DOI:
10.1039/D5TC02600H
(Paper)
J. Mater. Chem. C, 2025,
13, 20571-20579
Prediction of high-temperature superconductivity in LaH4 at low pressures
Received
7th July 2025
, Accepted 7th September 2025
First published on 15th September 2025
Abstract
Superconducting hydrides have received significant attention in the last decade due to their potential for room-temperature superconductivity. However, achieving high critical temperatures (Tcs) typically requires extreme pressures exceeding 150 GPa. Recently, a new, low-pressure R
m-LaH4 phase was observed to form above approximately 20 GPa. Here, we perform first-principles calculations to investigate the electron–phonon interactions and superconducting properties of the new phase across a range of pressures. At the harmonic level, the system is found to be dynamically unstable, but this is remedied through the inclusion of anharmonic effects. We estimate that Tc reaches up to 115 K at 25 GPa, driven by a high density of states at the Fermi level (NF) and soft phonon modes. However, superconductivity is suppressed with increasing pressure, as Tc rapidly decreases to 34 K at 60 GPa and 11 K at 100 GPa, due to a reduction in NF and phonon hardening.
1 Introduction
Room-temperature superconductivity has long been a highly sought after goal in materials science. The discovery of superconductivity in H3S at 203 K and 155 GPa1 marked a pivotal breakthrough, demonstrating that high-Tc superconductivity could be realized in hydrogen-rich systems under pressure. This was soon followed by the realization of near-room-temperature superconductivity in LaH10 at 250 K and 170 GPa,2,3 firmly establishing binary hydrides as front-runners in the search for high-Tc materials. Yttrium hydrides, such as YH6 and YH9, have since been both predicted and experimentally confirmed to reach Tc values above 240 K at megabar pressures.4,5 Other binary hydrides have been experimentally reported to exhibit superconductivity, including CaH6 (215 K at 172 GPa),6,7 CeH9 (100 K at 100 GPa),8–10 and ThH10 (161 K at 175 GPa).11 These findings demonstrate that binary hydrides have a strong potential to achieve high-Tc, particularly when the hydrogen fraction in a compound exceeds 0.6.12 The high phonon frequencies of hydrogen and the lower frequency modes from the heavier ions both couple strongly to result in significant electron–phonon (e–ph) coupling, and, therefore, Tc. Although these superhydrides significantly reduce the pressure needed to dissociate molecular H2 by providing a lattice structure that “chemically pre-compresses” the hydrogen,13 the required pressures remain prohibitively high for practical applications.
Among binary hydrides, significant research has focused on the lanthanum hydrides following the report of superconductivity in LaH10.2,3,14–27 In this case, studies have demonstrated strong e-ph coupling,20 stability at lower pressures,17 and the importance of anharmonic corrections for achieving dynamical stability.18 Recently, several lanthanum hydrides with different La
:
H ratios have been successfully synthesized in the pressure range of 50 to 180 GPa,28 with the cubic La4H23 phase shown to exhibit superconductivity with a maximum Tc of 105 K at 118 GPa.29,30
Building on these results and motivated by the need to identify hydrogen-rich compounds that can achieve high-Tc superconductivity at lower pressures, some of the present authors investigated the formation of La hydrides in cryomilled elemental lanthanum exposed to ammonia borane31. Structural evolution under pressures up to 60 GPa revealed a distortion from the Fm
m phase to an R
m phase, with the new rhombohedral structure becoming kinetically stable at pressures above approximately 20 GPa. Using the observed excess cell volume and comparison with the equation of state calculated via density functional theory (DFT) for various lanthanum hydrides, the stoichiometry of the R
m phase was determined to correspond to LaHx with x ∼ 4.31
Based on these experimental observations, we perform first-principles calculations to investigate the superconducting properties of the rhombohedral LaH4 structure shown in Fig. 1. Our calculations predict a Tc reaching 115 K at 25 GPa, nearly half that of LaH1015,19 and comparable to La4H23,29 but at significantly lower pressures. For instance, a tetragonal LaH4 phase was previously predicted to exhibit a Tc of only 10 K at 300 GPa.15 Our findings reinforce the potential of the lanthanum hydride family for the development of high-temperature superconducting materials at low to ambient pressures.
 |
| | Fig. 1 Crystal structure of R m-LaH4 at 25 GPa. The two nonequivalent hydrogen atoms are labeled as H1 and H2 and the shortest H–H distance is indicated as dH–H. The figures were generated with VESTA.47 | |
2 Methods
First principle calculations were carried out with the Quantum ESPRESSO (QE) package.32,33 We employed optimized norm-conversing Vanderbilt pseudopotentials (ONCVPSP)34 from the Pseudo Dojo library35 generated with the scalar-relativistic revised Perdew–Burke–Ernzerhof (PBEsol) parametrization.36 A plane wave kinetic-energy cutoff of 120 Ry for the wavefunctions, and 480 Ry for the charge density and potential, were used. For the Brillouin-zone integration of the 5-atom unit cell, we used a Γ-centered 12 × 12 × 12 k-mesh37 with a Methfessel–Paxton38 smearing of 0.02 Ry. The atomic positions and lattice parameters were optimized until the total energy was converged within 10−6 Ry and the force on each atom was less than 10−4 Ry Å−1. The dynamical matrices and the linear variation of the self-consistent potential were first computed using density-functional perturbation theory (DFPT)39 on the irreducible set of a regular 6 × 6 × 6 q-mesh.
The force constants computed by DFPT were used as a starting point to calculate phonon anharmonicity via the anharmonic special displacement method (ASDM)40–42 within the framework of the self-consistent phonon theory.43 The ASDM is used as implemented in the Zacharias–Giustino (ZG) code of the EPW package.44 For our ASDM calculations, we employed finite differences of amplitude 0.2 Å to iteratively compute the interatomic force constants of ZG configurations at 300 K in 2 × 2 × 2 supercells. We find that such large displacements in low-symmetry systems involving light-mass atoms improve numerical stability and provide a better representation of the anharmonic behavior, particularly for capturing higher-order contributions to the effective force constants. In contrast to high-symmetry hydrides, such as the Im
m phase of H3S, where small 0.01 Å displacements work effectively,45 such small displacement values lead to phonon instabilities in the soft-mode regions near Γ, as shown in Fig. S1.46 With the aid of iterative mixing,42 the convergence of the interatomic force constants was achieved at the 10th iteration, where three successive iterations showed little, to no, change in the anharmonic phonon dispersion. The converged anharmonic force constants were then used to construct the dynamical matrix via standard Fourier interpolation on a 6 × 6 × 6 q-mesh.
