Potential high- superconductivity in YCeHx and LaCeHx under pressure
Abstract
Lanthanum, yttrium, and cerium hydrides are the three most well-known superconducting binary hydrides (La-H, Y-H, and Ce-H systems), which have gained great attention in both theoretical and experimental studies. Recent studies have shown that ternary hydrides composed of lanthanum and yttrium can achieve high superconductivity around 253 K. In this study we employ the evolutionary-algorithm-based crystal structure prediction (CSP) method and first-principles calculations to investigate the stability and superconductivity of ternary hydrides composed of (Y, Ce) and (La, Ce) under high pressure. Our calculations show that there are multiple stable phases in Y-Ce-H and La-Ce-H systems, among which -YCeH8, -YCeH18, -YCeH20, -LaCeH8, and -LaCeH20 possessing H18, H29 and H32 clathrate structures can maintain both the thermodynamic and lattice-dynamic stabilities. In addition, we also find that these phases also maintain a strong resistance to decomposition at high temperature. Electron-phonon coupling calculations show that only three of these five phases can exhibit high-temperature superconductivity. The superconducting transition temperatures () of -YCeH20, -LaCeH20, and -YCeH18 are predicted using the Allen-Dynes-modified McMillan formula to be 122 K at 300 GPa, 116 K at 250 GPa, and 173 K at 150 GPa, respectively. Moreover, the pressure to stabilize -YCeH18 can be lowered to 150 GPa, suggesting an accessible condition for its high-pressure synthesis.
I Introduction
Superconductivity has gotten a remarkable progress and attention owing to its unique phenomenon and wide range of industrial applications. Tinkham (2004); Gennes (2018); Linder and Robinson (2015); Josephson (1962); Bergen et al. (2019); Islam et al. (2014); Thomas et al. (2016) On the other hand, the extremely low temperature environment required for the manifestation of these properties severely limits the breadth of their application. Therefore, the pursuit for high(room)-temperature superconductivity has historically been one of the most competitive and sought goals in superconductivity. Tinkham (1988); Kalsi (2011); Flores-Livas et al. (2020) According to the Bardeen-Cooper-Schrieffer (BCS) theory, the superconducting transition temperature () significantly correlates with the phonon vibration frequency. Bardeen (1955); Cooper (1956); Bardeen et al. (1957a, b) In this sense, metallic hydrogen is considered to be the most ideal candidate for high- superconductivity due to its ultra-high phonon vibration frequency. Ashcroft (1968) However, hydrogen metallization, also known as the Wigner-Huntington transition, typically requires extremely high pressures, making it difficult to study its superconductivity-related electrical properties at these pressures. Wigner and Huntington (1935); Loubeyre et al. (1996); McMinis et al. (2015); Eremets and Troyan (2011); Dalladay-Simpson et al. (2016); Dias and Silvera (2017); Castelvecchi (2017); Celliers et al. (2018) Ashcroft proposed the introduction of other elements into hydrogen can provide the necessary pre-compression for the entire system, allowing it to maintain metallicity and superconductivity at lower pressures. Ashcroft (2004) Thus, metal hydrides are one of the most ideal candidates for high- superconductivity.
According to Ashcroft, almost all binary hydrides have been theoretically screened out by density functional theory (DFT) calculations, and some predictions have been verified by experiments. Semenok et al. (2020); Peng et al. (2017); Zhang et al. (2020); Li and Peng (2017); Yu et al. (2015); Einaga et al. (2016); Drozdov et al. (2015, 2019); Salke et al. (2019); Chen et al. (2021a); Shao et al. (2021a, b); Troyan et al. (2021); Chen et al. (2021b); Kong et al. (2021); Li et al. (2014); Somayazulu et al. (2019) Compared to binary superconducting hydrides, research on multi-component superconducting hydrides appears to be slightly less. Since multi-component hydrides contain more combinations of elements, however, they are more promising in the search for high(room)-temperature superconductivity. Song et al. (2021); Ma et al. (2017a, b); Shao et al. (2019); Zheng et al. (2021); Rahm et al. (2017); Liang et al. (2021); Sun et al. (2019); Liang et al. (2019a); Wei et al. (2020); Shi et al. (2021); Song et al. (2022a); Semenok et al. (2021); Liang et al. (2019b); Jiang et al. (2021); Cui et al. (2020); Cataldo et al. (2020); Song et al. (2022b); Cataldo et al. (2022); Vocaturo et al. (2022); Somayazulu et al. (2019) For example, the DFT calculations by Sun et al. Sun et al. (2019) predicted that magnesium hydride can dissociate the H2 molecule into atomic H by lithium doping. This doping increases its electron concentration at the Fermi level (), enabling it to reach a value of 473 K at 250 GPa. Sun et al. (2019) Furthermore, a recent high-pressure experiment has revealed that a small amount of carbon doping in H3S can result in room-temperature superconductivity ( = 287.7 1.2 K at 267 10 GPa), Snider et al. (2020) although the anomaly of AC susceptibility in low temperature for the carbonaceous sulfur hydride has been questioned by Hirsch et al. Hirsch and Marsiglio (2021); Hirsch (2022) and more strong experimental data are needed to support the claim of the observation of room-temperature superconductivity. Therefore, increasing the density of electronic states at by electrons doping in the binary hydride can affect the value; however this modification also can result in a decrease in in some cases. Nakanishi et al. (2018); Amsler (2019); Guan et al. (2021)
Furthermore, most of the high- superconducting hydrides reported thus far are concentrated in the Mendeleev table’s ”lability belt”, and many of these hydrides have unusual structural properties, such as a cage-like structure surrounded by H. Semenok et al. (2020) Some ternary superconducting hydrides (e.g., ScYH6, ScCaH8, CaYH12, and so on) can be easily obtained by substituting one of the metal atoms in the binary parent compound. Liang et al. (2019a); Wei et al. (2020); Shi et al. (2021); Song et al. (2022a); Semenok et al. (2021) The crystal structure, pressure range of stability, unit cell volume, and other properties of the parent binary superconducting hydrides that correspond to these ternary superconducting hydrides resemble each other. Semenok et al. (2020); Flores-Livas et al. (2020) For instance, by combining two types of binary hydrides with comparable structural features, it is possible to produced a more stable ternary hydride with similar structural properties. To test this hypothesis, we noticed that CeH9, YH9, and LaH10, CeH10 have comparable qualities in a variety of properties, and their structural properties have been proven in high-pressure experiments. Salke et al. (2019); Chen et al. (2021b); Kong et al. (2021); Drozdov et al. (2019) The related ternary superconducting hydrides, one of which is (La,Y)H10, has been successfully synthesized experimentally, and the observed under high pressure also matched the theoretically predicted quite well. Semenok et al. (2021) During this work, we are aware that a very recent high-pressure experiment study shows that La-Ce-H may have stable -(La,Ce)H4 and -(La,Ce)H6 around 100 GPa. Chen et al. (2022a)
Therefore, this work will concentrate on the YCeHx and LaCeHx systems, which have not yet been well investigated, to explore whether there exist stable compounds under high pressure. By employing the evolutionary algorithm for crystal structure prediction, we have theoretically discovered thermodynamically stable compounds of YCeH5,YCeH7,YCeH8, YCeH18, YCeH20, LaCeH8, LaCeH18, and LaCeH20 in pressure range of 100-400 GPa. In addition, we have investigated the dynamic stability and superconductivity of these hydrogen-rich ternary hydrides within the harmonic approximation. All of them are dynamically stable with increasing pressure, More importantly, YCeH18 exhibits extremely strong electron-phonon coupling with superconducting transition temperatures of 173 K at 150 GPa.
