research letters\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

IUCrJ
Volume 6| Part 6| November 2019| Pages 990-995
ISSN: 2052-2525

R3c-type LnNiO3 (Ln = La, Ce, Nd, Pm, Gd, Tb, Dy, Ho, Er, Lu) half-metals with multiple Dirac cones: a potential class of advanced spintronic materials

CROSSMARK_Color_square_no_text.svg

aInstitute for Superconducting and Electronic Materials (ISEM), University of Wollongong, North Wollongong NSW 2500, Australia, bInstitute for Quantum Information and Spintronics (IQIS), School of Science, Chongqing University of Posts and Telecommunications, Chongqing 400065, People's Republic of China, cSchool of Physical Science and Technology, Southwest University, Chongqing 400715, People's Republic of China, dLaboratoire de Physique Quantique de la Matière et de Modélisation Mathématique (LPQ3M), Université de Mascara, Mascara 29000, Algeria, and eState Key Laboratory for Magnetism, Beijing National Laboratory for Condensed Matter Physics and Institute of Physics, Chinese Academy of Sciences, Beijing 100190, People's Republic of China
*Correspondence e-mail: cheng@uow.edu.au, yhk10@swu.edu.cn, yangtie@swu.edu.cn

Edited by A. N. Cormack, Alfred University, USA (Received 20 February 2019; accepted 9 September 2019; online 16 October 2019)

In the past three years, Dirac half-metals (DHMs) have attracted considerable attention and become a high-profile topic in spintronics becuase of their excellent physical properties such as 100% spin polarization and massless Dirac fermions. Two-dimensional DHMs proposed recently have not yet been experimentally synthesized and thus remain theoretical. As a result, their characteristics cannot be experimentally confirmed. In addition, many theoretically predicted Dirac materials have only a single cone, resulting in a nonlinear electromagnetic response with insufficient intensity and inadequate transport carrier efficiency near the Fermi level. Therefore, after several attempts, we have focused on a novel class of DHMs with multiple Dirac crossings to address the above limitations. In particular, we direct our attention to three-dimensional bulk materials. In this study, the discovery via first principles of an experimentally synthesized DHM LaNiO3 with many Dirac cones and complete spin polarization near the Fermi level is reported. It is also shown that the crystal structures of these materials are strongly correlated with their physical properties. The results indicate that many rhombohedral materials with the general formula LnNiO3 (Ln = La, Ce, Nd, Pm, Gd, Tb, Dy, Ho, Er, Lu) in the space group R3c are potential DHMs with multiple Dirac cones.

What are the key challenges lying ahead for the next generation of spintronics (Wang et al., 2016[Wang, X., Cheng, Z., Wang, J., Wang, X. & Liu, G. (2016). J. Mater. Chem. C. 4, 7176-7192.], 2018[Wang, X., Li, T., Cheng, Z., Wang, X. & Chen, H. (2018). Appl. Phys. Rev. 5, 041103.]; Wang, 2017[Wang, X. L. (2017). Natl Sci. Rev. 4, 252-257.])? This question relates to a key issue encountered in the current study – how to achieve ultra-fast transmission and zero-energy dissipation. To obtain a dissipation-free spin current, we found that the key to addressing this issue is identifying new materials with linear energy band crossings (Dirac cone features) and high spin polarization near the Fermi level. For this purpose, two types of materials with high-polarization, Dirac half-metals (DHMs) (Kan et al., 2011[Kan, E., Wu, F., Xiang, H., Yang, J. & Whangbo, M. (2011). J. Phys. Chem. C, 115, 17252-17254.]; Liu et al., 2017[Liu, Z., Liu, J. & Zhao, J. (2017). Nano Res. 10, 1972-1979.]; Li & Yang, 2017[Li, X. & Yang, J. (2017). WIREs Comput. Mol. Sci. 7, e1314.]; He et al., 2016[He, J., Ma, S., Lyu, P. & Nachtigall, P. (2016). J. Mater. Chem. C. 4, 2518-2526.]) and Dirac spin-gapless semiconductors (DSGSs) (He et al., 2017[He, J., Li, X., Lyu, P. & Nachtigall, P. (2017). Nanoscale, 9, 2246-2252.]; Wang et al., 2010[Wang, X. L., Dou, S. X. & Zhang, C. (2010). NPG Asia Mater. 2, 31.]; Wang, 2008[Wang, X. L. (2008). Phys. Rev. Lett. 100, 156404.]), were predicted. DHMs exhibit a novel Dirac cone in one spin direction and a semiconductor/insulator property in the other; thus, in theory, the entire material has a spin polarization of 100%. Unlike non-spin-polarized Dirac structures (Neto et al., 2009[Castro Neto, A. H., Guinea, F., Peres, N. M. R., Novoselov, K. S. & Geim, A. K. (2009). Rev. Mod. Phys. 81, 109-162.]; Zhang et al., 2017[Zhang, X., Jin, L., Dai, X. & Liu, G. (2017). J. Phys. Chem. Lett. 8, 4814-4819.]) such as graphene, DHMs can break the time reversal symmetry (TRS) in spin-resolved orbital physics because of the differences in the orders of charge current and spin current under TRS and the non-dissipative property of intrinsic coupling in DHMs. True DSGSs, which can be considered as extreme cases of DHMs, are relatively difficult to achieve. Therefore, the theoretical search for DHMs is critical for overcoming the current bottleneck in spintronics.