The EPW44,48–50 code was employed to investigate the e–ph interactions and superconducting properties. The electronic wavefunctions required for the Wannier–Fourier interpolation51,52 were calculated on a Γ-centered 12 × 12 × 12 k-mesh. Nine maximally localized Wannier functions (five d orbitals for La and one s orbital for each H) were used to describe the electronic structure. The anisotropic full-bandwidth Migdal–Eliashberg equations44,53 were solved with a sparse intermediate representation of the Matsubara frequencies54 on fine uniform 80 × 80 × 80 k- and 40 × 40 × 40 q-grids with an energy window of ± 0.2 eV around the Fermi level. The semiempirical Coulomb parameter, μ*, was varied from 0.1–0.2.
3 Results and discussion
Guided by experiments, we consider the R
m rhombohedral phase (space group no. 166) for the LaH4 structure. Fig. 2(a) shows our X-ray Diffraction (XRD) results for cryomilled La mixed with ammonia borane, measured in a diamond anvil cell at a pressure of 29.1 GPa. The XRD data reveal the co-existence of R
m and FCC La-hydride phases. Fig. 2(b) shows the cell volume of La-hydride inferred from XRD data as a function of pressure. The experimental cell volume is systemically larger than that of the FCC LaH3 phase over the investigated pressure range.
 |
| | Fig. 2 (a) Powder X-ray diffraction data for cryomilled La mixed with ammonia borane, showing the co-existence of R m and FCC La-hydride phases under high pressure (29.1 GPa). (b) The cell volume of the La-hydride phase inferred from X-ray diffraction data as a function of pressure. DFT simulations of the equation of state are shown for the proposed R m LaH4 phase and the FCC LaH3 phase, as indicated. The data presented in this figure are from ref. 31. | |
Utilizing the experimentally determined unit cell and positions of the La atoms, the positions of the H atoms were derived in ref. 31 based on symmetry considerations and atomic relaxation using DFT. The enthalpy of the predicted rhombohedral R
m LaH4 structure was found to be lower by about 0.2 eV f.u.−1 than that of several alternative, symmetry-broken LaH4 configurations obtained via ab initio molecular dynamics followed by atomic relaxation. On the other hand, the I4/mmm proposed in ref. 15 remains thermodynamically more favorable than the R
m phase in the considered pressure range, as shown in Fig. S2.46 While the R
m phase is not the thermodynamic ground state, our theoretical calculations (including anharmonic effects) indicate that it is dynamically stable, meaning it is kinetically stable in the limit of low temperature. Our experimental data corroborate this, suggesting that the R
m phase is observable (thermodynamically metastable) at room temperature. This indicates that it is even more likely to be observable (kinetically stabilized) at low temperatures, particularly near the predicted transition temperature Tc. A similar kinetics-based stabilization mechanism has been shown to overcome the highly unfavorable thermodynamics of Alane at room temperature.55
The R
m-LaH4 phase is predicted to adopt a clathrate structure that is characteristic of other high-temperature superconducting hydrides.7,18,56–63 The structure consists of a La atom occupying the 3a (0, 0, 0) Wyckoff site and two nonequivalent H atoms, labeled as H1 and H2, located at 6c (0, 0, z1) and 6c (1/3, 2/3, z2) Wyckoff sites. The calculated structural parameters at 25, 60, and 100 GPa are provided in Table 1. Each La atom is enclosed in a cage formed by 14 H atoms, as seen in Fig. 1. This matches the crystal structure of R
m-TeH4 predicted to stabilize above 230 GPa and reach a Tc of about 70 K.64 The H-hexagonal rings are slightly buckled at 25 GPa (∼0.1 Å) but, as in R
m-TeH4, become nearly flat at higher pressures. The H1 atoms form a H2 unit as marked by the red bond in Fig. 1. This bond marks the shortest H–H distance in the structure, with a value dH–H of 0.96, 0.92, and 0.88 Å for 25, 60, and 100 GPa, respectively. These bond length values are close to the H–H distance of 1 Å predicted for metallic hydrogen near 500 GPa65 and 0.86 Å for TeH4 at 300 GPa,64 but slightly smaller than 1.1 Å for LaH10 in the 150–300 GPa pressure range15,66 and 1.3 Å for La4H23 at 118 GPa.29
Table 1 Calculated lattice parameters, shortest H–H distance, and atomic coordinates for the conventional cell of LaH4 at 25, 60, and 100 GPa. The structure adopts the R
m space group with γ = 120° at all pressures
| Pressure (GPa) |
a = b (Å) |
c (Å) |
d
H–H (Å) |
Atomic coordinates (fractional) |
| 25 |
3.54 |
9.87 |
0.96 |
La |
(3a) |
0.000 |
0.000 |
0.000 |
| H1 |
(6c) |
0.000 |
0.000 |
0.231 |
| H2 |
(6c) |
0.333 |
0.666 |
0.217 |
| 60 |
3.20 |
9.89 |
0.92 |
La |
(3a) |
0.000 |
0.000 |
0.000 |
| H1 |
(6c) |
0.000 |
0.000 |
0.215 |
| H2 |
(6c) |
0.333 |
0.666 |
0.213 |
| 100 |
2.99 |
9.78 |
0.88 |
La |
(3a) |
0.000 |
0.000 |
0.000 |
| H1 |
(6c) |
0.000 |
0.000 |
0.209 |
| H2 |
(6c) |
0.333 |
0.666 |
0.212 |
The electronic band structure and density of state (DOS) plots in Fig. 3 indicate that the R
m-LaH4 phase is metallic across all investigated pressures. The DOS at the Fermi level (NF) comes from a combination of La and H states with La boasting about 50%, 68%, and 81% increase over H states at 25, 60, and 100 GPa. For hydrogen, the unit formed by the vertically adjacent H2 atoms, as seen in Fig. 1, account for the bulk of the hydrogen states at the Fermi level, contributing 86%, 82%, and 81% of the total hydrogen states at the Fermi level, as depicted in the second panel of Fig. 3. At 25 GPa, NF reaches its highest value of 0.81 states per eV f.u.−1, comparable to values reported for other high-Tc La–H phases (e.g., ∼ 0.93 states per eV f.u.−1 for LaH10 at 150 GPa18 and 0.79 states per eV per La for La4H23 at 120 GPa29). This high DOS is largely attributed to a flat electronic band around the Γ point. Furthermore, approximately 0.2 eV below the Fermi level, a Van-Hove-like singularity emerges due to another flat band along the L-B1 and B-Z high-symmetry directions. The presence of these dispersionless bands at 25 GPa induces a significant variation in the DOS in the vicinity of EF that is captured by the full-bandwidth approach used to solve the superconducting Migdal–Eliashberg equations.53 At higher pressures, the flat bands evolve into parabolic dispersions, resulting in a marked decrease in NF by approximately 33% at 60 GPa and 46% at 100 GPa, respectively. To note, the effect of spin–orbit coupling on the electronic structure was considered and found to be negligible, as shown in Fig. S3.46