II Method
The crystal-structure search for YCeHx and LaCeHx ( = 210, 12, 14, 16, 18, and 20) at 100, 200, and 300 GPa was performed using the USPEX (Universal Structure Predictor: Evolutionary Xtallography) Glass et al. (2006); Lyakhov et al. (2013) code. In the USPEX run, 200 structures with 1-2 formula units were randomly built for the initial generation, and then 100 structures for each of the subsequent generations were produced by 40% heredity, 40% random, 10% mutation, and 10% soft mutation. Each structure was optimized through 4 times of relaxation from low to high accuracy levels. The structure relaxation was carried out by the density functional theory calculations within the generalized gradient approximation (GGA) with Perdew-Burke-Ernzerhof (PBE) exchange-correlation functional Perdew et al. (1996) as implemented in the VASP (Vienna ab initio simulation package) Kresse and Furthmüller (1996); Kresse and Joubert (1999) code. The electron-ion interaction was treated using projector-augmented-wave (PAW) potentials. Kresse and Furthmüller (1996); Kresse and Joubert (1999) The Ce 4f electrons were explicitly considered into the valence electrons. The cutoff energy for plane waves was set to 600 eV and the smallest allowed spacing between the -points in the irreducible Brillouin zone was set to be 0.2 . To study the chemical bonding and electronic structures of some selected stable YCeHx and LaCeHx, we have carried out the static calculations using the VASP code and then performed the Crystal Orbital Hamilton Population (COHP) and Crystal Orbital Bond Index (COBI) analyses using the Lobster 4.1.0 package Müller et al. (2021) together with the pbeVaspFit2015 basis set. The electronic band structures were also checked by taking into account the spin-orbit coupling (SOC) interaction and the correction of the on-site Coulomb interaction of Ce 4f orbitals. The DFT+U method of Dudarev et al. Dudarev et al. (1998) was employed in the latter case. The Hubbard U parameter for Ce 4f orbitals was determined by the density-functional perturbation theory approach, as implemented in the HP package Timrov et al. (2022). The obtained U for Ce 4f orbitals is 4.2 eV, which is very close to the value (i.e., 4.5 eV) adopted by Wang et al. Wang et al. (2021) for their study on CeH9. The exchange interaction parameter for Ce 4f orbitals was taken from Ref. 77 and set to 0.5 eV.
The formation enthalpies of La-Ce-H and Y-Ce-H ternary systems (herein abbreviated to -Ce-Y with = La and Y), relative to the elemental solids Y Chen et al. (2012), Ce Vohra et al. (1999), La Chen et al. (2022b), and H Pickard and Needs (2007), were calculated at each pressure defined as below:
(1) |
where is the enthalpy per formula unit (f.u.) of the A-Ce-H compound; , , and are the reference enthalpies of , Ce, and H single phases, respectively. The thermodynamic stability of the predicted phase was determined by comparing the formation enthalpies with respect to the convex hull energy:
(2) |
where is a convex hull energy obtained by constraining the minimum value of the total enthalpies of a linear combination of stable phases Akbarzadeh et al. (2007), which can be computed using the ConvexHull module in scipy. Virtanen et al. (2020) is energy above the convex hull. means that the corresponding ternary phase is stable, namely, such a phase would not decompose into any combination of elementary, binary, or other ternary phases. The Gibbs free energies at finite temperatures were calculated within the quasiharmonic approximation using the PHONOPY code. Togo and Tanaka (2015) After incorporating the Gibbs free energies, we can drive the stability for the whole system at high temperatures by computing the convex hull.
The phonon dispersion and electron-phonon coupling (EPC) of the stable ternary phases were carried out using the Quantum ESPRESSO (QE) code Giannozzi et al. (2009) with the PAW method and the GGA-PBE exchange-correlation functional. In particular, the Ce PAW potential was generated with a valence configuration of 4d105s25p64f0.55d1.56s2, as provided by the PSLibray of Dal Corso Corso (2014). The cutoff energy for plane-wave basis sets in the QE calculations was set to 180 Ry. The -point mesh (and integral -points mesh) in the first Brillouin-zone for the EPC calculations was set as follows: () for CeH8, () for CeH18 and () for CeH20 ( = La and Y). The the Allen-Dynes-modified McMillan formula (AD) and the Eliashberg function derived from the EPC calculation were used to predict the superconducting critical temperature as follows McMillan (1968); Allen and Dynes (1975):
(3) |
with
(4) | |||||
When = 1, Eq. (3) is restored to the original McMillan formula (McM). The is the Coulomb pseudopotential, which is defined as below Allen and Dynes (1975)
(5) |
where is instantaneous repulsion, the is a ratio of propagation time. For most metal compounds, the value of is roughly in the range of 0.1 to 0.16 by incorporating previous experimental and calculated data. The widely accepted value of 0.1 for was used herein. The electron-phonon coupling constant (), logarithmic average phonon frequency (), and mean square frequency () are defined as below
(6) |
(7) |
and
(8) |
respectively.
III Results and discussion
III.1 Phase stability and crystal structural of YCeHx and LaCeHx
Because predicting the ternary phase diagram in its entirety is computationally expensive, and our main target is to determine the stability and superconductivity of the hydrogen-rich phase, this research focuses on stable compounds and crystal structures at 100 GPa, 200 GPa, and 300 GPa along a line chosen by YCeHx and LaCeHx ( = 2-8, 10, 12, 14, 16, 18, and 20) for the fixed composition search. This approach is currently seen to be effective for alkaline earths and rare earth metal ternary hydrides. Liang et al. (2019a); Shao et al. (2021a); Cataldo et al. (2020) For the stability analysis of YCeHx and LaCeHx, the corresponding Y-H, La-H, and Ce-H binary hydrides have been extensively studied. Drozdov et al. (2019); Chen et al. (2021a); Kong et al. (2021) Theoretical calculations show that there is a hydrogen-rich structure H9 ( = Y, La, and Ce) that can be stabilized below 100 GPa in the three binary systems. In addition, YH10 is capable of exhibiting room temperature superconductivity. Peng et al. (2017) Therefore, this study mainly used these stable binary hydrides as well as single elements as a reference for thermodynamic stability assessment.
The constructed convex hulls of YCeHx and LaCeHx at 200 GPa are presented in Fig. 1. Their stable and metastable phases are highlighted by the symbols of circles and red diamond in Fig. 1, respectively. The more detailed results for the stabilites of YCeHx and LaCeHx at difference pressures are presented in Figs. S1-S6 in the Supplementary Material (SM). YCeH5, YCeH7, YCeH8, LaCeH8, and LaCeH18 are found to be stabilized at 100 GPa. When the pressure increases up to 200 GPa, YCeH5, YCeH7, and LaCeH18 vanish from the ternary convex, while LaCeH20 and YCeH18 appear in the stable phases of convex hull. In the pressure range of 100-300 GPa, YCeH20 fails to achieve stability. But further increasing the pressure above 300 GPa, -YCeH20 becomes stable, and its stable pressure interval is very close to that of YH10. Peng et al. (2017) In addition, in the pressure interval of 300-400 GPa, LaCeH8, LaCeH20, and YCeH20 undergo a phase transition to the more stable , , and phases, respectively. The ranges of the stabilization pressures of these new predicted phases have been summarized in Fig. 2. To further explore the resistance of these phases to decomposition at high temperatures, we calculated the ternary convex hull at a finite temperature at 200 GPa. We note that the stable phase (YCeH8, YCeH18, LaCeH8, and LaCeH20) at this point is also stable at high temperatures, and its specific energy values have been placed in the SM. Therefore, the above-mentioned can be synthesized in the same way as (La,Y)Hx Semenok et al. (2021) by heating LaCe and YCe alloys with NH3BH3 at high pressure.
It is interesting to note that most of the above predicted stable phases, except YCeH5 and YCeH7, are composed of clathrate structures enclosed by H. For instance, the clathrate structures of stable YCeH8, LaCeH8, YCeH18, LaCeH18, and YCeH20 are shown in Fig. 3. Since the hydrogen-rich compounds are more likely to achieve high- superconductivity Flores-Livas et al. (2020), the hydrogen-less ones such as YCeH5 and YCeH7 will be discussed elsewhere. The stable hydrides of YCeHx and LaCeHx with clathrate structures can be roughly divided into three groups.
(i) The first one consists of H18 cages, as found in -YCeH8, -LaCeH8, and -LaCeH8, which are derived from their corresponding parent compounds (-YH4, -CeH4, and -LaH4) Peng et al. (2017) with the same cage structure. If half of Ce atoms at the site of the -CeH4 phase are orderly replaced by Y or La atoms, the ternary phase would be formed. Peng et al. (2017) The relative enthalpies of these two structures of LaCeH4 and YCeH4 are shown in Figs. S1 and S4. We find that the most stable candidate structures prefer the H18 cages. In addition, LaCeH8 would be transformed into the phase when the pressure is above 300 GPa. (ii) The second group consists of H29 cages, as found in the -YCeH18 and -LaCeH18 phases. The -YCeH18 phase can be derived by replacing half of Ce atoms at the site 2 of -CeH9 with Y atoms. As for the parent compounds of -YCeH18 and -LaCeH18, the previous study by Peng et al. Peng et al. (2017) has shown that only stable -YH9 and -CeH9 phases exist in the pressure range of 100-400 GPa, however LaH9 does not. For the -LaCeH18 phase, only its internal energy plays a superior role to hinder its possible decomposition pathway of 1/6 LaH4 + 1/6 LaH10 + CeH9, and thus it cannot be stabilized at higher pressures. Liu et al. (2017); Peng et al. (2017) (iii) The third group consists of the H32 cages, as found in YCeH20 and LaCeH20. The -YCeH20 and -LaCeH20 phases exhibit the same spatial structure with LaYH20, Semenok et al. (2021) and all of them are the supercell structures of their parent phases -H10. More specifically speaking, the equivalent primitive unit cell of -YCeH20 can be obtained by replacing half of Ce atoms with Y atoms in the extension of the primitive unit cell of -CeH10. LaCeH20 undergoes a phase transition from to . For YCeH20, it also undergoes a phase transition at 370 GPa and the new high-pressure phase can be derived by replacing half of Ce atoms in the site of the parent phase -CeH10 with Y atoms. Considering the chemical similarity of elements, one may choose the stable phases of binary hydrides as the starting structures to accelerate the search for multi-component hydrides.
To further check the lattice dynamic stability of the above predicted YCeHx and LaCeHx phases, we have employed the Phonopy Togo and Tanaka (2015) code to initially calculate their phonon band structures, which are presented in Fig. S7 in the SM. Among them, the -LaCeH18 phases cannot maintain dynamic stability in the pressure range of their thermodynamic stability. To make the -LaCeH20 phase dynamically stable, the pressure needs to be above 200 GPa. In addition, the high pressure applied in the current mainstream experimental studies of metal hydrides falls into a range of 100-300 GPa. As for the phases stabilized above 300 GPa, both -YCeH20 and -LaCeH20 have a certain degree of similarity. Therefore, we pay much attention to the -YCeH8, -YCeH18, -YCeH20, -LaCeH8, and -LaCeH20 phases for their electronic properties and superconductivity.
III.2 Electronic Properties and Superconductivity
The electronic band structure, the electronic density of states (eDOS), of -YCeH8, -YCeH18, -YCeH20, -LaCeH8, and -LaCeH20 at high pressures are presented in Figs. S8-S10 in the SM, respectively. All these compounds in the pressure range of their thermodynamical stabilities possess the metallic features in their energy band structures. For -YCeH8 and -LaCeH8, there are very steep conduction bands along the direction crossing the Fermi level (), resulting in electron pockets near the point. For -YCeH20 and -LaCeH20, a band inversion can be found around the point and near the , which is mainly attributed to the H 1s and Ce 4f orbitals, as seen from the atom- and orbital-weighted band structures in Figs. S8 and S9 in the SM, respectively. The eDOS at () for -YCeH8, -YCeH18, -YCeH20, -LaCeH8, and -LaCeH20 are 1.72, 1.39, 1.10, 1.34, and 1.26 states/eV/f.u., respectively. As further analyzed by the atom-projected eDOS, the difference in the aforementioned values of mainly comes from the different contribution of H and Ce atoms to the . The contribution of Ce atoms at is continuously suppressed as the hydrogen content increases, while the opposite trend is observed for H. However, the decrease of Ce is higher than the increase of H, leading to a constant decrease of the total DOS with increasing H content. According to the arguments of Belli et al., Belli et al. (2021) the superconductivity of hydrides maintains a strong positive correlation with the contribution of H in the total DOS at (). The for these five phases are 9.5%, 34.4%, 38.2%, 11.2%, and 34.7%, respectively. This indicates that the H-rich phase has a higher H-derived DOS at although the corresponding total DOS is low, which may suggest that the possibility of high- in H-rich phases to some extent. Belli et al. (2021)
From the orbital-weighted electronic band structures of -YCeH18, -YCeH20 and, -LaCeH20, as shown in Fig. S9, it can be seen that the bands around are mainly contributed by the hybridization between the Ce 4f and H 1s orbitals. We should point out that they were predicted by the GGA-PBE functional. This raises a question how strong the SOC interaction and the correlated effect would be associated with the Ce 4f orbitals in these compounds. So, we further checked the electronic band structures with the SOC interaction, the correction of the on-site Coulomb interaction of Ce 4f orbitals, and their combination, which are presented in Fig. S10 in the SM. We find that both the SOC interaction and the correction of the on-site Coulomb interaction have much weak effect on the bands around , which could be neglected. The superconducting transition temperature is largely affected by the electronic properties at the . Therefore, the computational methodology employed in the present study would be acceptable to give a reasonable prediction on the superconducting properties of Ce-containing ternary hydrides. The similar treatment with ignoring the SOC interaction and the correlated effect has also been employed in several of the previous theoretical studies Peng et al. (2017); Wang et al. (2021); Li et al. (2019) on the electron-phonon coupling of binary CeHx. Unfortunately, the combination of the DFT+U method and the density-functional perturbation theory approach in the latest version (v7.1) of QE code does not support the electron-phonon interaction calculations. Because of the above reasons, our superconductivity calculations did not consider the SOC interaction and the strongly-correlated effect.
The phonon dispersions with the mode-resolved electron-phonon coupling (EPC) constant , atom-projected phonon density of states (PHDOS), and EPC spectra of -YCeH8, -YCeH18, -YCeH20, -LaCeH8, and -LaCeH20 at high pressures are presented Fig. 4 and Fig. S12 in the SM. For these five phases, the phonons with frequencies above 10 THz are dominated by the H contribution, accounting for more than 70%. The phonons with frequencies below 10 THz are mainly contributed by La, Ce, and Y atoms. Because of the difference in the atomic masses of Y, La, and Ce, the low-frequency range in the phonon spectrum of YCeHx is slightly wider than that of LaCeHx. The EPC constants of YCeH8 at 100 GPa, YCeH18 at 150 GPa, YCeH20 at 300 GPa, LaCeH8 at 100 GPa, and LaCeH20 at 250 GPa are 0.41, 2.51, 1.02, 0.35, and 0.99, respectively. The EPC constant of YCeH18 is so significant compared to the other four phases. From the mode-resolved EPC constant , we find that the largest contribution to the total EPC constant of YCeH18 comes from the optical phonon branch in the frequency range of 10 to 20 THz, which contributes about 0.64 (i.e., 25.7% to the total EPC constant). The rare-earth-atom associated vibrational modes with the frequencies of 10 THz for YCeH18 also have moderate contribution to the total EPC constant, as compared with the other four phases. In contrast, the contribution of rare-earth atoms to the EPC for YCeH20 with higher H content is only 0.09. When the pressure is increased from 150 GPa to 300 GPa, the phonons dominated by the H-atom-associated vibration in YCeH18 keep hardening, which can been seen from Fig. 4(a) and Fig. S11(c). The increase of pressure leads to a decrease of the EPC constant of YCeH18, so that the EPC constant of YCeH18 at 300 GPa is very close to that of LaCeH20 at 300 GPa. The EPC constants of CeH8 ( = Y and La) are so small, compared to CeH18 and CeH20, since both the highest phonon frequencies and the H-derived electronic density of states in the former cases are lower than those in the latter cases. These results suggest that once the hydrogen atoms in metal hydrides simultaneously provide sufficient electron density of states at and sufficient phonon density of states in the high-frequency region, a strong EPC strength would be achieved.
Based on the Allen-Dynes-modified McMillan formula Allen and Dynes (1975) with a widely accepted value of the Coulomb pseudopotential ( = 0.1 used herein), the values of -YCeH8 at 100 GPa, -YCeH18 at 150 GPa, -YCeH20 at 300 GPa, -LaCeH8 at 100 GPa, and -LaCeH8 at 250 GPa are predicted to be 4.8, 173.8, 122.1, 1.8, and 116.1 K, respectively. We note that very recently three independent high-pressure experiments Chen et al. (2022a); Bi et al. (2022); Huang et al. (2022) on the La-Ce-H system have been reported, in which the measurement pressures by Chen et al. Chen et al. (2022a) and Bi et al. Bi et al. (2022) were both less than 130 GPa and the observed was around 180 K. One interesting thing is that even though these two research groups Chen et al. (2022a); Bi et al. (2022) used different ratios of the starting materials for synthesis, the final measurement results were very close, suggesting that both of them synthesized probably the same high-temperature superconducting phase. Our calculation results show that at this pressure (130 GPa) the thermodynamically stable phase is LaCeH18, which is not explored further here because the stabilization of this phase may involve anharmonic approximation effect not the focus of this work. The measurement pressure on La-Ce-H by Huang et al. Huang et al. (2022) was in the range of 140-160 GPa. As mentioned above, LaCeH20 can be stabilized in this pressure range. A high-pressure experiment study Chen et al. (2022c) on the Y-Ce-H system has also been reported very recently, in which the observed values are between 97-140 K. The predicted value of YCeH18 by either the McM or AD formula shows a good agreement with this experiment data. To a certain extent, these experiment results show strong support to the validity of our theoretical prediction. In short, the high-pressure phases of LaCeH18, LaCeH20, and YCeH18 discovered computationally in this work have all been confirmed in the very recent high-pressure experiments, among which LaCeH20 and YCeH18 are theoretically predicted for the first time.
III.3 Discussion
As presented above, CeH8, YCeH18 and CeH20 ( = La and Y) have the space groups of , and , respectively, and the rare earth atoms are encased in the H18, H29 and H32 cages, respectively. Herein we performed the COHP and integrated COHP (ICOHP) analyses to further investigate the chemical bonding in these cage structures. The critical bond lengths of some representative H-H, Ce-H, and -H ( - La, Y) atom pairs and their corresponding ICOHP values that can reflect the bonding strength are shown in Figure 5. According to the convention of COHP analysis, the positive and negative values of COHP indicate the bonding and anti-bonding characteristics, respectively. As seen from Fig. 3, the H18 cage in YCeH8 can be viewed as consisting of eight quadrangles and four hexagons. Accordingly, the H-H bond in the H18 cages can be roughly classified into three groups: i) H-H bonds shared by two neighboring quadrangles, ii) H-H bonds shared by two neighboring hexagons, and iii) H-H bonds shared by a quadrangle and a hexagon.
The H-H bond lengths in the first and second groups are denoted as D1 and D2, respectively, as indicated in Figs. S14(a) and S14(d) in the SM. These two distinct bond lengths D1 and D2 in YCeH8 and LaCeH8 at 100 GPa can be sorted in the following descending order: D2(LaCeH8) D1(YCeH8) D1(LaCeH8) D2(YCeH8). Although the H-H bonds with D1 in YCeH8 and LaCeH8 have the very close bond lengths, the projected COHPs for them shown in Figs. 5(a) and 5(c) indicate that this H-H bond in LaCeH8 exhibits a stronger bonding interaction than in YCeH8 because of the less anti-bonding states appearing just below and a slightly greater absolute value of the corresponding ICOHP in LaCeH8. However, compared to the -H and Ce-H bonds from their corresponding ICOHP values, the interaction of this H-H bond would be significantly weaker. For YCeH18 and CeH20 with higher H content, the H-H bonds are significantly shrunken because of a higher pressure required to stabilize YCeH18 and CeH20, the ICOHP values of the H-H bonds in YCeH18 and ACeH20 are almost three times than those in CeH8, and much less (and even no) anti-bonding states appear below for the H-H bonds in YCeH20 (and LaCeH20). From the H18 cages in CeH8 to the H29 cages in YCeH18 and H32 cages in CeH20, the cage size increases and the number of quadrangles (hexagons) decreases (increases) due to incorporating more H atoms. These results suggest that the H-H bonds within anti-bonding states in CeH8 could be much stabilized in YCeH18 and CeH20 because they seem ready to be hybridized with the incorporated additional H atoms. In addition, YCeH20 has some remarkable anti-bonding states appearing around -11 eV for the Ce-H atom pair, as seen from Fig. 5(c), which mainly come from the hybridization between the H-1s and Ce-5px orbitals, suggesting that the semi-core states (such as 5p orbitals) of Ce in YCeH20 at the high pressure may contribute to the Ce-H bonding.
The crystal orbital bond index (COBI) can be used to quantify the strength of covalent bonds in covalently bonded solid materials. Müller et al. (2021) The integrated-COBI (abbreviated to ICOBI) values for the H-H bonds in these caged structures of CeH8, YCeH18 and CeH20 are less than 0.15, indicating very weak bonding interactions between H-H. The overall trend is that the ICOBI of H-H bonds increases with the increase of pressure, but the change is less than 0.02, suggesting that the effect of pressure on the H-H bonding is small. The similar trend is also observed from the analysis of ICOHP. One interesting thing is that the bonding strength of the H-H bonds in -YCeH20 in the pressure range of 300-400 GPa is slightly stronger than the one in -LaCeH20 according to both the electron localization function (ELF) and ICOHP analyses, although the corresponding ICOBI values in these two structures are quite close to each other and also exhibit a very weak dependence on the pressure.
The five superconducting compounds reported above, which are both thermodynamically and dynamically stable, are mainly of the CeH8, CeH18 and CeH20 types. The previously reported ScCaH8 Shi et al. (2021) and YMgH8 Song et al. (2022a) also have the same structure as CeH8. For the CeH8 type, there are three distinct types of H-H bonds, as mentioned above, in which the D1 decreases as the pressure increases, while the D2 exhibits a weaker dependence on the pressure. In additional, the bonding strength of the H-H bonds can also be described as the ELF.
Recently, Belli et al. Belli et al. (2021) have proposed a strong correlation between the value of superconducting hydride and the weak covalent H-H bonds. They also proposed a possible networking value , which is defined as the highest value of the ELF that creates an isosurface spanning through the whole crystal in all three Cartesian directions, to describe the nature of the H-H bond in hydrides. As noted by Belli et al., Belli et al. (2021) the precise acquisition of the value is not uniquely defined. From the radial distribution function of H atoms in our studied YCeHx and LaCeHx phases as a function of pressure, as presented in Fig. S14 in the SM, we note that within the radial range of Å there is a main peak representing the most important distribution of H-H bond lengths. In this way, we obtain the value by taking the average of all ELF values at the valley points of the lines plots of ELF along different H-H bonds that correspond to such a main peak. And according to the model proposed by Belli et al., Belli et al. (2021) the of metal hydride can be predicted by the following equation with an accuracy in the range of 60 K:
(9) |
where is the hydrogen fraction and is the hydrogen fraction of the total DOS at . As presented in Table 1, the network values of our studied phases are all less than 0.5 and they are very close. Consequently, the values predicted by Eq. (9) are more significantly affected by the contribution of . We also noted that the values predicted by Eq. (9) are still somewhat deviation from those predicted by the AD formula, i.e., Eq. (3). Through the examination on YCeH18 at 150 and 300 GPa, it can be seen that when the pressure increases, the H-H distance decreases, the overall network value exhibits an increasing trend, and the contribution of H to the density of states at also increases. And thus the Eq. (9) based on ELF predicts that the for YCeH18 increases with the increase of pressure, which is opposite to the trend in the prediction of the AD formula. This indicates that one possible limitation in the use of Eq. (9) to predict may be the pressure-dependence of . From the predicted results of Peng et al. Peng et al. (2017) for all binary rare-earth metal hydrides, a trend can be seen that the elements containing more f electrons exhibit a suppressing effect on the value. However, the dataset to build Eq. (9) in the work of Belli et al. Belli et al. (2021) contains too few the f-electron-containing cases (i.e., only CeH9 and PrH9 included therein). Therefore, the does exhibit a certain relationship with the and values, however, the generalization of Eq. (9) to the pressure-dependence and the f-electron-containing systems needs a further improvement.
IV Conclusion
In summary, we employed the evolutionary algorithm and first-principles calculations to study the ternary phase diagrams of YCeHx and LaCeHx systems under the pressure range of 100-400 GPa. The clathrate structure surrounded by H18, H29, and H32 can maintain thermodynamic stability in the searched pressure range. These cage structures correspond to YCeH8, YCeH18, YCeH20, LaCeH8, LaCeH18, LaCeH20, respectively. However, under the harmonic approximation, LaCeH18 cannot achieve dynamic stability. We further studied the superconductivity of YCeH8, YCeH18,YCeH20, LaCeH8, and LaCeH20. The results show that all four can achieve high temperature superconductivity, and -YCeH18 can have the highest 173 K at 150GPa. In addition, in terms of the structure of these phases, -YCeH8, -LaCeH8, -YCeH18, -YCeH20 can all be obtained by element replacement in binary superconducting hydrides. In other words, the search for new superconducting hydrides can start from the combination of binary superconducting hydrides. When two kinds of binary superconducting hydrides with similar properties are combined together, it is possible to form a novel stable three ternary hydride.
Acknowledgments
The computations in this work have been performed using the facilities of Research Center for Advanced Computing Infrastructure (RCACI) at JAIST. P.S. is grateful for financial support from Grant-in-Aid for JSPS Research Fellow (JSPS KAKENHI Grant No. 22J10527). K.H. is grateful for financial support from the HPCI System Research Project (Project ID: hp190169) and MEXT-KAKENHI (JP16H06439, JP17K17762, JP19K05029, and JP19H05169). R.M. is grateful for financial supports from MEXT-KAKENHI (19H04692 and 16KK0097), FLAGSHIP2020 (project nos. hp1 90169 and hp190167 at K-computer), Toyota Motor Corporation, I-O DATA Foundation, the Air Force Office of Scientific Research (AFOSR-AOARD/FA2386-17-1-4049;FA2386-19-1-4015), and JSPS Bilateral Joint Projects (with India DST).
References
- Tinkham (2004) M. Tinkham, Introduction to superconductivity (Courier Corporation, 2004).
- Gennes (2018) P. G. D. Gennes, Superconductivity of Metals and Alloys (CRC Press, 2018).
- Linder and Robinson (2015) J. Linder and J. W. A. Robinson, Nat. Phys. 11, 307 (2015).
- Josephson (1962) B. Josephson, Phys. Lett 1, 251 (1962).
- Bergen et al. (2019) A. Bergen, R. Andersen, M. Bauer, H. Boy, M. ter Brake, P. Brutsaert, C. Bührer, M. Dhallé, J. Hansen, H. ten Kate, J. Kellers, J. Krause, E. Krooshoop, C. Kruse, H. Kylling, M. Pilas, H. Pütz, A. Rebsdorf, M. Reckhard, E. Seitz, H. Springer, X. Song, N. Tzabar, S. Wessel, J. Wiezoreck, T. Winkler, and K. Yagotyntsev, Supercond. Sci. Technol. 32, 125006 (2019).
- Islam et al. (2014) M. R. Islam, Y. Guo, and J. Zhu, Renew. Sust. Energ. Rev. 33, 161 (2014).
- Thomas et al. (2016) H. Thomas, A. Marian, A. Chervyakov, S. Stückrad, D. Salmieri, and C. Rubbia, Renew. Sust. Energ. Rev. 55, 59 (2016).
- Tinkham (1988) M. Tinkham, Phys. Rev. Lett. 61, 1658 (1988).
- Kalsi (2011) S. S. Kalsi, Applications of high temperature superconductors to electric power equipment (John Wiley & Sons, 2011).
- Flores-Livas et al. (2020) J. A. Flores-Livas, L. Boeri, A. Sanna, G. Profeta, R. Arita, and M. Eremets, Phys. Rep. 856, 1 (2020).
- Bardeen (1955) J. Bardeen, Phys. Rev. 97, 1724 (1955).
- Cooper (1956) L. N. Cooper, Phys. Rev. 104, 1189 (1956).
- Bardeen et al. (1957a) J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Phys. Rev. 106, 162 (1957a).
- Bardeen et al. (1957b) J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Phys. Rev. 108, 1175 (1957b).
- Ashcroft (1968) N. W. Ashcroft, Phys. Rev. Lett. 21, 1748 (1968).
- Wigner and Huntington (1935) E. Wigner and H. B. Huntington, J. Chem. Phys. 3, 764 (1935).
- Loubeyre et al. (1996) P. Loubeyre, R. LeToullec, D. Hausermann, M. Hanfland, R. J. Hemley, H. K. Mao, and L. W. Finger, Nature 383, 702 (1996).
- McMinis et al. (2015) J. McMinis, R. C. Clay, D. Lee, and M. A. Morales, Phys. Rev. Lett. 114, 105305 (2015).
- Eremets and Troyan (2011) M. I. Eremets and I. A. Troyan, Nat. Mater. 10, 927 (2011).
- Dalladay-Simpson et al. (2016) P. Dalladay-Simpson, R. T. Howie, and E. Gregoryanz, Nature 529, 63 (2016).
- Dias and Silvera (2017) R. P. Dias and I. F. Silvera, Science 355, 715 (2017).
- Castelvecchi (2017) D. Castelvecchi, Nature 542, 17 (2017).
- Celliers et al. (2018) P. M. Celliers, M. Millot, S. Brygoo, R. S. McWilliams, D. E. Fratanduono, J. R. Rygg, A. F. Goncharov, P. Loubeyre, J. H. Eggert, J. L. Peterson, N. B. Meezan, S. L. Pape, G. W. Collins, R. Jeanloz, and R. J. Hemley, Science 361, 677 (2018).
- Ashcroft (2004) N. W. Ashcroft, Phys. Rev. Lett. 92, 187002 (2004).
- Semenok et al. (2020) D. V. Semenok, I. A. Kruglov, I. A. Savkin, A. G. Kvashnin, and A. R. Oganov, Curr. Opin. Solid State Mater. Sci. 24, 100808 (2020).
- Peng et al. (2017) F. Peng, Y. Sun, C. J. Pickard, R. J. Needs, Q. Wu, and Y. Ma, Phys. Rev. Lett. 119, 107001 (2017).
- Zhang et al. (2020) J. Zhang, J. M. McMahon, A. R. Oganov, X. Li, X. Dong, H. Dong, and S. Wang, Phys. Rev. B 101, 134108 (2020).
- Li and Peng (2017) X. Li and F. Peng, Inorg. Chem. 56, 13759 (2017).
- Yu et al. (2015) S. Yu, X. Jia, G. Frapper, D. Li, A. R. Oganov, Q. Zeng, and L. Zhang, Sci. Rep. 5, 1 (2015).
- Einaga et al. (2016) M. Einaga, M. Sakata, T. Ishikawa, K. Shimizu, M. I. Eremets, A. P. Drozdov, I. A. Troyan, N. Hirao, and Y. Ohishi, Nat. Phys. 12, 835 (2016).
- Drozdov et al. (2015) A. P. Drozdov, M. I. Eremets, I. A. Troyan, V. Ksenofontov, and S. I. Shylin, Nature 525, 73 (2015).
- Drozdov et al. (2019) 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, Nature 569, 528 (2019).
- Salke et al. (2019) N. P. Salke, M. M. D. Esfahani, Y. Zhang, I. A. Kruglov, J. Zhou, Y. Wang, E. Greenberg, V. B. Prakapenka, J. Liu, A. R. Oganov, and J.-F. Lin, Nat. Commun. 10, 1 (2019).
- Chen et al. (2021a) W. Chen, D. V. Semenok, A. G. Kvashnin, X. Huang, I. A. Kruglov, M. Galasso, H. Song, D. Duan, A. F. Goncharov, V. B. Prakapenka, A. R. Oganov, and T. Cui, Nat. Commun. 12, 1 (2021a).
- Shao et al. (2021a) M. Shao, W. Chen, K. Zhang, X. Huang, and T. Cui, Phys. Rev. B 104, 174509 (2021a).
- Shao et al. (2021b) M. Shao, S. Chen, W. Chen, K. Zhang, X. Huang, and T. Cui, Inorg. Chem. 60, 15330 (2021b).
- Troyan et al. (2021) I. A. Troyan, D. V. Semenok, A. G. Kvashnin, A. V. Sadakov, O. A. Sobolevskiy, V. M. Pudalov, A. G. Ivanova, V. B. Prakapenka, E. Greenberg, A. G. Gavriliuk, I. S. Lyubutin, V. V. Struzhkin, A. Bergara, I. Errea, R. Bianco, M. Calandra, F. Mauri, L. Monacelli, R. Akashi, and A. R. Oganov, Adv. Mater. 33, 2006832 (2021).
- Chen et al. (2021b) W. Chen, D. Semenok, X. Huang, H. Shu, X. Li, D. Duan, T. Cui, and A. Oganov, Phys. Rev. Lett. 127, 117001 (2021b).
- Kong et al. (2021) 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, Nat. Commun. 12, 1 (2021).
- Li et al. (2014) Y. Li, J. Hao, H. Liu, Y. Li, and Y. Ma, J. Chem. Phys 140, 174712 (2014).
- Somayazulu et al. (2019) M. Somayazulu, M. Ahart, A. K. Mishra, Z. M. Geballe, M. Baldini, Y. Meng, V. V. Struzhkin, and R. J. Hemley, Phys. Rev. Lett. 122, 027001 (2019).
- Song et al. (2021) P. Song, Z. Hou, P. B. de Castro, K. Nakano, K. Hongo, Y. Takano, and R. Maezono, Chem. Mater. 33, 9501 (2021).
- Ma et al. (2017a) Y. Ma, D. Duan, Z. Shao, D. Li, L. Wang, H. Yu, F. Tian, H. Xie, B. Liu, and T. Cui, Phys. Chem. Chem. Phys. 19, 27406 (2017a).
- Ma et al. (2017b) Y. Ma, D. Duan, Z. Shao, H. Yu, H. Liu, F. Tian, X. Huang, D. Li, B. Liu, and T. Cui, Phys. Rev. B 96, 144518 (2017b).
- Shao et al. (2019) Z. Shao, D. Duan, Y. Ma, H. Yu, H. Song, H. Xie, D. Li, F. Tian, B. Liu, and T. Cui, npj Comput. Mater. 5, 1 (2019).
- Zheng et al. (2021) J. Zheng, W. Sun, X. Dou, A.-J. Mao, and C. Lu, J. Phys. Chem. C 125, 3150 (2021).
- Rahm et al. (2017) M. Rahm, R. Hoffmann, and N. W. Ashcroft, J. Am. Chem. Soc. 139, 8740 (2017).
- Liang et al. (2021) X. Liang, A. Bergara, X. Wei, X. Song, L. Wang, R. Sun, H. Liu, R. J. Hemley, L. Wang, G. Gao, and Y. Tian, Phys. Rev. B 104, 134501 (2021).
- Sun et al. (2019) Y. Sun, J. Lv, Y. Xie, H. Liu, and Y. Ma, Phys. Rev. Lett. 123, 097001 (2019).
- Liang et al. (2019a) X. Liang, S. Zhao, C. Shao, A. Bergara, H. Liu, L. Wang, R. Sun, Y. Zhang, Y. Gao, Z. Zhao, X.-F. Zhou, J. He, D. Yu, G. Gao, and Y. Tian, Phys. Rev. B 100, 184502 (2019a).
- Wei et al. (2020) Y. K. Wei, L. Q. Jia, Y. Y. Fang, L. J. Wang, Z. X. Qian, J. N. Yuan, G. Selvaraj, G. F. Ji, and D. Q. Wei, Int. J. Quantum Chem. 121, 26459 (2020).
- Shi et al. (2021) L.-T. Shi, Y.-K. Wei, A.-K. Liang, R. Turnbull, C. Cheng, X.-R. Chen, and G.-F. Ji, J. Mater. Chem. C 9, 7284 (2021).
- Song et al. (2022a) P. Song, Z. Hou, P. B. de Castro, K. Nakano, Y. Takano, R. Maezono, and K. Hongo, Adv. Theory Simul. 5, 2100364 (2022a).
- Semenok et al. (2021) D. V. Semenok, I. A. Troyan, A. G. Ivanova, A. G. Kvashnin, I. A. Kruglov, M. Hanfland, A. V. Sadakov, O. A. Sobolevskiy, K. S. Pervakov, I. S. Lyubutin, K. V. Glazyrin, N. Giordano, D. N. Karimov, A. L. Vasiliev, R. Akashi, V. M. Pudalov, and A. R. Oganov, Mater. Today 48, 18 (2021).
- Liang et al. (2019b) X. Liang, A. Bergara, L. Wang, B. Wen, Z. Zhao, X.-F. Zhou, J. He, G. Gao, and Y. Tian, Phys. Rev. B 99, 100505 (2019b).
- Jiang et al. (2021) M.-J. Jiang, H.-L. Tian, Y.-L. Hai, N. Lu, P.-F. Tong, S.-Y. Wu, W.-J. Li, C.-L. Yang, and G.-H. Zhong, ACS Appl. Electron. Mater 3, 4172 (2021).
- Cui et al. (2020) W. Cui, T. Bi, J. Shi, Y. Li, H. Liu, E. Zurek, and R. J. Hemley, Phys. Rev. B 101, 134504 (2020).
- Cataldo et al. (2020) S. D. Cataldo, W. von der Linden, and L. Boeri, Phys. Rev. B 102, 014516 (2020).
- Song et al. (2022b) P. Song, Z. Hou, P. B. de Castro, K. Nakano, K. Hongo, Y. Takano, and R. Maezono, J. Phys. Chem. C 126, 2747 (2022b).
- Cataldo et al. (2022) S. D. Cataldo, W. von der Linden, and L. Boeri, Npj Comput. Mater. 8, 1 (2022).
- Vocaturo et al. (2022) R. Vocaturo, C. Tresca, G. Ghiringhelli, and G. Profeta, J. Appl. Phys. 131, 033903 (2022).
- Snider et al. (2020) E. Snider, N. Dasenbrock-Gammon, R. McBride, M. Debessai, H. Vindana, K. Vencatasamy, K. V. Lawler, A. Salamat, and R. P. Dias, Nature 586, 373 (2020).
- Hirsch and Marsiglio (2021) J. E. Hirsch and F. Marsiglio, Nature 596, E9 (2021).
