rev. Lett. 130, 266001 (2023). [DOI] [PubMed] [Google Scholar]
  • 26.Liang X., et al. , Potential high-Tc superconductivity in CaYH12 under pressure. Phys. Rev. B 99, 100505(R) (2019). [Google Scholar]
  • 27.Sukmas W., Tsuppayakorn-aek P., Pinsook U., Bovornratanaraks T., Near-room-temperature superconductivity of Mg/Ca substituted metal hexahydride under pressure. J. Alloys Compd. 849, 156434 (2020). [Google Scholar]
  • 28.Song P., Hou Z., Nakano K., Hongo K., Maezono R., Potential high-Tc superconductivity in YCeHx and LaCeHx under pressure. Mater. Today Phys. 28, 100873 (2022). [Google Scholar]
  • 29.Du M., Song H., Zhang Z., Duan D., Cui T., Room-temperature superconductivity in Yb/Lu substituted clathrate hexahydrides under moderate pressure. Research 2022, 9784309 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Liu L., et al. , Generic rules for achieving room-temperature superconductivity in ternary hydrides with clathrate structures. Phys. Rev. B 107, L020504 (2023). [Google Scholar]
  • 31.Semenok D. V., et al. , Superconductivity at 253 K in lanthanum-yttrium ternary hydrides. Mater. Today 48, 18–28 (2021). [Google Scholar]
  • 32.Bi J., et al. , Giant enhancement of superconducting critical temperature in substitutional alloy (La, Ce)H9. Nat. Commun. 13, 5952 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Chen W., et al. , Enhancement of superconducting properties in the La-Ce-H system at moderate pressures. Nat. Commun. 14, 2660 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Semenok D. V., et al. , Effect of magnetic impurities on superconductivity in LaH10. Adv. Mater. 34, 2204038 (2022). [DOI] [PubMed] [Google Scholar]
  • 35.Geballe Z. M., et al. , Synthesis and stability of lanthanum superhydrides. Angew. Chem. Int. Ed. Engl. 130, 696–700 (2018). [DOI] [PubMed] [Google Scholar]
  • 36.Ge Y., Zhang F., Hemley R. J., Room-temperature superconductivity in boron- and nitrogen-doped lanthanum superhydride. Phys. Rev. B 104, 214505 (2021). [Google Scholar]
  • 37.Wang Y., Lv J., Zhu L., Ma Y., Crystal structure prediction via particle-swarm optimization. Phys. Rev. B 82, 094116 (2010). [Google Scholar]
  • 38.Wang Y., Lv J., Zhu L., Ma Y., CALYPSO: A method for crystal structure prediction. Comput. Phys. Commun. 183, 2063–2070 (2012). [Google Scholar]
  • 39.Pickard C. J., Needs R. J., Structure of phase III of solid hydrogen. Nat. Phys. 3, 473–476 (2007). [Google Scholar]
  • 40.Liu H., Wang H., Ma Y., Quasi-molecular and atomic phases of dense solid hydrogen. J. Phys. Chem. C 116, 9221–9226 (2012). [Google Scholar]
  • 41.Kruglov I. A., et al. , Superconductivity of LaH10 and LaH16 polyhydrides. Phys. Rev. B 101, 024508 (2020). [Google Scholar]
  • 42.Ye X., Zarifi N., Zurek E., Hoffmann R., Ashcroft N. W., High hydrides of scandium under pressure: Potential superconductors. J. Phys. Chem. C 122, 6298–6309 (2018). [Google Scholar]
  • 43.Porsch F., Holzapfel W. B., Novel reentrant high pressure phase transtion in lanthanum. Phys. Rev. Lett. 70, 4087–4089 (1993). [DOI] [PubMed] [Google Scholar]
  • 44.Chen W., et al. , Superconductivity and equation of state of lanthanum at megabar pressures. Phys. Rev. B 102, 134510 (2020). [Google Scholar]
  • 45.Chen L., et al. , Phase transitions and properties of lanthanum under high pressures. J. Phys.: Condens. Matter 34, 204005 (2022). [DOI] [PubMed] [Google Scholar]
  • 46.Akahama Y., Fujihisa H., Kawamura H., New helical chain structure for scandium at 240 GPa. Phys. Rev. Lett. 94, 195503 (2005). [DOI] [PubMed] [Google Scholar]
  • 47.Hofmann W., Jäniche W., Die struktur von aluminiumborid AlB2. Z. Phys. Chem. 31B, 214–222 (1936). [Google Scholar]