 |
| | Fig. 3 Electronic band structure with orbital characters, and total and projected density of states (DOS) of LaH4 at (a)–(c) 25, 60, and 100 GPa. The orbital character is shown in red for H and blue for La states. The total DOS is shown in black, while the projected DOS corresponding to H1, H2, and La atoms are shown with red, dashed red, and blue lines, respectively. | |
To further analyze the electronic structure, we calculated the charge density difference between the LaH4 structure and the sum of its constituent atoms at 25 and 100 GPa. As shown in Fig. 4(a) and (b), there is a buildup of charge on the hydrogen atoms, but, surprisingly, no charge depletion around the La atoms. This suggests minimal to no charge transfer between La and H atoms. Meanwhile, charge depletion is observed between two vertically adjacent H2 atoms, becoming significantly more pronounced at 100 GPa. This trend indicates the formation of H2 units at higher pressure, resembling molecular hydrogen. This picture is further supported by the electron localization function (ELF) plots in Fig. 4(c) and (d), taken along the (1 1 0) and (0 0 1) planes. Panel (c) reveals the formation of H2 units between adjacent H2 atoms, as well as the repulsion between La and H1 atoms, which gives rise to the teardrop-shaped features around the H1 atoms. Panel (d) illustrates how the H atoms arrange into a hexagonal ring surrounding the La atom, highlighting the structural characteristics of R
m-LaH4.
 |
| | Fig. 4 Charge density difference between the LaH4 crystal structure and the sum of its constituent atoms at (a) 25 GPa and (b) 100 GPa. The yellow and green colors represent charge accumulation and depletion regions, respectively, with an isosurface value set to 7 × 10−2 e Å−3. Electron localization function (ELF) for LaH4 at 25 GPa taken along the (c) (1 1 0) and (d) (0 0 1) Miller planes, respectively, with a contour spacing set to 0.1. All plots were generated with VESTA.47 | |
Fig. 5 presents the vibrational properties of LaH4. At the harmonic level, depicted by the solid gray lines, the structure exhibits dynamic instabilities across all investigated pressures, with the most pronounced negative phonon modes observed at 25 GPa. This behavior is common among many superconducting hydrides and is typically resolved by accounting for quantum anharmonic effects.18,68–71 When applying the ASDM, these instabilities are eliminated, as shown by the solid black lines. Notably, the low-frequency La modes at 25 GPa undergo significant renormalization under the ASDM, which has important implications for the superconducting properties of the material.
 |
| | Fig. 5 (a)–(c) Harmonic and anharmonic phonon dispersion, phonon density of states (PHDOS), isotropic Eliashberg spectral function (α2F), and electron–phonon coupling strength (λ) of LaH4 at 25, 60, and 100 GPa, respectively. The total PHDOS is decomposed with respect to the vibrations of H (red) and La (blue) atoms. The anharmonic phonon dispersions were computed for a temperature of 300 K. | |
The phonon dispersion displays a distinct separation between modes associated with lanthanum and hydrogen. The acoustic modes arise solely from the vibration of the La atoms, while the intermediate- and high-frequency modes originate exclusively from the motion of the H atoms. From the comparison of the Eliashberg spectral function (α2F) and the phonon density of states (PHDOS) in the middle and right panels of Fig. 5, we can see that most of the e–ph coupling comes from the hydrogen phonon modes. At 25 GPa, for example, 66% of the total e–ph coupling constant λ of 1.64 comes from the H modes. This λ value is comparable to those computed for LaH10 (1.78 and 1.86 at 300 GPa15,19) and La4H23 (1.49 at 100 GPa72), and almost four times larger than that of a predicted tetragonal LaH4 phase (0.43 at 300 GPa15). As pressure increases, the phonon frequencies harden across all regions, and the DOS at the Fermi level decreases. These combined effects lead to a reduction in λ, with values of 0.79 and 0.53 at 60 GPa and 100 GPa, respectively. A similar pressure dependence has been observed in LaH10 and La4H23 superconductors.3,15,29
To understand the superconducting properties, we solved the anisotropic full-bandwidth Migdal–Eliashberg equations for a Coulomb parameter μ* in the 0.10–0.20 range, consistent with values used in other studies on superconductivity in lanthanum hydrides.15,19,29Fig. 6 presents the energy distribution of the superconducting gap, Δnk, as a function of temperature along with its momentum-resolved k-dependence on the Fermi surface at select pressure and μ* values. Our calculations reveal that LaH4 exhibits a single anisotropic gap, with an estimated TaMEc of 115 K for μ* = 0.10 at 25 GPa. In comparison, the McMillan (McM) equation,73 the Allen–Dynes (AD) formula,74 the machine-learned (ML) SISSO model,75 and the isotropic Migdal–Eliashberg (iME) formalism yield TMcMc = 72.5 K, TADc = 85.4 K, TMLc = 102.7 K, and TiMEc = 117 K, respectively. To put this into context, TiMEc was predicted to be 254 K at 300 GPa in LaH10,15 95 K at 100 GPa in La4H23,29 and only 10 K at 300 GPa in a tetragonal LaH4 phase15 for μ* = 0.10. The estimated TaMEc in R
m-LaH4 is less than half that of LaH10 but slightly higher than that in La4H23, and more importantly, it could be achieved at significantly lower pressures than in these cases. The critical temperature decreases significantly with pressure, with TaMEc dropping to 33 K at 60 GPa and 14 K at 100 GPa. Results for different μ* values and various approaches for calculating the critical temperature are summarized in Table 2.
 |
| | Fig. 6 (a) Histograms of the energy-dependent distribution of the superconducting gap Δnk as a function of temperature in LaH4 at 25 and 60 GPa with μ* = 0.10 (black) and μ* = 0.15 (blue). The solid lines are guides to the eye. (b) and (c) Momentum-resolved superconducting gap on the Fermi surface in LaH4 at 25 and 60 GPa with μ* = 0.10, generated with FermiSurfer.67 | |
Table 2 Properties of LaH4 at 25, 60, and 100 GPa: density of states at the Fermi level (NF), logarithmic average phonon frequency (ωlog), total electron–phonon coupling strength (λ), semiempirical Coulomb parameter (μ*), and superconducting critical temperature (Tc). The Tc values are calculated using different methodologies: McMillan equation (TADc), Allen–Dynes formula (TADc), machine-learned SISSO mode (TMLc), isotropic Migdal–Eliashberg formalism (TiMEc), and anisotropic Migdal–Eliashberg formalism (TaMEc)
| Pressure (GPa) |
N
F (states per eV f.u.−1) |
ω
log (meV) |
λ
|
μ* |
T
McMc (K) |
T
ADc (K) |
T
MLc (K) |
T
iMEc (K) |
T
aMEc (K) |
| 25 |
0.81 |
50.6 |
1.64 |
0.10 |
72.5 |
85.4 |
102.7 |
117 |
115 |
| 0.15 |
62.8 |
71.5 |
85.1 |
103 |
102 |
| 0.20 |
53.1 |
59.1 |
68.7 |
93 |
92 |
| 60 |
0.54 |
48.7 |
0.79 |
0.10 |
26.0 |
27.4 |
25.3 |
29 |
34 |
| 0.15 |
17.8 |
18.6 |
16.2 |
21 |
24 |
| 0.20 |
10.9 |
11.2 |
10.3 |
16 |
15 |
| 100 |
0.44 |
56.1 |
0.53 |
0.10 |
9.8 |
10.1 |
7.3 |
9 |
11 |
| 0.15 |
4.4 |
4.5 |
3.0 |
9 |
9 |
| 0.20 |
1.3 |
1.3 |
0.9 |
— |
6 |
As mentioned above, the rapid decrease in Tc with increasing pressure can be directly attributed to a reduction in the DOS at the Fermi level and the hardening of the phonon frequencies. Delving deeper, superconducting hydrides are designed to provide a lattice that promotes the dissociation of molecular H2, facilitating the formation of metallic hydrogen under pressure.13 Consequently, a reduction in the H–H distance within the H2 units may suppress superconductivity, aligning with our observations. As summarized in Table 1, the highest Tc of 115 K is achieved at 25 GPa, where the shortest H–H distance is dH−H = 0.96 Å. However, Tc drops sharply as pressure increases, reaching just 11 K at 100 GPa where dH−H = 0.88 Å. For reference, the bond length of a H2 molecule is approximately 0.74 Å,76 while the H–H distance in metallic hydrogen is predicted to be around 1 Å at 500 GPa.65 This trend, along with the increased charge depletion between the H2 units at higher pressures, suggests that hydrogen may transition toward a molecular H2-like state rather than a metallic phase under compression.
4 Conclusion
We carried out an ab initio study to explore the superconducting potential of a new low-pressure rhombohedral LaH4 phase. Our results show that the inclusion of anharmonic corrections is crucial for stabilizing the crystal structure, highlighting the importance of these effects when investigating lanthanum hydrides under pressure. Furthermore, we find that the superconducting temperature can reach up to 115 K at 25 GPa but decreases rapidly with increasing pressure, a common trend in hydride superconductors. Besides experimentally confirming superconductivity of the rhombohedral phase in the 25–100 GPa range, it would be interesting to explore whether this phase could be stabilized down to ambient pressure. These results are exciting as they contribute to the ongoing search for hydrogen-rich compounds capable of achieve high-temperature superconductivity at lower pressures.
Conflicts of interest
There are no conflicts to declare.
Data availability
The data supporting this article is included in the manuscript.
Supplementary information is available. See DOI: https://doi.org/10.1039/d5tc02600h.
Acknowledgements
We would like to thank Dr. Vitalie Stavila for useful discussions. C. R. and E. R. M. acknowledge support from the National Science Foundation (NSF) under award no. DMR-2035518. This work used the Frontera supercomputer at the Texas Advanced Computing Center via the Leadership Resource Allocation (LRAC) award DMR22004 and the Expanse system at the San Diego Supercomputer Center through the ACCESS allocation TG-DMR180071. M. Z. acknowledges funding by the European Union (project ULTRA-2DPK/HORIZON-MSCA-2022-PF-01/grant agreement no. 101106654) and computational resources from the EuroHPC Joint Undertaking and supercomputer LUMI https://lumi-supercomputer.eu/, hosted by CSC (Finland) and the LUMI consortium through a EuroHPC Extreme Scale Access call. Sandia National Laboratories is a multimission laboratory managed and operated by National Technology and Engineering Solutions of Sandia, LLC, a wholly owned subsidiary of Honeywell International, Inc., for the U.S. Department of Energy (DOE)'s National Nuclear Security Administration under contract no. DE-NA-0003525. The views expressed in the article do not necessarily represent the views of the U.S. DOE or the U.S. Government. The XRD work was performed on APS beam time award from the Advanced Photon Source, a U.S. DOE Office of Science user facility operated for the DOE Office of Science by Argonne National Laboratory under contract no. DE-AC02-06CH11357. HPCAT operations are supported by DOE-NNSA's Office of Experimental Sciences.
References
- A. P. Drozdov, M. I. Eremets, I. A. Troyan, V. Ksenofontov and S. I. Shylin, Conventional superconductivity at 203 kelvin at high pressures in the sulfur hydride system, Nature, 2015, 525, 73 CrossRef CAS.
- M. Somayazulu, M. Ahart, A. K. Mishra, Z. M. Geballe, M. Baldini, Y. Meng, V. V. Struzhkin and R. J. Hemley, Evidence for Superconductivity above 260 K in Lanthanum Superhydride at Megabar Pressures, Phys. Rev. Lett., 2019, 122, 027001 CrossRef CAS.
- A. P. Drozdov, P. P. Kong, V. S. Minkov, S. P. Besedin, M. A. Kuzovnikov, S. Mozaffari, L. Balicas, F. F. Balakirev, D. E. Graf, V. B. Prakapenka, E. Greenberg, D. A. Knyazev, M. Tkacz and M. I. Eremets, Superconductivity at 250 K in lanthanum hydride under high pressures, Nature, 2019, 569, 528 CrossRef CAS PubMed.
- P. P. Kong, V. S. Minkov, M. A. Kuzovnikov, S. P. Besedin, A. P. Drozdov, S. Mozaffari, L. Balicas, F. F. Balakirev, V. B. Prakapenka, E. Greenberg, D. A. Knyazev and M. I. Eremets, Superconductivity up to 243 K in yttrium hydrides under high pressure, Nat. Commun., 2021, 12, 5075 CrossRef CAS PubMed.
- A. V. Troyan, A. O. Lyakhov, A. O. Kurnosov, I. A. Abrikosov, A. O. Shorikov, A. O. Ivanovskii, A. V. Baranov and A. V. Chulkov, Anomalous High-Temperature Superconductivity in YH6, Adv. Mater., 2021, 33, 2006832 CrossRef.
- Z. Li, X. He, C. Zhang, X. Wang, S. Zhang, Y. Jia, S. Feng, K. Lu, J. Zhao, J. Zhang, B. Min, Y. Long, R. Yu, L. Wang, M. Ye, Z. Zhang, V. Prakapenka, S. Chariton, P. A. Ginsberg, J. Bass, S. Yuan, H. Liu and C. Jin, Superconductivity above 200 K discovered in superhydrides of calcium, Nat. Commun., 2022, 13, 2863 CrossRef CAS PubMed.
- L. Ma, K. Wang, Y. Xie, X. Yang, Y. Wang, M. Zhou, H. Liu, X. Yu, Y. Zhao, H. Wang, G. Liu and Y. Ma, High-Temperature Superconducting Phase in Clathrate Calcium Hydride CaH6 up to 215 K at a Pressure of 172 GPa, Phys. Rev. Lett., 2022, 128, 167001 CrossRef CAS.
- N. P. Salke, M. M. Davari Esfahani, Y. Zhang, I. A. Kruglov, J. Zhou, Y. Wang, E. Greenberg, V. B. Prakapenka, J. Liu, A. R. Oganov and J.-F. Lin, Synthesis of clathrate cerium superhydride CeH9 at 80–100 GPa with atomic hydrogen sublattice, Nat. Commun., 2019, 10, 4453 CrossRef.
- W. Chen, D. V. Semenok, X. Huang, H. Shu, X. Li, D. Duan, T. Cui and A. R. Oganov, High-Temperature Superconducting Phases in Cerium Superhydride with a Tc up to 115 K below a Pressure of 1 Megabar, Phys. Rev. Lett., 2021, 127, 117001 CrossRef CAS PubMed.
-
Z.-Y. Cao, S. Choi, L.-C. Chen, P. Dalladay-Simpson, H. Jang, F. A. Gorelli, J.-F. Yan, S.-G. Jung, G. Huang, L. Yu, Y. Lee, J. Kim, T. Park and X.-J. Chen, Probing superconducting gap in CeH9 under pressure, arXiv, 2024, preprint, arXiv:2401.12682 cond-mat.supr-con DOI:10.48550/arXiv.2401.1268.
- D. V. Semenok, A. G. Kvashnin, A. G. Ivanova, V. Svitlyk, V. Y. Fominski, A. V. Sadakov, O. A. Sobolevskiy, V. M. Pudalov, I. A. Troyan and A. R. Oganov, Superconductivity at 161 K in thorium hydride ThH10: Synthesis and properties, Mater. Today, 2020, 33, 36 CrossRef CAS.
- I. A. Wrona, P. Niegodajew and A. P. Durajski, A recipe for an effective selection of promising candidates for high-temperature superconductors among binary hydrides, Mater. Today Phys., 2024, 46, 101499 CrossRef CAS.
- N. W. Ashcroft, Hydrogen Dominant Metallic Alloys: High Temperature Superconductors?, Phys. Rev. Lett., 2004, 92, 187002 CrossRef CAS PubMed.
- Z. M. Geballe, H. Liu, A. K. Mishra, M. Ahart, M. Somayazulu, Y. Meng, M. Baldini and R. J. Hemley, Synthesis and stability of lanthanum superhydrides, Angew. Chem., Int. Ed., 2018, 57, 688 CrossRef CAS.
- H. Liu, I. I. Naumov, R. Hoffmann, N. W. Ashcroft and R. J. Hemley, Potential high-Tc superconducting lanthanum and yttrium hydrides at high pressure, Proc. Natl. Acad. Sci. U. S. A., 2017, 114, 6990 Search PubMed.