- Hirsch (2022) J. E. Hirsch, Europhys. Lett. 137, 36001 (2022).
- Nakanishi et al. (2018) A. Nakanishi, T. Ishikawa, and K. Shimizu, J. Phys. Soc. Jpn. 87, 124711 (2018).
- Amsler (2019) M. Amsler, Phys. Rev. B 99, 060102 (2019).
- Guan et al. (2021) H. Guan, Y. Sun, and H. Liu, Phys. Rev. Res. 3, 043102 (2021).
- Chen et al. (2022a) W. Chen, X. Huang, D. V. Semenok, S. Chen, K. Zhang, A. R. Oganov, and T. Cui, arXiv preprint arXiv:2203.14353 (2022a).
- Glass et al. (2006) C. W. Glass, A. R. Oganov, and N. Hansen, Comput. Phys. Commun. 175, 713 (2006).
- Lyakhov et al. (2013) A. O. Lyakhov, A. R. Oganov, H. T. Stokes, and Q. Zhu, Comput. Phys. Commun. 184, 1172 (2013).
- Perdew et al. (1996) J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996).
- Kresse and Furthmüller (1996) G. Kresse and J. Furthmüller, Phys. Rev. B 54, 11169 (1996).
- Kresse and Joubert (1999) G. Kresse and D. Joubert, Phys. Rev. B 59, 1758 (1999).
- Müller et al. (2021) P. C. Müller, C. Ertural, J. Hempelmann, and R. Dronskowski, J. Phys. Chem. C. 125, 7959 (2021).
- Dudarev et al. (1998) S. L. Dudarev, G. A. Botton, S. Y. Savrasov, C. J. Humphreys, and A. P. Sutton, Phys. Rev. B 57, 1505 (1998).
- Timrov et al. (2022) I. Timrov, N. Marzari, and M. Cococcioni, arXiv preprint arXiv:2203.15684 (2022).
- Wang et al. (2021) C. Wang, S. Liu, H. Jeon, S. Yi, Y. Bang, and J.-H. Cho, Phys. Rev. B 104, l020504 (2021).
- Chen et al. (2012) Y. Chen, Q.-M. Hu, and R. Yang, Phys. Rev. Lett. 109, 157004 (2012).
- Vohra et al. (1999) Y. K. Vohra, S. L. Beaver, J. Akella, C. A. Ruddle, and S. T. Weir, J. Appl. Phys. 85, 2451 (1999).
- Chen et al. (2022b) L. Chen, T. Liang, Z. Zhang, H. Song, Z. Liu, Q. Jiang, Y. Chen, and D. Duan, J. Phys.: Condens. Matter 34, 204005 (2022b).
- Pickard and Needs (2007) C. J. Pickard and R. J. Needs, Nat. Phys. 3, 473 (2007).
- Akbarzadeh et al. (2007) A. R. Akbarzadeh, V. Ozoliņš, and C. Wolverton, Adv. Mater. 19, 3233 (2007).
- Virtanen et al. (2020) P. Virtanen, R. Gommers, T. E. Oliphant, M. Haberland, T. Reddy, D. Cournapeau, E. Burovski, P. Peterson, W. Weckesser, J. Bright, S. J. van der Walt, M. Brett, J. Wilson, K. J. Millman, N. Mayorov, A. R. J. Nelson, E. Jones, R. Kern, E. Larson, C. J. Carey, İ. Polat, Y. Feng, E. W. Moore, J. VanderPlas, D. Laxalde, J. Perktold, R. Cimrman, I. Henriksen, E. A. Quintero, C. R. Harris, A. M. Archibald, A. H. Ribeiro, F. Pedregosa, and P. van Mulbregt, Nat. Methods 17, 352 (2020).
- Togo and Tanaka (2015) A. Togo and I. Tanaka, Scr. Mater. 108, 1 (2015).
- Giannozzi et al. (2009) 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. Mauri, 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, J. Phys.: Condens. Matter 21, 395502 (2009).
- Corso (2014) A. D. Corso, Comp. Mater. Sci. 95, 337 (2014).
- McMillan (1968) W. L. McMillan, Phys. Rev. 167, 331 (1968).
- Allen and Dynes (1975) P. B. Allen and R. C. Dynes, Phys. Rev. B 12, 905 (1975).
- Liu et al. (2017) H. Liu, I. I. Naumov, R. Hoffmann, N. W. Ashcroft, and R. J. Hemley, Proc. Natl. Acad. Sci. 114, 6990 (2017).
- Belli et al. (2021) F. Belli, T. Novoa, J. Contreras-García, and I. Errea, Nat. Commun. 12, 5381 (2021).
- Li et al. (2019) B. Li, Z. Miao, L. Ti, S. Liu, J. Chen, Z. Shi, and E. Gregoryanz, J. Appl. Phys. 126, 235901 (2019).
- Bi et al. (2022) J. Bi, Y. Nakamoto, K. Shimizu, M. Zhou, H. Wang, G. Liu, and Y. Ma, arXiv preprint arXiv:2204.04623 (2022).
- Huang et al. (2022) G. Huang, T. Luo, P. Dalladay Simpson, L. Chen, Z. Cao, D. Peng, F. A. Gorelli, G. Zhong, H. Lin, and X. Chen, arXiv preprint arXiv:2208.05199 (2022).
- Chen et al. (2022c) L. Chen, T. Luo, P. Dalladay Simpson, G. Huang, Z. Cao, D. Peng, F. A. Gorelli, G. Zhong, H. Lin, and X. Chen, arXiv preprint arXiv:2208.05191 (2022c).
Phase | Space group | Natom | (ELF) | (McM) | (AD) | ||||||
---|---|---|---|---|---|---|---|---|---|---|---|
(GPa) | Å | (K) | (K) | (K) | (K) | ||||||
YCeH8 | 100 | 20 | 1.52 | 0.40 | 9.5% | 0.41 | 987.5 | 24.5 | 4.7 | 4.8 | |
YCeH18 | 150 | 22 | 1.07 | 0.43 | 34.4% | 2.51 | 812.2 | 118.3 | 133.8 | 173.8 | |
YCeH18 | 300 | 22 | 1.05 | 0.46 | 39.2% | 0.97 | 1547.7 | 142.2 | 103 | 110 | |
YCeH20 | 300 | 66 | 1.05 | 0.45 | 38.2% | 1.03 | 1595.2 | 137.6 | 114.2 | 122.1 | |
LaCeH8 | 100 | 20 | 1.56 | 0.38 | 11.2% | 0.35 | 975.8 | 24.9 | 1.73 | 1.76 | |
LaCeH20 | 250 | 66 | 1.06 | 0.48 | 34.7% | 1.00 | 1566.9 | 145.0 | 109.1 | 116.1 |




