This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Hole-Doping Effect on Superconductivity in Compressed CeH9 at High Pressure

Chongze Wang1, Shuyuan Liu1, Hyunsoo Jeon1, Seho Yi1, Yunkyu Bang2,3, and Jun-Hyung Cho1,3∗ 1 Department of Physics, Research Institute for Natural Science, and Institute for High Pressure at Hanyang University, Hanyang University, 222 Wangsimni-ro, Seongdong-Ku, Seoul 04763, Republic of Korea
2 Department of Physics, Pohang University of Science and Technology, Pohang 37673, Republic of Korea
3 Asia Pacific Center for Theoretical Physics (APCTP), Pohang-si, Gyeongsangbuk-do 37673, Republic of Korea
Abstract

The experimental realization of high-temperature superconductivity in compressed hydrides H3S and LaH10 at high pressures over 150 GPa has aroused great interest in reducing the stabilization pressure of superconducting hydrides. For cerium hydride CeH9 recently synthesized at 80-100 GPa, our first-principles calculations reveal that the strongly hybridized electronic states of Ce 4ff and H 1ss orbitals produce the topologically nontrivial Dirac nodal lines around the Fermi energy EFE_{F}, which are protected by crystalline symmetries. By hole doping, EFE_{F} shifts down toward the topology-driven van Hove singularity to significantly increase the density of states, which in turn raises a superconducting transition temperature TcT_{c} from 74 K up to 136 K at 100 GPa. The hole-doping concentration can be controlled by the incorporation of Ce3+ ions with varying their percentages, which can be well electronically miscible with Ce atoms in the CeH9 matrix because both Ce3+ and Ce behave similarly as cations. Therefore, the interplay of symmetry, band topology, and hole doping contributes to enhance TcT_{c} in compressed CeH9. This mechanism to enhance TcT_{c} can also be applicable to another superconducting rare earth hydride LaH10.

Doping in condensed matters is a well-established means of manipulating their electronic structures, which may lead to the emergence of various quantum phases with exotic physical properties [1, 2, 3, 4, 5, 6, 7]. For example, in the unconventional high-temperature superconductors such as cuprates [8] and pnictides [9, 10], doping by holes or electrons has been demonstrated not only to induce complex quantum phase transitions including magnetism, pseudogap, charge density wave, superconductivity (SC), and Fermi liquid phases, but also to vary TcT_{c} in their superconducting phases [4, 5, 6, 7]. Due to the emergence of such many electronic states in unconventional high-TcT_{c} superconductors, identifying the mechanism responsible for the doping-induced changes of TcT_{c} has been elusive. By contrast, doping effect in conventional Bardeen-Cooper-Schrieffer (BCS) [11] superconductors has been relatively well understood in terms of the influence of electron-phonon coupling (EPC), and therefore various dopants can be employed to tune TcT_{c}. It is thus very interesting and challenging to investigate the effect of doping on the EPC-driven SC of recently discovered hydrides at high pressures [12].

During the past six years, compressed hydrides under megabar pressures have attracted much attention because of their unprecedented records of TcT_{c}. Motivated by the theoretical predictions of SC in a number of hydrides [13, 16, 14, 15, 17, 18, 19, 20, 21, 22], experiments have confirmed that sulfur hydride H3S and lanthanum hydride LaH10 exhibit TcT_{\rm c} around 203 K at {\sim}155 GPa [23] and 250-260 K at {\sim}170 GPa [24, 25], respectively. More recently, carbonaceous sulfur hydride was experimentally realized to reach a room-temperature SC with a TcT_{\rm c} of 288 K at {\sim}267 GPa [26]. Despite such an achievement of room-temperature SC, it is highly demanding to discover high-TcT_{c} superconducting hydrides synthesized at moderate pressures below {\sim}100 GPa, which can be normally achievable with the diamond anvil cell [27, 28]. Near simultaneously, two experimental groups [29, 30] reported the successful synthesis of cerium hydride CeH9 at 80-100 GPa. The subsequent density-functional theory (DFT) calculation of CeH9 revealed that the delocalized nature of Ce 4ff electrons is an essential ingredient in the high chemical precompression of clathrate H cage around Ce atom [see Fig. 1(a)]. It is noticeable that, even though the synthesis of CeH9 was made at lower pressures below {\sim}100 GPa, its theoretically predicted TcT_{c} value was around 75 K [19], much lower than those of H3S and LaH10 [23, 24, 25]. Therefore, the main bottleneck for the research of high-TcT_{c} superconducting hydrides has been associated with difficulties both raising TcT_{c} and lowering the pressure of stability simultaneously. In order to alleviate this bottleneck in CeH9, we here investigate the effect of hole doping on SC, which leads to a significant increase in TcT_{c}.