  • 48.Zhong X., et al. , Prediction of above-room-temperature superconductivity in lanthanide/actinide extreme superhydrides. J. Am. Chem. Soc. 144, 13394–13400 (2022). [DOI] [PubMed] [Google Scholar]
  • 49.The database of zeolite structures, http://asia.iza-structure.org/IZA-SC/ftc_table.php.
  • 50.Labet V., Gonzalez-Morelos P., Hoffmann R., Ashcroft N. W., A fresh look at dense hydrogen under pressure. I. An introduction to the problem, and an index probing equalization of H-H distances. J. Chem. Phys. 136, 074501 (2012). [DOI] [PubMed] [Google Scholar]
  • 51.McMahon J. M., Ceperley D. M., Ground-state structures of atomic metallic hydrogen. Phys. Rev. Lett. 106, 165302 (2011). [DOI] [PubMed] [Google Scholar]
  • 52.Semenok D. V., et al. , Sr-doped superionic hydrogen glass: Synthesis and properties of SrH22. Adv. Mater. 34, 2200924 (2022). [DOI] [PubMed] [Google Scholar]
  • 53.Baettig P., Zurek E., Pressure-stabilized sodium polyhydrides: NaHn (n > 1). Phys. Rev. Lett. 106, 237002 (2011). [DOI] [PubMed] [Google Scholar]
  • 54.Chen B., et al. , Phase stability and superconductivity of lead hydrides at high pressure. Phys. Rev. B 103, 035131 (2021). [Google Scholar]
  • 55.Liu Z., et al. , Emergence of near room-temperature superconductivity in hydrides with H2 molecular units. Phys. Rev. B 109, L180501 (2024). [Google Scholar]
  • 56.Hilleke K. P., Zurek E., Rational design of superconducting metal hydrides via chemical pressure tuning. Angew. Chem. Int. Ed. Engl. 61, e202207589 (2022). [DOI] [PubMed] [Google Scholar]
  • 57.Sanna A., et al. , Ab initio eliashberg theory: Making genuine predictions of superconducting features. J. Phys. Soc. Jpn. 87, 041012 (2018). [Google Scholar]
  • 58.The elk code, which is an all-electron full-potential linearised augmented-plane wave code with many advanced features, https://elk.sourceforge.io/.
  • 59.Errea I., Calandra M., Mauri F., First-principles theory of anharmonicity and the inverse isotope effect in superconducting palladium-hydride compounds. Phys. Rev. Lett. 111, 177002 (2013). [DOI] [PubMed] [Google Scholar]
  • 60.Errea I., et al. , High-pressure hydrogen sulfide from first principles: A strongly anharmonic phonon-mediated superconductor. Phys. Rev. Lett. 114, 157004 (2015). [DOI] [PubMed] [Google Scholar]
  • 61.Errea I., et al. , Quantum crystal structure in the 250-kelvin superconducting lanthanum hydride. Nature 578, 66–69 (2020). [DOI] [PubMed] [Google Scholar]
  • 62.Monacelli L., et al. , The stochastic self-consistent harmonic approximation: Calculating vibrational properties of materials with full quantum and anharmonic effects. J. Phys.: Condens. Matter 33, 363001 (2021). [DOI] [PubMed] [Google Scholar]
  • 63.Perdew J. P., Burke K., Ernzerhof M., Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868 (1996). [DOI] [PubMed] [Google Scholar]
  • 64.Kresse G., Furthmüller J., Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 54, 11169–11186 (1996). [DOI] [PubMed] [Google Scholar]
  • 65.Blöchl P. E., Projector augmented-wave method. Phys. Rev. B 50, 17953–17979 (1994). [DOI] [PubMed] [Google Scholar]
  • 66.Giannozzi P., et al. , QUANTUM ESPRESSO: A modular and open-source software project for quantum simulations of materials. J. Phys.: Condens. Matter 21, 395502 (2009). [DOI] [PubMed] [Google Scholar]
  • 67.Prandini G., Marrazzo A., Castelli I. E., Mounet N., Marzari N., Precision and efficiency in solid-state pseudopotential calculations. npj Comput. Mater. 4, 72 (2018). [Google Scholar]

Associated Data

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

Supplementary Materials

Appendix 01 (PDF)

Data Availability Statement

All study data are included in this article and/or SI Appendix.


Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES