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

Revelation of Mott insulating state in layered honeycomb lattice Li2RuO3

Sakshi Bansal, Asif Ali, B. H. Reddy and Ravi Shankar Singh rssingh@iiserb.ac.in Department of Physics, Indian Institute of Science Education and Research Bhopal, Bhopal Bypass Road, Bhauri, Bhopal - 462 066, INDIA
Abstract

We investigate the role of electron correlation in the electronic structure of honeycomb lattice Li2RuO3 using photoemission spectroscopy and band structure calculations. Monoclinic Li2RuO3 having Ru network as honeycomb lattice undergoes magneto-structural transition at Tc \sim 540 K from high temperature phase C2/mC2/m to low temperature dimerized phase P21/mP2_{1}/m. Room temperature valence band photoemission spectra reveal an insulating ground state with no intensity at Fermi level (EFE_{F}). Ru 4dd band extracted from high and low photon energy valence band photoemission spectra reveal that the surface and bulk electronic structures are very similar in this system. Band structure calculations using generalized gradient approximation (GGA) leads to metallic ground state while screened hybrid (YS-PBE0) functional reveals opening up of a gap in almost degenerate dzxd_{zx}/dyzd_{yz} orbital, whereas dxyd_{xy} orbital is already gapped. Ru 3dd core level spectra with prominent unscreened feature provides direct evidence of strong electron correlation among Ru 4dd electrons which is also manifested by |EEF|2|E-E_{F}|^{2} dependence of spectral density of states (DOS) in the vicinity of EFE_{F} in the high-resolution spectra, establishing Li2RuO3 as Mott insulator.

preprint: preprint

I INTRODUCTION

Quasi two dimensional transition metal oxides (TMOs) with layered honeycomb lattice have recently been topic of great interest due to plethora of novel phases in these compounds K-H-model ; QSHE-Ir ; yogesh-Ir ; Miura-struc ; Miura-MO ; VBLiquid and their potential applications Li-ion1 ; Li-ion2 . In these systems, the transition metal at the center of the octahedra forms a layered honeycomb network. Among these, honeycomb iridates AA2IrO3 (AA = Na, Li) have received considerable interest due to anisotropic exchange interaction between spin and orbital moments in the strong spin-orbit coupling (SOC) limit, leading to novel ground states K-H-model ; QSHE-Ir ; yogesh-Ir . Ru doped Li2IrO3 demonstrates that models based on the assumption of only SOC do not describe this system properly Ir-Ru . In addition to SOC, on-site electron correlation (UU) among dd electrons is also expected to play an important role, as found in relativistic Mott insulator Na2IrO3 NIO-j1/2 as well as in many other transition metal oxides BJ-j1/2 ; cairo3 ; mgir-znir ; rss-STIO . On-site electron correlation and its role in deciding the novel ground state properties have been extensively studied in TMOs MIT where the systems can be defined by a single parameter UU/WW (WW = width of the dd band) within the Mott-Hubbard model and the Mott criteria UU/WW \sim 1 separates the Mott insulators (UU/WW ¿ 1) and correlated metals (UU/WW ¡ 1) Mott ; Hubbard . 3dd TMOs fall in the strongly correlated regime due to the narrow 3dd band. Since, UU inversely depends on the extension of the radial wave function of the dd orbital, correlation effects are expected to be less important in wide band 4dd and further in 5dd TMOs RuO2 ; casrruo3 ; IrO2 ; bairo3 . Surprisingly, moderate to large correlation effect has been realized in various 4dd and 5dd systems exhibiting bad metallic to Mott insulating ground states y2ru2o7 ; ca2ruo4 ; ru-core ; Ca2FeReO6 ; rss-YIO ; sr3iro7 .

4dd TMOs, particularly Ru-based oxides, exhibit a plethora of interesting properties such as superconductivity, non-Fermi liquid behaviour, unusual magnetic ground states, etc., while exhibiting varying electron correlation strength Sr2RuO4 ; kalo-epl ; rana-sci-rep . 4dd analogue of the honeycomb iridates is Li2RuO3 which crystallizes in monoclinic structure with Ru network forming honeycomb lattice. It undergoes first order magneto-structural transition at \sim 540 K from high temperature phase C2/mC2/m to low temperature dimerized phase P21/mP2_{1}/m accompanied by a steep increase in resistivity (insulating to highly insulating behaviour) and a steep decrease in magnetic susceptibility with the decrease in temperature Miura-struc . In the dimerized phase, strong disproportionation of about 19% between short and long Ru-Ru bonds has been observed at room temperature. Dimerization occurs due to strong overlap between one pair of Ru 4dd orbitals resulting into formation of molecular orbitals triggering the structural transition below 540 K Miura-struc ; Miura-MO ; yogesh-Ru . Li2RuO3, with partially filled t2g4t_{2g}^{4} band, is expected to be metallic. In contrast, transport measurements reveal it to be highly insulating suggests correlation effects may also be important, as also emphasized in cluster dynamical mean field theory (LDA+cDMFT) calculations dmft .

In this letter, we investigate the role of electron correlation in the electronic structure of dimerized Li2RuO3 using photoemission spectroscopy and band structure calculations. Valence band collected using xx-ray and ultra-violet photoemission spectroscopy reveal that the surface and bulk electronic structures are very similar. Ru 4dd band exhibits a broad peak at around 1 eV binding energy along with a shoulder at higher binding energy. No intensity at the Fermi level confirms the insulating ground state. Band structure calculations using YS-PBE0 functional opens up a gap in correlated dyzd_{yz}/dzxd_{zx} orbitals whereas dxyd_{xy} orbital is already gapped in GGA calculation, revealing the orbital dependency of correlation effect in this system. Prominent unscreened features in Ru 3dd core level spectra and parabolic energy dependence of spectral DOS near EFE_{F} suggest the influence of strong electron correlation among Ru 4dd electrons.

II EXPERIMENTAL AND CALCULATION DETAILS

Polycrystalline samples of Li2RuO3 were synthesized by solid-state reaction method using high purity ingredients Li2CO3 (99.995%) and RuO2 (99.99%). Stoichiometric amounts of ingredients were mixed, and the well ground mixture was pressed into pellets and calcined at 700 oC for 12 hours. To achieve large grain size, samples in highly pressed pellet form was sintered at 1000 oC for 6 days with two intermediate grindings. Samples were furnace cooled at the end of each heat treatment. A 10% excess of Li2CO3 was used to compensate for the evaporation of Li during heat treatment. The phase purity was confirmed by powder xx-ray diffraction (XRD) pattern collected at room temperature using PANalytical X’Pert Pro diffractometer equipped with Cu KαK\alpha radiation (λ\lambda = 1.540 Å). Rietveld refinement of xx-ray diffraction pattern was performed using the HighScore Plus software. XRD pattern shows no secondary phase including the absence of any peak corresponding to RuO2. The Rietveld refinement reveals that the crystal structure is monoclinic with lattice parameters aa = 4.933 Å, bb = 8.778 Å, cc = 5.894 Å, and β\beta = 124.41o and crystallizes in P21/mP2_{1}/m space group which are in excellent agreement with earlier reports Miura-struc . Photoemission spectra were collected at room temperature on in-situ (base pressure \sim 4 ×\times 10-11 mbar) fractured samples using spectrometer equipped with a Scienta-R4000 electron energy analyzer with a total instrumental resolution of 400 meV, 8 meV and 5 meV for monochromatic Al KαK\alpha (1486.6 eV), He II (40.8 eV) and He I (21.2 eV) photons (energy), respectively. Multiple samples were fractured to ensure the reproducibility of data and cleanliness of the surface was ensured by the negligibly small feature at higher binding energy in O 1ss core level spectra and absence of feature corresponding to C 1ss core level. Polycrystalline silver was used to determine EFE_{F} and the energy resolutions for different photon energies at 30 K.

Band structure calculations were performed using full-potential linearized augmented plane wave (FPLAPW) method as implemented in Wien2k for experimentally found structural parameters consisting of 4 formula units (fu) in the unit cell wien2k . 17×\times8×\times14 kk mesh within the first Brillouin zone and GGA GGA were used to calculate the DOS. To compare with the experimental valence band and to take care of the electron correlation in the band structure calculations, we also performed the calculations using YS-PBE0 functional YS-PBE0 and DOS was calculated for 10×\times4×\times8 kk mesh within the first Brillouin zone. The local coordinate system for Ru atom, where xx and yy axis points to the oxygen atoms on the common edge shared by two RuO6 octahedra (forming the Ru-Ru dimer), is chosen for obtaining orbital resolved DOS for Ru 4dd states. For all the calculations, the energy and charge convergence was better than 0.1 meV and 10-4 electronic charge, respectively.

Refer to caption
Figure 1: O 2pp and Ru 4dd PDOS for Li2RuO3 from (a) GGA calculations and (b) YS-PBE0 calculations. Insets in respective figures show different dd orbital contributions in t2gt_{2g} band in the local coordinate system (see text).
Refer to caption
Figure 2: (a) XP, He II and He I valence band spectra of Li2RuO3 at room temperature. Lines show the O 2pp band contribution. (b) Total DOS for Li2RuO3 using YS-PBE0 calculations.

III RESULTS AND DISCUSSION

In Fig. 1(a), we show the results of GGA calculations. Grey shaded area shows the O 2pp partial density of states (PDOS), and Ru 4dd PDOS has been shown with black line. PDOS corresponding to Li has not been shown here due to negligibly small contribution; thus, the valence band is formed by the hybridization of O 2pp with Ru 4dd states only. Three distinct groups of features are evident in the occupied part. Nonbonding O 2pp states appear between 2–5.5 eV, and the bonding states with dominant O 2pp contributions appear at higher binding energies (centered at \sim6 eV). The feature below 2 eV binding energy are antibonding states primarily having Ru 4dd character. In local coordinate system, ege_{g} band, comprising of dz2d_{z}^{2} and dx2y2d_{x^{2}-y^{2}} orbitals, is completely empty and appears below -2 eV binding energy while t2gt_{2g} band contribute in the vicinity of EFE_{F} appearing between 2 to -1 eV binding energy. Total width of t2gt_{2g} states is about 2.7 eV and of ege_{g} states is about 1.5 eV. A dip like structure is observed, but no hard gap is found at EFE_{F}. The inset of Fig. 1(a) shows various dd orbitals of t2gt_{2g} band. It is to note here that dxyd_{xy} orbital is gapped and almost degenerate dyzd_{yz} and dzxd_{zx} orbitals are partially filled. The ideal Ru-honeycomb lattice (with perfect RuO6 octahedra) exhibits C3C_{3} symmetry, where dxyd_{xy}, dyzd_{yz} and dzxd_{zx} orbitals are expected to be perfectly degenerate. While in case of distorted honeycomb lattice (low temperature dimerized phase), one of the Ru-Ru bond length has reduced, thereby breaking the C3C_{3} symmetry and stronger direct overlap of the orbital along the dimer (dxyd_{xy} in the present case) leads to the formation of molecular orbital Miura-MO ; dmft .

Fig. 1(b) shows the PDOS corresponding to O 2pp and Ru 4dd states from band structure calculations using YS-PBE0 functional. Screened hybrid functional have been shown to be very successful in capturing the band gap in correlated systems where the ground states are often found to be semiconducting or insulating YS-PBE0 ; YS-PBE0-1 ; CSPO . These results exhibit an insulating ground state with the gap opening in the t2gt_{2g} band; thus the EFE_{F} has been set in the middle of the gap. This Mott gap has also been observed in the calculation using TB-mBJ anisotropy and LDA+cDMFT dmft calculations. The value of the energy gap of 0.84 eV is somewhat larger than other calculations anisotropy and as well as the activation gap (\sim0.3 eV) from transport measurements gap . The O 2pp band moves towards higher binding energy (\sim0.5 eV) in comparison to GGA results. The inset of Fig. 1(b) shows orbital resolved t2gt_{2g} band. Interestingly, strong electron correlation leads to opening of the gap in partially filled dyzd_{yz} and dzxd_{zx} orbitals having a small effect on dxyd_{xy} orbital (already gapped in GGA), as also seen in the LDA+cDMFT results dmft . These results manifest the orbital selectivity of the correlation effects in Li2RuO3 where dyzd_{yz} and dzxd_{zx} orbitals are strongly correlated atomic orbitals and dxyd_{xy} forms the molecular orbital. Single site DMFT (LDA+sDMFT) calculations required quite large value of U/WU/W (\sim 2.2) to exhibit the Mott gap in honeycomb lattice jafari which later was shown to be U/WU/W \sim 1 using LDA+cDMFT calculations. While the LDA+cDMFT has also been successful in describing the ground state of dimerized VO2 VO2 , it is interesting to observe the Mott gap along with the orbital selectivity of correlation effects in dimerized Li2RuO3 using screened hybrid calculations.

Room temperature valence band spectra obtained using 1486.6 eV (XP spectra), 40.8 eV (He II spectra) and 21.2 eV (He I spectra) excitation energies are shown in Fig. 2(a). Similar to GGA and YS-PBE0 results, three discernible features A, B, and C are observed in the valence band spectra. The features B and C corresponding to O 2pp states are enhanced in the low excitation energy spectra, while feature A corresponding to Ru 4dd states is enhanced in XP spectra. This is due to the strong dependence of the relative transition matrix elements on excitation energies. The ratio of photoemission cross-section of Ru 4dd states to O 2pp states is significantly higher in XP spectra compared to that in He II and He I spectra yeh .

The overall comparison of experimental spectra with GGA results in Fig. 1(a) suggests that a rigid shift of O 2pp band from GGA results (\sim0.5 eV towards higher binding energy) is required to match the observed peak position and widths of features B and C of the experimental spectra. Such a shift of completely filled O 2pp bands is often observed due to underestimation of correlation effects in the band structure calculations dd .To compare the experimental results and to take care of the electron correlations in the band structure calculations we show the occupied part of the total DOS obtained using YS-PBE0 functional in Fig. 2(b). The total width of the O 2pp band as well as positions of the features A, B, and C are remarkably similar to the experimental spectra.

Refer to caption
Figure 3: (a) Ru 4dd band extracted from XP, He II and He I valence band spectra. The resolution broadened He II spectra is shown by line. (b) Spectral DOS corresponding to He I spectra. |EEF|2|E-E_{F}|^{2} fit is shown by line.
Refer to caption
Figure 4: Ru 3dd core level spectra of Li2RuO3 at room temperature. Red and blue curves are corresponding to RuO2 RuO2 and Y2Ru2O7 y2ru2o7 respectively.

It is clear from Fig. 2(a) that the O 2pp and Ru 4dd bands are distinctly separated in high and low excitation energy spectra. Thus, Ru 4dd contributions can reliably be delineated by fitting O 2pp bands using two Gaussians representing feature B and C as performed in other systems casrruo3 ; bairo3 ; rss-YIO ; rss-STIO . Lines in Fig. 2(a) show the resultant fit obtained by the least-squares error method and the extracted Ru 4dd band is shown in Fig. 3(a). All the spectra, normalized by the total integrated intensity, show the main feature at around 1 eV binding energy, with a shoulder structure appearing between 1.5-2 eV binding energy. No intensity at the Fermi level in all the spectra suggests insulating character. Interestingly, the spectral line shape is very similar in both high excitation energy XP spectra as well as low excitation energy (He I and He II) spectra, despite having large difference in probing depth. The major difference in the low and high excitation energy spectra arises due to the energy resolution; thus, the resolution broadened (Gaussian broadening of 0.4 eV) He II spectra (shown by line) is compared with XP spectra. The very similar lineshape of these two spectra establishes that the surface and bulk electronic structures are essentially similar in contrast to the observations in other 4dd and 5dd transition metal oxides casrruo3 ; rss-YIO . The spectral DOS can be obtained by dividing photoemission spectra with Fermi Dirac distribution function since the hole and electron lifetime broadening around Fermi level and energy broadening due to high resolution can be neglected. Thus obtained spectral DOS corresponding to He I spectra has been shown in Fig. 3(b). The line shows |EEF|2|E-E_{F}|^{2} behavior of the spectral DOS in the vicinity of EFE_{F}.

It is to note here that lithium deficiency in Li2RuO3 (hole doping) would lead to larger DOS at EFE_{F}, which is reflected as well-screened feature in the Ru 3dd core level spectra sakshi-AIP . Increased DOS at EFE_{F} is also observed in specific heat measurements for highly disordered Li2RuO3 gap . The value of Sommerfeld coefficient γ\gamma, for the least disordered Li2RuO3 is similar to that of the band insulating Li2TiO3, suggests negligible DOS gap as observed here in the high resolution He I spectra. The parabolic energy dependence of spectral DOS manifests strong correlation induced soft Coulomb gap at EFE_{F} in this finitely disordered system rss-STIO ; rss-YIO ; Efros .

In Fig. 4, we show Ru 3dd core-level spectra of Li2RuO3 collected at room temperature. The core level spectra exhibit spin-orbit split peaks at 282.2 eV and 286.3 eV corresponding to Ru 3d5/2d_{5/2} and Ru 3d3/2d_{3/2} with the spin-orbit splitting of about 4.1 eV. Absence of any signal at \sim280 eV binding energy in the present case of in-situ fractured sample confirms the high quality sample (with least disorder/lithium deficiency) sakshi-AIP . For reference, we also show the Ru 3dd core level spectra for RuO2 and Y2Ru2O7. Ru 3dd core level spectra, in various ruthenates having Ru in 4+ valence state, exhibit two peak structures ru-core for each spin-orbit split peak as seen in case of metallic RuO2 RuO2 . The peak appearing at around 281 eV and 285 eV correspond to well-screened features while weak and broad features at around 282.5 eV and 286.5 eV correspond to poorly-screened features. These features are associated with the final state effects where a core hole generated in the photoemission process are screened by the electrons at the Fermi level. In contrast to metallic RuO2, Mott insulating Y2Ru2O7 only exhibits broad peaks corresponding to unscreened features since the valence electrons are localized in the presence of strong electron correlation leading to the disappearance of screened feature y2ru2o7 . The spectral features with respect to peak position and broadening in the case of Li2RuO3 are very similar to Y2Ru2O7, confirming the strong correlation and thus the Mott insulating state in the present system.

IV CONCLUSION

In conclusion, we have investigated the electronic structure of layered honeycomb lattice Li2RuO3 to understand the role of electron correlation on the electronic structure. Room temperature valence band spectra suggest an insulating ground state with no intensity at EFE_{F}. The line shape of Ru 4dd band extracted from high and low energy photoemission spectra, having different probing depths, suggests surface and bulk electronic structures are very similar in this system. The influence of electron correlation is manifested by parabolic energy dependence of spectral DOS in the vicinity of EFE_{F}. Band structure calculations using GGA and screened hybrid functional reveals orbital selective Mott state in this system. Ru 3dd core level spectra with prominent unscreened feature provide direct evidence for strong electron correlation among 4dd electrons.

ACKNOWLEDGEMENTS

We thank R. Kewat and N. Ganguly for the fruitful discussion. Authors acknowledge the support of Central Instrumentation Facility and HPC Facility at IISER Bhopal. Support from DST-FIST (Project No. SR/FST/PSI-195/2014(C)) is also thankfully acknowledged.

Present address: Department of Physics, Government College (A), Rajamundry – 533105, A. P., India

References

  • (1) J. Chaloupka, G. Jackeli, and G. Khaliullin, Phys. Rev. Lett. 105, 027204 (2010).
  • (2) A. Shitade, H. Katsura, J. Kuneš, X.-L. Qi, S.-C. Zhang, and N. Nagaosa, Phys. Rev. Lett. 102, 256403 (2009).
  • (3) Y. Singh, S. Manni, J. Reuther, T. Berlijn, R. Thomale, W. Ku, S. Trebst, and P. Gegenwart, Phys. Rev. Lett. 108, 127203 (2012); Y. Singh, and P. Gegenwart, Phys. Rev. B 82, 064412 (2010).
  • (4) Y. Miura, Y. Yasui, M. Sato, N. Igawa, and K. Kakurai, J. Phys. Soc. Jpn. 76, 033705 (2007).
  • (5) Y. Miura, M. Sato, Y. Yamakawa, T. Habaguchi, and Y. Ōno, J. Phys. Soc. Jpn. 78, 094706 (2009).
  • (6) S. A. J. Kimber, I. I. Mazin, J. Shen, H. O. Jeschke, S. V. Streltsov, D. N. Argyriou, R. Valentí, and D. I. Khomskii, Phys. Rev. B, 89, 081408(R) (2014).
  • (7) R. Marom, S. F. Amalraj, N. Leifer, D. Jacob, and D. Aurbach, J. Mater. Chem 21, 9938 (2011).
  • (8) M. M. Thackeray, S.-H. Kang, C. S. Johnson, J. T. Vaughey, R. Benedek, and S. A. Hackney, J. Mater. Chem. 15, 2257 (2005).
  • (9) H. Lei, W.-G. Yin, Z. Zhong, and H. Hosono, Phys. Rev. B 89, 020409(R) (2014).
  • (10) R. Comin, G. Levy, B. Ludbrook, Z.-H. Zhu, C. N. Veenstra, J. A. Rosen, Y. Singh, P. Gegenwart, D. Stricker, J. N. Hancock, D. van der Marel, I. S. Elfimov, and A. Damascelli, Phys. Rev. Lett. 109, 266406 (2012).
  • (11) B. J. Kim, H. Jin, S. J. Moon, J.-Y. Kim, B.-G. Park, C. S. Leem, J. Yu, T. W. Noh, C. Kim, S.-J. Oh, J.-H. Park, V. Durairaj, G. Cao, and E. Rotenberg, Phys. Rev. Lett. 101, 076402 (2008).
  • (12) J. Fujioka, R. Yamada, M. Kawamura, S. Sakai, M. Hirayama, R. Arita, T. Okawa, D. Hashizume, M. Hoshino, and Y. Tokura., Nat. Commun. 10, 362 (2019).
  • (13) G. Cao, A. Subedi, S. Calder, J.-Q. Yan, J. Yi, Z. Gai, L. Poudel, D. J. Singh, M. D. Lumsden, A. D. Christianson, B. C. Sales, and D. Mandrus, Phys. Rev. B 87, 155136 (2013).
  • (14) B. H. Reddy, A. Ali, and R. S. Singh, Europhys. Lett. 127, 47003 (2019).
  • (15) M. Imada, A. Fujimori, and Y. Tokura, Rev. Mod. Phys. 70, 1039 (1998).
  • (16) N. F. Mott Metal-Insulator Transitions, 2nd ed. (Taylor & Francis, London, 1990).
  • (17) J. Hubbard, Proc. R. Soc. A 276, 238 (1963).
  • (18) Y. J. Kim, Y. Gao, and S. A.Chambers, Appl. Surf. Sci. 120, 250 (1997).
  • (19) K. Maiti and R. S. Singh, Phys. Rev. B 71, 161102(R) (2005).
  • (20) J. M. Kahk, C. G. Poll, F. E. Oropeza, J. M. Ablett, D. Céolin, J.-P. Rueff, S. Agrestini, Y. Utsumi, K. D. Tsuei, Y. F. Liao, F. Borgatti, G. Panaccione, A. Regoutz, R. G. Egdell, B. J. Morgan, D. O. Scanlon, and D. J. Payne, Phys. Rev. Lett. 112, 117601 (2014).
  • (21) K. Maiti, R. S. Singh, V. R. R. Medicherla, S. Rayaprol, and E. V. Sampathkumaran, Phys. Rev. Lett. 95, 016404 (2005).
  • (22) P. A. Cox, R. G. Egdell, J. B. Goodenough, A. Hamnett, and C. C. Naish, J. Phys. C: Solid State Phys. 16, 6221 (1983).
  • (23) S. Nakatsuji, S. Ikeda, and Y. Maeno, J. Phys. Soc. Jpn. 66, 126404 (1997).
  • (24) H.-D. Kim, H.-J. Noh, K. H. Kim, and S.-J. Oh, Phys. Rev. Lett. 93, 126404 (2004).
  • (25) H. Iwasawa, T. Saitoh, Y. Yamashita, D. Ishii, H. Kato, N. Hamada, Y. Tokura, and D. D. Sarma, Phys. Rev. B 71, 075106 (2005).
  • (26) R. S. Singh, V. R. R. Medicherla, Kalobaran Maiti, and E. V. Sampathkumaran, Phys. Rev. B 77, 201102(R) (2008).
  • (27) G. Ahn, S. J. Song, T. Hogan, S. D. Wilson, and S. J. Moon, Sci. Rep. 6, 32632 (2016).
  • (28) Y. Maeno, H. Hashimoto, K. Yoshida, S. Nishizaki, T. Fujita, J. G. Bednorz, and F. Lichtenberg, Nature 372, 532–534 (1994).
  • (29) K. Maiti, R. S. Singh, and V. R. R. Medicherla, Europhys. Lett. 78, 17002 (2007); R. S. Singh, V. R. R. Medicherla, and K. Maiti, Appl. Phys. Lett. 91, 132503 (2007).
  • (30) S. Tripathi, R. Rana, S. Kumar, P. Pandey, R. S. Singh, and D. S. Rana, Sci. Rep. 4, 3877 (2014).
  • (31) K. Mehlawat and Y. Singh, Phys. Rev. B 95, 075105 (2017).
  • (32) Z. V. Pchelkina, A. L. Pitman, A. Moewes, E. Z. Kurmaev, T.-Y. Tan, D. C. Peets, J.-G. Park, and S. V. Streltsov, Phys. Rev. B 91, 115138 (2015).
  • (33) P. Blaha et al., 2001 WIEN2K, An Augmented Plane Wave + Local Orbitals Program for Calculating Crystal Properties (Technical Universitat Wien, Austria: Karlheinz Schwarz) ISBN 3-9501031-1-2; P. Blaha et al., J. Chem. Phys. 152, 074101 (2020).
  • (34) J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996).
  • (35) F. Tran and P. Blaha, Phys. Rev. B 83, 235118 (2011).
  • (36) F. Tran and P. Blaha, J. Phys. Chem. A 121, 3318 (2017).
  • (37) B. H. Reddy, A. Ali, and R. S. Singh, J. Phys.: Condens. Matt. 33, 185502 (2021).
  • (38) S. Yun, K. H. Lee, S. Y. Park, T.-Y. Tan, J. Park, S. Kang, D. I. Khomskii, Y. Jo, and J.-G. Park, Phys. Rev. B 100, 165119 (2019).
  • (39) J. Park, T.-Y. Tan, D. T. Adroja, A. Daoud-Aladine, S. Choi, D.-Y. Cho, S.-H. Lee, J. Kim, H. Sim, T. Morioka, H. Nojiri, V. V. Krishnamurthy, P. Manuel, M. R. Lees, S. V. Streltsov, D. I. Khomskii, and J.-G. Park, Sci. Rep. 6, 25238 (2016).
  • (40) S. A. Jafari, Eur. Phys. J. B 68, 537 (2009); M. Ebrahimkhas and S. A. Jafari, Europhys. Lett. 98, 27009 (2012).
  • (41) S. Biermann, A. Poteryaev, A. I. Lichtenstein, and A. Georges, Phys. Rev. Lett. 94, 026404 (2005).
  • (42) J. J. Yeh and I. Lindau, At. Data and Nucl. Data Tables 32, 1 (1985).
  • (43) D. D. Sarma, N. Shanthi, S. R. Barman, N. Hamada, H. Sawada, and K. Terakura, Phys. Rev. Lett. 75, 1126 (1995).
  • (44) S. Bansal, A. Ali, B. H. Reddy, and R. S. Singh, AIP Conf. Proc. 2265, 030385 (2020).
  • (45) A. F. Efros and B. I. Shklovskii, J. Phys. C 8, L49 (1975).; J. G. Massey and M. Lee, Phys. Rev. Lett. 75, 4266 (1995).