For high-pressure rare earth hydrides with clathrate H-cage structures, the electronic states tend to have a strong hybridization between rare earth-4ff and H-1ss orbitals near EFE_{\rm F} [20, 31, 32, 33, 34, 35]. This electronic characteristic of rare earth hydrides having high-symmetry structures could be favorable for hosting topological states through band inversions, identified in recent studies of topological materials [36, 37]. As compelling examples, topologically nontrivial Dirac-nodal-line (DNL) states are jointly protected by the space inversion symmetry PP and time-reversal symmetry TT supplemented by additional crystalline symmetry [38, 39, 40, 41]. However, exploration of the cooperative interplay of crystal symmetry and band topology has so far been overlooked in high-pressure superconducting hydrides. These new ingredients of symmetry and topology together with hole doping will provide a promising playground to enhance TcT_{c} in high-pressure superconducting hydrides, as will be demonstrated below.

In this Letter, using first-principles calculations, we discover that CeH9 possessing a hexagonal-close-packed (hcp) structure has symmetry-enforced DNL states. It is revealed that the two-dimensional (2D) nodal surface guaranteed by the nonsymmorphic crystal symmetry S2zS_{2z} (equivalent to the combination of twofold rotation symmetry C2zC_{2z} about the zz axis and a half translation along the zz direction) is converted to one-dimensional (1D) DNLs in the presence of spin-orbit coupling (SOC) [42]. Moreover, two DNL states composed of strongly hybridized Ce 4ff and H 1ss orbitals touch each other, leading to the formation of a van Hove singularity (vHs) around -1.6 eV below EFE_{F}. Consequently, hole doping shifts EFE_{F} toward the vHs, which in turn increases EPC and therefore raises TcT_{c} from 74 K (without hole doping) up to 136 K at 100 GPa. Considering that Ce atoms in the CeH9 matrix behave as cations, Ce3+ ions is expected to be well electronically miscible with Ce atoms and their incorporation percentages can control hole-doping concentrations. Our findings provide a new avenue for using hole doping to enhance TcT_{c} in recently synthesized rare earth hydrides CeH9 [29, 30] as well as LaH10 [24, 25].

Refer to caption
Figure 1: (a) Optimized hcp structure of Ce atoms in CeH9. The inset shows the top view of Ce atoms, and the isolated H29 cage surrounding a Ce atom is also included. The calculated band structure of CeH9 together with the PDOS for Ce 4ff, Ce 5dd, and H 1ss orbitals is given in (b). The unit of DOS is states/eV per unit cell that contains two Ce atoms. N1N_{1} and N2N_{2} represent the fourfold degenerate bands along the high-symmetry HH-AA-LL-HH paths, and the energy zero is EFE_{F}. The Brillouin zone of hcp structure is also included in (b).

We first present the electronic band structure of hcp CeH9, obtained using first-principles DFT calculations [43]. In most of the calculations hereafter, we fix a pressure of 100 GPa at which the hcp phase with the lattice parameters aa = bb = 3.698 Å and cc = 5.596 Å [see Fig. 1(a)] is thermodynamically stable [see Fig. S1(a) in the Supplemental Material [53]]. It is noted that at 70 GPa, the hcp phase becomes dynamically unstable with the presence of imaginary phonon frequencies [see Fig. S1(b)]. Figure 1(b) shows the calculated band structure and partial density of states (PDOS) of CeH9. We find that the Ce 4ff and H 1ss orbitals are more dominant components in the electronic states around EFE_{F}, compared to other orbitals (see Fig. S2 in the Supplemental Material [53]). Interestingly, the PDOS for Ce 4ff and H 1ss orbitals exhibits a sharp peak around -1.6 eV below EFE_{F} [see Fig. 1(b)], indicating a strong hybridization of the two orbitals. The existence of such a vHs having large DOS leads to an increase of TcT_{c} via hole doping, as discussed below.

Figure 1(b) represents the DFT band structure computed without including SOC. The presence of PP and TT symmetries ensures Kramer’s double degeneracy in the whole Brillouin zone (BZ). We find that there are fourfold degenerate bands N1N_{1} and N2N_{2} along the high-symmetry H-A-L-H paths, formed by touching of two bands. It is noted that N1N_{1} and N2N_{2} touch each other at AA and between AA and LL [marked by dashed circles in Fig. 1(c)], thereby giving rise to eigthfold accidental degeneracies. Using the tight-binding Hamiltonian with a basis of maximally localized Wannier functions [54, 55], we reveal the existence of 2D nodal surfaces NS1NS_{1} and NS2NS_{2} throughout the kzk_{z} = π{\pi}/cc plane, as shown in Fig. 2(a). Here, each nodal surface is formed by a touching of two doubly-degenerate bands at the boundary of BZ. Since the crystalline symmetry of hcp CeH9 belongs to the space group P63/mmcP6_{3}/mmc (No. 194) with the point group D6hD_{6h}, the fourfold degeneracy of NS1NS_{1} and NS2NS_{2} is respected by the combined symmetry PS2zPS_{2z}, whose eigenvalues are ±1{\pm}1 because of (PS2zPS_{2z})2 = 1 (see symmetry analysis in the Supplemental Material [53]). The inclusion of SOC lifts the degeneracy of N1N_{1} and N2N_{2} along the HH-AA-LL-HH paths except AA-LL (see Fig. S3 in the Suppelemental Material [53]), where the SOC-induced gap opening is less than {\sim}0.1 eV [see Fig. 2(b)]. It is noted that the nodal surfaces NS1NS_{1} and NS2NS_{2} are converted into 1D nodal lines along the high-symmetry paths kxk_{x} = 0 and kxk_{x} = ±3ky{\pm}\sqrt{3}k_{y} as well as with circular patterns around the AA point [see Fig. 2(b)]. These DNLs showing C3zC_{3z} rotation symmetry are protected by additional mirror symmetry (see symmetry analysis in the Supplemental Material [53]). For example, MxM_{x} : (xx,yy,zz) {\rightarrow} (x-x,yy,zz) anticommuting with PS2zPS_{2z} allows the existence of the fourfold degenerate nodal line at kxk_{x} = 0 on the kzk_{z} = π{\pi}/cc plane. The topological characterizations of these DNLs are demonstrated by calculating the topological index [41], defined as ζ1{\zeta}_{1} = 1π{\frac{1}{\pi}} {\oint}c dkdk{\cdot}A(kk), along a closed loop encircling any of the DNLs. Here, A(k) = i-i<<uku_{k}\mid\partialk\miduku_{k}>> is the Berry connection of the related Bloch bands. We obtain ζ1{\zeta}_{1} = ±{\pm}1 for the DNLs, indicating that they are stable against PS2zPS_{2z} and MM symmetries conserving perturbations.

Refer to caption
Figure 2: (a) Energy of 2D nodal surfaces NS1NS_{1} and NS2NS_{2} throughout the kzk_{z} = π{\pi}/cc plane, obtained without including SOC and (b) 1D nodal lines converted from NS1NS_{1} and NS2NS_{2} with including SOC. In (b), the SOC-induced gap is represented using the color scale in the range between 0 and 100 meV.

As shown in Figs. 1(b) and 2(a), the position of vHs arising from the saddle points of energy dispersion is located near the band touching of N1N_{1} (NS1NS_{1}) and N2N_{2} (NS2NS_{2}). It is thus likely that crystalline symmetries producing the nontrivial band topology of CeH9 are correlated with the existence of vHs. In other words, the present vHs is rendered emergent by the band touching of the two fourfold degenerate topological states below EFE_{F} and yield a power law divergence in the DOS [56], which in turn enhances TcT_{c} via hole doping, as discussed below.

To estimate TcT_{c} of CeH9 at 100 GPa, we calculate the phonon spectrum, projected phonon DOS onto Ce and H atoms, Eliashberg function α2F(ω){\alpha}^{2}F({\omega}), and integrated EPC constant λ(ω){\lambda}({\omega}) as a function of phonon frequency. Figure 3(a) shows that the phonon spectrum is divided into two regimes: i.e., one is the low-frequency regime arising from the vibrations of Ce atoms and the other is the high-frequency regime driven by H atoms. Therefore, we find that the acoustic phonon modes of Ce atoms contribute to {\sim}19% of the total EPC constant λ{\lambda} = λ{\lambda}({\infty}), whereas the optical phonon modes of H atoms contribute to {\sim}81% of λ{\lambda}. Specifically, the H-derived low-frequency optical modes close to the Ce-derived acoustic modes show larger EPC strength, as represented by circles on the phonon dispersion in Fig. 3(a). Based on these results, we can say that the optical vibrations of H atoms are strongly coupled to the hybridized electronic states of Ce 4ff and H 1ss orbitals around EFE_{F}, giving rise to λ{\lambda} = 1.04. By numerically solving the isotropic Migdal-Eliashberg equations [57, 58, 59], we calculate the superconducting gap versus temperature with varying Coulomb pseudopotential parameter μ{\mu}^{*} [19, 30], and estimate TcT_{c} {\approx} 84 and 74 K with μ{\mu}^{*} = 0.1 and 0.13, respectively [see Fig. 3(b)] [60]. These predicted TcT_{c} values of CeH9 are much lower than the experimentally observed TcT_{c} {\approx} 260 K of LaH10 [24, 25]. The lower TcT_{c} in CeH9 is associated with relatively lower EPC constant compared to the case of LaH10 [31, 32, 33]. It is also noted that the H-derived DOS of CeH9 at EFE_{F} is smaller than that of LaH10 [see fig. 4(a)].

Refer to caption
Figure 3: (a) Calculated phonon spectrum, phonon DOS projected onto Ce and H atoms, Eliashberg function α2F(ω){\alpha}^{2}F({\omega}), and integrated EPC constant λ(ω){\lambda}({\omega}) of CeH9, (b) superconducting energy gap Δ{\Delta} as a function of temperature with μ{\mu}^{*} = 0.1 and 0.13, and (c) λ{\lambda} and TcT_{c} as a function of nhn_{h}.

Since a vHs in the electronic DOS of CeH9 is located below EFE_{F} [see Fig. 1(b)], hole doping is expected to induce a shift of EFE_{F} toward the vHs. The calculated band structure at a hole doping of nhn_{h} = 1.0ee per Ce atom shows that EFE_{F} approaches the vHs, thereby giving rise to an increase of DOS around EFE_{F} (see Fig. S4 in the Supplemental Material [53]). In order to examine how the hole doping influences SC, we use the isotropic Migdal-Eliashberg formalism [57, 58, 59] to estimate the variations of λ{\lambda} and TcT_{c} as a function of nhn_{h}. As shown in Fig. 3(c), λ{\lambda} is enhanced from 1.04 (without hole doping) to 1.08 and 1.21 at nhn_{h} = 0.5 and 1.0ee, respectively, which in turn increases TcT_{\rm c} up to 136 K at nhn_{h} = 1.0ee. It is thus likely that the increased DOS around EFE_{F} via hole doping increases the EPC channels, resulting in an increase of TcT_{\rm c}. We note that the hole doping with nhn_{h} << 1.2ee preserves structural stability without imaginary phonon frequencies (see Fig. S5 in the Supplemental Material [53]).

Although ordinary hole doping is achieved by the introduction of electron acceptor dopants in the host matrix, we here propose the hole doping of CeH9 using the substitution of Ce3+ ions for Ce atoms. The percentage of Ce3+ could change hole-doping concentrations to tune the DOS at EFE_{F}. Note that the hole doping of nhn_{h} = 1.0ee can be enabled by the substitution of 33% Ce3+. In order to examine the electronic miscibility of Ce3+ ions in the CeH9 matrix, we calculate the charge density of CeH9 without hole doping (see Fig. S6 in the Suppelemental Material [53]). Interestingly, we find that the total charge inside the Ce muffin-tin sphere with radius 1.40 Å is 9.55ee with including the 5s2s^{2}5p6p^{6} semicore electrons, close to that (9.63ee) obtained at nhn_{h} = 1.0ee. This nearly invariance of Ce charges between the two systems implies that both Ce and Ce3+ could behave similarly as cations without hole localization. Based on our results for the charge distribution and structural stability of hole doping nhn_{h} << 1.2ee, Ce3+ ions are most likely to be electronically miscible with Ce atoms in the CeH9 matrix. We note that in the present calculations, the hole charges are compensated by uniform background charge to maintain charge neutrality, as implemented in the VASP code [44, 45]. This simulation of hole doping is believed to properly describe the incorporated Ce3+ ions in the CeH9 matrix, because both Ce and Ce3+ with similar cation characters can be equally screened by their surrounding anionic H cages.

Finally, we also explore the hole-doping effect on SC in a recently observed [24, 25] rare earth hydride LaH10. As shown in Fig. 4(a), this hydride has a vHs near EFE_{F} with a strong hybridization of La 4ff and H 1ss orbitals [31], similar to the characteristic of vHs in CeH9 [see Fig. 1(b)]. The calculated λ{\lambda} and TcT_{\rm c} values of LaH10 are displayed as a function of nhn_{h} in Fig. 4(b). Since the DOS around EFE_{F} increases with hole doping [see Fig. 4(a)], λ{\lambda} increases monotonously with increasing nhn_{h}. Consequently, hole doped LaH10 raises TcT_{c} from 233 K (without hole doping) to 245 K at nhn_{h} = 0.3ee. Here, the hole doping-induced increase of TcT_{c} is 12 K, much smaller than the corresponding ΔTc{\Delta}T_{c} {\approx} 62 K in CeH9 [see Fig. 3(c)]. The relatively smaller value of ΔTc{\Delta}T_{c} in LaH10 is partly associated with the weak variation of DOS around EFE_{F} via hole doping [see Fig. 4(a)].

Refer to caption
Figure 4: (a) Calculated PDOS of LaH10, obtained without hole doping (left) and at nhn_{h} = 0.1ee (right). The unit of DOS is states/eV per unit cell that contains one La atom. In (b), the calculated λ{\lambda} and TcT_{c} values are displayed as a function of nhn_{h}.

In summary, based on first-principles calculations, we proposed that hole doping significantly enhances TcT_{c} in a recently synthesized [29, 30] hydride CeH9. It was revealed that hole doping induces the shift of EFE_{F} toward a vHs, thereby leading to the enhancement of EPC through an increased DOS around EFE_{F}. Interestingly, the vHs was found to be created by the band touching of two crystalline symmetry-protected DNLs whose electronic states are mostly composed of hybridized Ce 4ff and H 1ss orbitals. Therefore, the symmetry, band topology, and hole doping are cooperated to increase TcT_{c} of compressed CeH9. The proposed hole-doping effect on SC is rather generic and hence, it can also be applicable to another experimentally observed [24, 25] high-TcT_{c} rare earth hydride LaH10. We anticipate that future experimental work will be stimulated to adopt hole doping for rasing TcT_{c} in high-pressure superconducting hydrides.

Acknowledgements. This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korean Government (Grants No. 2019R1A2C1002975, No. 2016K1A4A3914691, and No. 2015M3D1A1070609). The calculations were performed by the KISTI Supercomputing Center through the Strategic Support Program (Program No. KSC-2020-CRE-0163) for the supercomputing application research.

C.W. and S.L. contributed equally to this work.

Corresponding author: [email protected]

References

  • [1] H. J. Queisser and E. E. Haller, Science 281, 945 (1998).
  • [2] Y. Tokura, H. Takagi, and S. Uchida, Nature (London) 337, 345 (1989).
  • [3] Y. Cao, V. Fatemi, S. Fang, K. Watanabe, T. Taniguchi, E. Kaxiras, and P. Jarillo-Herrero, Nature (London) 556, 34 (2018).
  • [4] P. A. Lee, N. Nagaosa, and X.-G. Wen, Rev. Mod. Phys. 78, 17 (2006).
  • [5] B. Keimer, S. A. Kivelson, M. R. Norman, S. Uchida, J. Zaanen, Nature (London) 518, 179 (2015).
  • [6] G. R. Stewart, Rev. Mod. Phys. 83, 1589 (2011).
  • [7] P. Dai, Rev. Mod. Phys. 87, 855 (2015).
  • [8] J. G. Bednorz and K. A. Müller, Z. Phys. B Condens. Matter 64, 189 (1986).
  • [9] Y. Kamihara, H. Hiramatsu, M. Hirano, R. Kawamura, H. Yanagi, T. Kamiya, and H. Hosono, J. Am. Chem. Soc. 128, 10012 (2006).
  • [10] Y. Kamihara, T. Watanabe, M. Hirano, and H. Hosono, J. Am. Chem. Soc. 130, 3296 (2008).
  • [11] J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Phys. Rev. 106, 162 (1957).
  • [12] J. A. Flores-Livas, L. Boeri, A. Sanna, G. Profeta, R. Arita, and M. Eremets, Phys. Rep. 856, 1 (2020), and references therein.
  • [13] E. Zurek, R. Hoffmann, N. W. Ashcroft, A. R. Oganov, and A. O. Lyakhov, Proc. Natl. Acad. Sci. USA 106, 42 (2009).
  • [14] J. Hooper and E. Zurek, J. Phys. Chem. C 116 13322 (2012).
  • [15] H. Wang, J. S. Tse, K. Tanaka, T. Litaka, and Y. Ma, Proc. Natl. Acad. Sci. USA 109, 6463 (2012).
  • [16] Y. Xie, Q. Li, A. R. Oganov, and H. Wang, Acta Cryst. C70 104 (2014).
  • [17] D. Duan, Y. Liu, F. Tian, D. Li, X. Huang, Z. Zhao, H. Yu, B. Liu, W. Tian, and T. Cui, Sci. Rep. 4, 6968 (2014).
  • [18] X. Feng, J. Zhang, G. Gao, H. Liu, and H. Wang, RSC Adv. 5, 59292 (2015).
  • [19] F. Peng, Y. Sun, C. J. Pickard, R. J. Needs, Q. Wu, and Y. Ma, Phys. Rev. Lett. 119, 107001 (2017).
  • [20] H. Liu, I. I. Naumov, R. Hoffmann, N. W. Ashcroft, and R. J. Hemley, Proc. Natl. Acad. Sci. USA 114, 6990 (2017).
  • [21] Y. Sun, J. Lv, Y. Xie, H. Liu, and Y. Ma, Phys. Rev. Lett. 123, 097001 (2019).
  • [22] H. Xie, Y. Yao, X. Feng, D. Duan, H. Song, Z. Zhang, S. Jiang, S. A. T. Redfern, V. Z. Kresin, C. J. Pickard, and T. Cui, Phys. Rev. Lett. 125, 217001 (2020).
  • [23] A. P. Drozdov, M. I. Eremets, I. A. Troyan, V. Ksenofontov, and S. I. Shylin, Nature (London) 525, 73 (2015).
  • [24] 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).
  • [25] 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 (London) 569, 528 (2019).
  • [26] E. Snider, N. Dasenbrock-Gammon, R. McBride, M. Debessai, H. Vindana, K. Vencatasamy, K. V. Lawler, A. Salamat, and R. P. Dias, Nature (London) 586, 373 (2020).
  • [27] W. A. Bassett, High Press. Res. 29, 163 (2009).
  • [28] H. K. Mao, X. J. Chen, Y. Ding, B. Li, and L. Wang, Rev. Mod. Phys. 90, 015007 (2018).
  • [29] X. Li, X. Huang, D. Duan, C. J. Pickard, D. Zhou, H. Xie, Q. Zhuang, Y. Huang, Q. Zhou, B. Liu, and T. Cui, Nat. Commun. 10, 3461 (2019).
  • [30] 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, Nat. Commun. 10, 4453 (2019).
  • [31] L. Liu, C. Wang, S. Yi, K. W. Kim, J. Kim, and J.-H. Cho, Phys. Rev. B 99, 140501(R) (2019).
  • [32] C. Wang, S. Yi, and J.-H. Cho, Phys. Rev. B 100, 060502(R) (2019).
  • [33] C. Wang, S. Yi, and J.-H. Cho, Phys. Rev. B 101, 104506 (2020).
  • [34] H. Jeon, C. Wang, S. Yi, and J.-H. Cho, Sci. Rep. 10, 16878 (2020).
  • [35] D. A. Papaconstantopoulos, M. J. Mehl, and P.-H. Chang, Phys. Rev. B 101, 060506(R) (2020).
  • [36] N. P. Armitage, E. J. Mele, and A. Vishwanath, Rev. Mod. Phys. 90, 015001 (2018), and references therein.
  • [37] S. Liu, C. Wang, L. Liu, J.-H. Choi, J.-H. Kim, Y. Jia, C. H. Park, and J.-H. Cho, Phys. Rev. Lett. 125, 187203 (2020).
  • [38] Y. Kim, B. J. Wieder, C. L. Kane, and A. M. Rappe, Phys. Rev. Lett. 115, 036806 (2015).
  • [39] C. Fang, Y. Chen, H.-Y. Kee, and L. Fu, Phys. Rev. B 92, 081201(R) (2015).
  • [40] G. Bian, T.-R. Chang, R. Sankar, S.-Y. Xu, H. Zheng, T. Neupert, C.-K. Chiu, S.-M. Huang, G. Chang, I. Belopolski, et al. Nat. Commun. 7 10556 (2016).
  • [41] C. Fang, H. Weng, X. Dai, and Z. Fang, Chinese Phys. B 25, 117106 (2016).
  • [42] Q.-F. Liang, J. Zhou, R. Yu, Z. Wang, and H. Weng, Phys. Rev. B 93, 085427 (2016).
  • [43] Our DFT calculations were performed using the Vienna ab initio simulation package with the projector-augmented wave method [44, 45, 46]. For the exchange-correlation energy, we employed the generalized-gradient approximation functional of Perdew-Burke-Ernzerhof (PBE) [47]. The 5s2s^{2}5p6p^{6} semicore electrons of Ce atom were included in the electronic-structure calculations. A plane-wave basis was used with a kinetic energy cutoff of 1000 eV. The 𝐤{\bf k}-space integration was done with the 18×{\times}18×{\times}12 kk points for the structure optimization and the 48×{\times}48×{\times}32 kk points for the DOS calculation. All atoms were allowed to relax along the calculated forces until all the residual force components were less than 0.001 eV/Å. The phonon spectrum and EPC calculations were carried out by using the QUANTUM ESPRESSO (QE) package [48] with the norm-conserved Hartwigsen-Goedecker-Hutter pseudopotentials [49], a kinetic energy cutoff of 1224 eV, and the 3×{\times}3×{\times}2 qq and 18×{\times}18×{\times}12 kk points for the computation of phonon frequencies. For the calculation of EPC, we used the software EPW [50, 51] with the 24×{\times}24×{\times}16 qq and 48×{\times}48×{\times}32 kk points. We also performed the DFT + UU calculation using the Dudarev’s approach [52], with the Hubbard parameter UU = 4.5 eV and exchange interaction parameter JJ = 0.5 eV. The DFT + UU bands for the hybridized Ce-4ff and H-1ss states around EFE_{F} are nearly the same as the corresponding DFT bands (see Fig. S7 in the Supplemental Material [53]). It is thus likely that the inclusion of UU would little change TcT_{c} of CeH9.
  • [44] G. Kresse and J. Hafner, Phys. Rev. B 48, 13115 (1993).
  • [45] G. Kresse and J. Furthmüller, Comput. Mater. Sci. 6, 15 (1996).
  • [46] P. E. Blöchl, Phys. Rev. B 50, 17953 (1994).
  • [47] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996); 78, 1396 (1997).
  • [48] P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni, I. Dabo et al., J. Phys.: Condens. Matter 21, 395502 (2009).
  • [49] C. Hartwigsen, S. Goedecker, and J. Hutter, Phys. Rev. B 58, 3641 (1998).
  • [50] F. Giustino, M. L. Cohen, and S. G. Louie, Phys. Rev. B 76, 165108(R) (2007).
  • [51] S. Poncé, E. R. Margine, C. Verdi, and F. Giustino, Comput. Phys. Commun. 209, 116 (2016).
  • [52] S. L. Dudarev, G. A. Botton, S. Y. Savrasov, C. J. Humphreys, and A. P. Sutton, Phys. Rev. B 57, 1505 (1998).
  • [53] The Supplemental Material at http://link.aps.org/supplemental/xxxxx for the symmetry and topology analysis, phonon frequencies, band projections onto the orbitals of Ce and H atoms, charge density plot, and band structures computed with including SOC and UU.
  • [54] A. A. Mostofi, J. R. Yates, Y.-S. Lee, I. Souza, D. Vanderbilt, and N. Marzari, Comput. Phys. Commun. 178, 685 (2008).
  • [55] Q. S. Wu, S. N. Zhang, H.-F. Song, M. Troyer, and A. A. Soluyanov, Comput. Phys. Commun. 224, 405 (2018).
  • [56] N. F.Q. Yuan, H. Isobe, and L. Fu, Nat. Commun. 10, 5769 (2019).
  • [57] A. B. Migdal, Sov. Phys. JETP 34, 996 (1958).
  • [58] G. M. Eliashberg, Sov. Phys. JETP 11, 696 (1960).
  • [59] P. B. Allen and B. Mitrović, Solid State Phys. 37, 1 (1982).
  • [60] The band dispersion of electronic states around EFE_{F} shows little change with respec to the absence [see Fig. 1(b)] and presence (see Fig. S3 in the Supplemental Material [53]) of SOC. Therefore, we can say that SOC hardly affects TcT_{c} in CeH9.