- H. Liu, I. I. Naumov, Z. M. Geballe, M. Somayazulu, J. S. Tse and R. J. Hemley, Dynamics and superconductivity in compressed lanthanum superhydride, Phys. Rev. B, 2018, 98, 100102 Search PubMed.
- D. Sun, V. S. Minkov, S. Mozaffari, Y. Sun, Y. Ma, S. Chariton, V. B. Prakapenka, M. I. Eremets, L. Balicas and F. F. Balakirev, High-temperature superconductivity on the verge of a structural instability in lanthanum superhydride, Nat. Commun., 2021, 12, 6863 CrossRef CAS PubMed.
- I. Errea, F. Belli, L. Monacelli, A. Sanna, T. Koretsune, T. Tadano, R. Bianco, M. Calandra, R. Arita, F. Mauri and J. A. Flores-Livas, Quantum crystal structure in the 250-kelvin superconducting lanthanum hydride, Nature, 2020, 578, 66 CrossRef CAS PubMed.
- C. Wang, S. Yi and J.-H. Cho, Multiband nature of room-temperature superconductivity in LaH10 at high pressure, Phys. Rev. B, 2020, 101, 104506 CrossRef CAS.
- Y. L. Wu, X. H. Yu, J. Z. L. Hasaien, F. Hong, P. F. Shan, Z. Y. Tian, Y. N. Zhai, J. P. Hu, J. G. Cheng and J. Zhao, Ultrafast dynamics evidence of strong coupling superconductivity in LaH10 ± δ, Nat. Commun., 2024, 15, 9683 CrossRef CAS PubMed.
- Z. Wu, Y. Sun, A. P. Durajski, F. Zheng, V. Antropov, K.-M. Ho and S. Wu, Effect of doping on the phase stability and superconductivity in LaH10, Phys. Rev. Mater., 2023, 7, L101801 CrossRef CAS.
- M. Caussé, G. Geneste and P. Loubeyre, Superionicity of Hδ− in LaH10 superhydride, Phys. Rev. B, 2023, 107, L060301 CrossRef.
- K. K. Ly and D. M. Ceperley, Stability and distortion of fcc LaH10 with path-integral molecular dynamics, Phys. Rev. B, 2022, 106, 054106 CrossRef CAS.
- Y. Watanabe, T. Nomoto and R. Arita, Quantum and temperature effects on the crystal structure of superhydride LaH10: A path integral molecular dynamics study, Phys. Rev. B, 2022, 105, 174111 CrossRef CAS.
- X. Liang, A. Bergara, X. Wei, X. Song, L. Wang, R. Sun, H. Liu, R. J. Hemley, L. Wang, G. Gao and Y. Tian, Prediction of high-Tc superconductivity in ternary lanthanum borohydrides, Phys. Rev. B, 2021, 104, 134501 CrossRef CAS.
- A. K. Verma, P. Modak, F. Schrodi, A. Aperis and P. M. Oppeneer, Prediction of an unusual trigonal phase of superconducting LaH10 stable at high pressures, Phys. Rev. B, 2021, 104, 174506 CrossRef CAS.
- Y. Ge, F. Zhang and R. J. Hemley, Room-temperature superconductivity in boron- and nitrogen-doped lanthanum superhydride, Phys. Rev. B, 2021, 104, 214505 CrossRef CAS.
- D. Laniel, F. Trybel, B. Winkler, F. Knoop, T. Fedotenko, S. Khandarkhaeva, A. Aslandukova, T. Meier, S. Chariton, K. Glazyrin, V. Milman, V. Prakapenka, I. A. Abrikosov, L. Dubrovinsky and N. Dubrovinskaia, High-pressure synthesis of seven lanthanum hydrides with a significant variability of hydrogen content, Nat. Commun., 2022, 13, 6987 CrossRef CAS PubMed.
- J. Guo, D. Semenok, G. Shutov, D. Zhou, S. Chen, Y. Wang, K. Zhang, X. Wu, S. Luther, T. Helm, X. Huang and T. Cui, Unusual metallic state in superconducting A15-type La4H23, Natl. Sci. Rev., 2024, 11, nwae149 CrossRef CAS PubMed.
- S. Cross, J. Buhot, A. Brooks, W. Thomas, A. Kleppe, O. Lord and S. Friedemann, High-temperature superconductivity in La4H23 below 100 GPa, Phys. Rev. B, 2024, 109, L020503 CrossRef CAS.
-
S. Duwal, V. Stavila, C. Spataru, M. Shivanna, P. Allen, T. Elmslie, C. T. Seagle, J. Jeffries, N. Velisavljevic, J. Smith, P. Chow, Y. Xiao, M. Somayazulu and P. A. Sharma, Enhancement of hydrogen absorption and hypervalent metal hydride formation in lanthanum using cryogenic ball milling, arXiv, 2025, preprint, arXiv:2506.23980 DOI:10.48550/arXiv.2506.23980.
- P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni, I. Dabo, A. D. Corso, S. de Gironcoli, S. Fabris, G. Fratesi, R. Gebauer, U. Gerstmann, C. Gougoussis, A. Kokalj, M. Lazzeri, L. Martin-Samos, N. Marzari, F. auri, R. Mazzarello, S. Paolini, A. Pasquarello, L. Paulatto, C. Sbraccia, S. Scandolo, G. Sclauzero, A. P. Seitsonen, A. Smogunov, P. Umari and R. M. Wentzcovitch, QUANTUM ESPRESSO: a modular and opensource software project for quantum simulations of materials, J. Phys.: Condens. Matter, 2009, 21, 395502 Search PubMed.
- P. Giannozzi, O. Andreussi, T. Brumme, O. Bunau, M. B. Nardelli, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, M. Cococcioni, N. Colonna, I. Carnimeo, A. D. Corso, S. de Gironcoli, P. Delugas, R. A. DiStasio, A. Ferretti, A. Floris, G. Fratesi, G. Fugallo, R. Gebauer, U. Gerstmann, F. Giustino, T. Gorni, J. Jia, M. Kawamura, H.-Y. Ko, A. Kokalj, E. Küçükbenli, M. Lazzeri, M. Marsili, N. Marzari, F. Mauri, N. L. Nguyen, H.-V. Nguyen, A. O. de-la Roza, L. Paulatto, S. Poncé, D. Rocca, R. Sabatini, B. Santra, M. Schlipf, A. P. Seitsonen, A. Smogunov, I. Timrov, T. Thonhauser, P. Umari, N. Vast, X. Wu and S. Baroni, Advanced capabilities for materials modelling with Quantum ESPRESSO, J. Phys.: Condens. Matter, 2017, 29, 465901 CrossRef CAS.
- D. R. Hamann, Optimized norm-conserving Vanderbilt pseudopotentials, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 88, 085117 CrossRef.
- M. van Setten, M. Giantomassi, E. Bousquet, M. Verstraete, D. Hamann, X. Gonze and G.-M. Rignanese, The PseudoDojo: Training and grading a 85 element optimized norm-conserving pseudopotential table, Comput. Phys. Commun., 2018, 226, 39 CrossRef CAS.
- J. P. Perdew, A. Ruzsinszky, G. I. Csonka, O. A. Vydrov, G. E. Scuseria, L. A. Constantin, X. Zhou and K. Burke, Restoring the Density-Gradient Expansion for Exchange in Solids and Surfaces, Phys. Rev. Lett., 2008, 100, 136406 Search PubMed.
- H. J. Monkhorst and J. D. Pack, Special points for Brillouin-zone integrations, Phys. Rev. B, 1976, 13, 5188 CrossRef.
- M. Methfessel and A. T. Paxton, High-precision sampling for Brillouin-zone integration in metals, Phys. Rev. B: Condens. Matter Mater. Phys., 1989, 40, 3616 CrossRef CAS PubMed.
- S. Baroni, S. de Gironcoli, A. Dal Corso and P. Giannozzi, Phonons and related crystal properties from density-functional perturbation theory, Rev. Mod. Phys., 2001, 73, 515 CrossRef CAS.
- M. Zacharias and F. Giustino, One-shot calculation of temperature-dependent optical spectra and phononinduced band-gap renormalization, Phys. Rev. B, 2016, 94, 075125 CrossRef.
- M. Zacharias and F. Giustino, Theory of the special displacement method for electronic structure calculations at finite temperature, Phys. Rev. Res., 2020, 2, 013357 CrossRef CAS.
- M. Zacharias, G. Volonakis, F. Giustino and J. Even, Anharmonic lattice dynamics via the special displacement method, Phys. Rev. B, 2023, 108, 035155 CrossRef CAS.
- N. R. Werthamer, Self-consistent phonon formulation of anharmonic lattice dynamics, Phys. Rev. B, 1970, 1, 572 CrossRef.
- H. Lee, S. Poncé, K. Bushick, S. Hajinazar, J. Lafuente-Bartolome, J. Leveillee, C. Lian, J.-M. Lihm, F. Macheda, H. Mori, H. Paudyal, W. H. Sio, S. Tiwari, M. Zacharias, X. Zhang, N. Bonini, E. Kioupakis, E. R. Margine and F. Giustino, Electron–phonon physics from first principles using the EPW code, npj Comput. Mater., 2023, 9, 156 CrossRef CAS.
-
S. B. Mishra, H. Mori and E. R. Margine, Electronphonon vertex correction effect in superconducting H3S, arXiv, 2025, preprint, arXiv:2507.01897 cond-mat.supr-con DOI:10.48550/arXiv.2507.01897.
- See Fig. S1–S3.
- K. Momma and F. Izumi, VESTA3 for three-dimensional visualization of crystal, volumetric and morphology data, J. Appl. Crystallogr., 2011, 44, 1272 CrossRef CAS.
- F. Giustino, M. L. Cohen and S. G. Louie, Electronphonon interaction using Wannier functions, Phys. Rev. B, 2007, 76, 165108 CrossRef.
- E. R. Margine and F. Giustino, Anisotropic Migdal–Eliashberg theory using Wannier functions, Phys. Rev. B, 2013, 87, 024505 CrossRef.
- S. Poncé, E. Margine, C. Verdi and F. Giustino, EPW: Electron–phonon coupling, transport and superconducting properties using maximally localizedWannier functions, Comput. Phys. Commun., 2016, 209, 116 CrossRef.
- N. Marzari, A. A. Mostofi, J. R. Yates, I. Souza and D. Vanderbilt, Maximally localized Wannier functions: Theory and applications, Rev. Mod. Phys., 2012, 84, 1419 CrossRef CAS.
- G. Pizzi, V. Vitale, R. Arita, S. Blügel, F. Freimuth, G. Géranton, M. Gibertini, D. Gresch, C. Johnson, T. Koretsune, J. Ibãnez-Azpiroz, H. Lee, J.-M. Lihm, D. Marchand, A. Marrazzo, Y. Mokrousov, J. I. Mustafa, Y. Nohara, Y. Nomura, L. Paulatto, S. Poncé, T. Ponweiser, J. Qiao, F. Thöle, S. S. Tsirkin, M. Wierzbowska, N. Marzari, D. Vanderbilt, I. Souza, A. A. Mostofi and J. R. Yates, Wannier90 as a community code: new features and applications, J. Phys.: Condens. Matter, 2020, 32, 165902 CrossRef CAS PubMed.
- R. Lucrezi, P. P. Ferreira, S. Hajinazar, H. Mori, H. Paudyal, E. R. Margine and C. Heil, Full-bandwidth anisotropic Migdal–Eliashberg theory and its application to superhydrides, Commun. Phys., 2024, 7, 33 CrossRef CAS.
- H. Mori, T. Nomoto, R. Arita and E. R. Margine, Efficient anisotropic Migdal–Eliashberg calculations with an intermediate representation basis and Wannier interpolation, Phys. Rev. B, 2024, 110, 064505 CrossRef CAS.
- V. Stavila, S. Li, C. Dun, M. A. T. Marple, H. E. Mason, J. L. Snider, J. E. r Reynolds, F. El Gabaly, J. D. Sugar, C. D. Spataru, X. Zhou, B. Dizdar, E. H. Majzoub, R. Chatterjee, J. Yano, H. Schlöombeg, B. V. Lotsch, J. J. Urban, B. C. Wood and M. D. Allendorf, Defying Thermodynamics: Stabilization of Alane Within Covalent Triazine Frameworks for Reversible Hydrogen Storage, Angew. Chem., Int. Ed., 2021, 60, 25815 CrossRef CAS.
- F. Peng, Y. Sun, C. J. Pickard, R. J. Needs, Q. Wu and Y. Ma, Hydrogen Clathrate Structures in Rare Earth Hydrides at High Pressures: Possible Route to Room-Temperature Superconductivity, Phys. Rev. Lett., 2017, 119, 107001 CrossRef.
- P. Kong, V. S. Minkov, M. A. Kuzovnikov, A. P. Drozdov, S. P. Besedin, S. Mozaffari, L. Balicas, F. F. Balakirev, V. B. Prakapenka, S. Chariton, D. A. Knyazev, E. Greenberg and M. I. Eremets, Superconductivity up to 243 K in the yttrium-hydrogen system under high pressure, Nat. Commun., 2021, 12, 5075 CrossRef CAS.
- Y. Li, J. Hao, H. Liu, J. S. Tse, Y. Wang and Y. Ma, Pressure-stabilized superconductive yttrium hydrides, Sci. Rep., 2015, 5, 9948 CrossRef CAS PubMed.
- X. Feng, J. Zhang, G. Gao, H. Liu and H. Wang, Compressed sodalite-like mgh6 as a potential hightemperature superconductor, RSC Adv., 2015, 5, 59292 RSC.
- J. Zhao, B. Ao, S. Li, T. Gao and X. Ye, Phase Diagram and Bonding States of Pu–H Binary Compounds at High Pressures, J. Phys. Chem. C, 2020, 124, 7361 CrossRef CAS.
- Y.-L. Hai, N. Lu, H.-L. Tian, M.-J. Jiang, W. Yang, W.-J. Li, X.-W. Yan, C. Zhang, X.-J. Chen and G.-H. Zhong, Cage Structure and Near Room-Temperature Superconductivity in TbHn (n = 1–12), The, J. Phys. Chem. C, 2021, 125, 3640 CrossRef CAS.
- D. V. Semenok, D. Zhou, A. G. Kvashnin, X. Huang, M. Galasso, I. A. Kruglov, A. G. Ivanova, A. G. Gavriliuk, W. Chen, N. V. Tkachenko, A. I. Boldyrev, I. Troyan, A. R. Oganov and T. Cui, Novel Strongly Correlated
Europium Superhydrides, The, J. Phys. Chem. Lett., 2021, 12, 32 CrossRef CAS PubMed.
- W. Sun, X. Kuang, H. D. J. Keen, C. Lu and A. Hermann, Second group of high-pressure high-temperature lanthanide polyhydride superconductors, Phys. Rev. B, 2020, 102, 144524 CrossRef CAS.
- X. Zhong, H. Wang, J. Zhang, H. Liu, S. Zhang, H.-F. Song, G. Yang, L. Zhang and Y. Ma, Tellurium hydrides at high pressures: High-temperature superconductors, Phys. Rev. Lett., 2016, 116, 057002 CrossRef PubMed.
- N. W. Ashcroft, Metallic Hydrogen: A High-Temperature Superconductor?, Phys. Rev. Lett., 1968, 21, 1748 Search PubMed.
- H. W. T. Morgan and A. N. Alexandrova, Structures of LaH10, EuH9, and UH8 superhydrides rationalized by electron counting and Jahn–Teller distortions in a covalent cluster model, Chem. Sci., 2023, 14, 6679 RSC.
- M. Kawamura, Fermisurfer: Fermi-surface viewer providing multiple representation schemes, Comput. Phys. Commun., 2019, 239, 197 Search PubMed.
- I. Errea, M. Calandra and F. Mauri, First-principles theory of anharmonicity and the inverse isotope effect in superconducting palladium-hydride compounds, Phys. Rev. Lett., 2013, 111, 177002 CrossRef PubMed.
- I. Errea, M. Calandra, C. J. Pickard, J. Nelson, R. J. Needs, Y. Li, H. Liu, Y. Zhang, Y. Ma and F. Mauri, High-pressure hydrogen sulfide from first principles: A strongly anharmonic phonon-mediated superconductor, Phys. Rev. Lett., 2015, 114, 157004 CrossRef PubMed.
- P. Hou, F. Belli, R. Bianco and I. Errea, Quantum anharmonic enhancement of superconductivity in P63/mmc ScH6 at high pressures: A first-principles study, J. Appl. Phys., 2021, 130, 175902 Search PubMed.
- M. Dogan, S. Oh and M. L. Cohen, High temperature superconductivity in the candidate phases of solid hydrogen, J. Phys.: Condens. Matter, 2022, 34, 15LT01 Search PubMed.
- J. J. Gilman, Lithium Dihydrogen Fluoride—An Approach to Metallic Hydrogen, Phys. Rev. Lett., 1971, 26, 546 Search PubMed.
- W. L. McMillan, Transition temperature of strongcoupled superconductors, Phys. Rev., 1968, 167, 331 Search PubMed.
- P. B. Allen and R. C. Dynes, Transition temperature of strong-coupled superconductors reanalyzed, Phys. Rev. B, 1975, 12, 905 Search PubMed.
- S. R. Xie, G. R. Stewart, J. J. Hamlin, P. J. Hirschfeld and R. G. Hennig, Functional form of the superconducting critical temperature from machine learning, Phys. Rev. B, 2019, 100, 174513 Search PubMed.
-
K. P. Huber and G. Herzberg, Constants of diatomic molecules, Molecular Spectra and Molecular Structure: IV. Constants of Diatomic Molecules, Springer, US, Boston, MA, 1979, pp. 8–689 Search PubMed.
|
| This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.