The concept of a DHM was first proposed in a triangular ferromagnet (Ishizuka & Motome, 2012[Ishizuka, H. & Motome, Y. (2012). Phys. Rev. Lett. 109, 237207.]). However, these materials are primarily limited to two-dimensional layered materials and heterojunction systems that have not yet been synthesized (Liu et al., 2017[Liu, Z., Liu, J. & Zhao, J. (2017). Nano Res. 10, 1972-1979.]) such as CrO2/TiO2 heterostructures and NiCl3 monolayers. Little progress was made until 2017, when Du's team discovered a novel category of three-dimensional DHMs with multiple Dirac cones (MDCs) based on MnF3 (Jiao et al., 2017[Jiao, Y., Ma, F., Zhang, C., Bell, J., Sanvito, S. & Du, A. (2017). Phys. Rev. Lett. 119, 016403.]), which have been realized experimentally. Based on first-principles calculations, Du and coworkers also investigated LaMnO3 (Ma et al., 2018[Ma, F., Jiao, Y., Jiang, Z. & Du, A. (2018). Appl. Mater. Interfaces, 10, 36088-36093.]) as a novel DHM and found that the material had multiple Dirac-like half-metallic properties.

Inspired by the above work, we extensively searched bulk materials for DHMs with MDCs and found that rhombohedral structures with the space group [R\bar 3 c] provide many potential DHMs for exploration. By comparing the DHMs identified in our search with MnF3, we discovered a series of materials with the general formula LnNiO3 that also has many Dirac cones near the Fermi level, similar to MnF3. Importantly, the LaNiO3 material with an [R\bar 3 c] structure has already been prepared by Sreedhar et al. (1992[Sreedhar, K., Honig, J. M., Darwin, M., McElfresh, M., Shand, P. M., Xu, J., Crooker, B. C. & Spalek, J. (1992). Phys. Rev. B, 46, 6382-6386.]).

All calculations in this manuscript were carried out in VASP code (Hafner, 2007[Hafner, J. (2007). Comput. Phys. Commun. 177, 6-13.]) using spin-polarized density functional theory (DFT). It should be noted that the GGA-PBE (Peverati & Truhlar, 2011[Peverati, R. & Truhlar, D. G. (2011). J. Phys. Chem. Lett. 2, 2810-2817.]) method has been widely used to predict the half-metallic and spin-gapless semiconducting properties of bulk materials (Wang et al., 2017a[Wang, X., Cheng, Z., Liu, G., Dai, X., Khenata, R., Wang, L. & Bouhemadou, A. (2017a). IUCrJ, 4, 758-768.]; Qin et al., 2017[Qin, G., Wu, W., Hu, S., Tao, Y., Yan, X., Jing, C., Li, X., Gu, H., Cao, S. & Ren, W. (2017). IUCrJ, 4, 506-511.]), and many spin-gapless semiconductors and half-metallic bulk materials predicted by GGA-PBE have been experimentally synthesized and confirmed (Ouardi et al., 2013[Ouardi, S., Fecher, G. H., Felser, C. & Kübler, J. (2013). Phys. Rev. Lett. 110, 100401.]; Bainsla et al.,2015[Bainsla, L., Mallick, A. I., Raja, M. M., Nigam, A. K., Varaprasad, B. S. D. C. S., Takahashi, Y. K., Alam, A., Suresh, K. G. & Hono, K. (2015). Phys. Rev. B, 91, 104408.]). More details about the methods of calculation can be found in the supporting information.

First, taking LaNiO3 as an example, we discuss its crystal structure and electronic structure in detail. The crystal structure of LaNiO3 is shown in Fig. S1 of the supporting information; a hexagonal structure with the space group [R\bar 3 c] (No. 167) which contains octahedrally coordinated metal centers with equal Ni—O bond lengths of 1.954 Å. The optimized equilibrium lattice parameters are a = b = 5.499 Å and c = 13.078 Å. The X-ray powder diffraction patterns of LaNiO3 have been studied by Sreedhar et al. (1992[Sreedhar, K., Honig, J. M., Darwin, M., McElfresh, M., Shand, P. M., Xu, J., Crooker, B. C. & Spalek, J. (1992). Phys. Rev. B, 46, 6382-6386.]) more than 20 years ago and the experimental lattice constants (a = b = 5.49 Å, c = 13.14 Å) of this material show a good qualitative agreement with the DFT results in the current study. The unit cell is composed of 6 La, 6 Ni and 18 O atoms. The total magnetic moment obtained through calculation was 6.236 µB, and the magnetic properties were mainly contributed by Ni atoms (see Table 1[link]). The magnetic contribution of each Ni atom (1.326 µB) was higher than that of each O and La atom. To further clarify the magnetic structure of LaNiO3, Fig. 1[link] shows the ferromagnetic (FM), antiferromagnetic (AFM) and nonmagnetic (NM) states of the 1 × 1 × 1 unit cell [Figs. 1[link](a)–1(d)]. The total energy selected under the FM magnetic structure was 0 meV. In the 1 × 1 × 1 unit cell, two AFM states were considered simultaneously [Figs. 1[link](c) and 1(d)]. The calculations revealed that the NM magnetic structure was the most unstable and the total energy of the two AFM states fell between the energies of the FM and NM states. This demonstrates that the FM structure of LaNiO3 is most stable in the case of the 1 × 1 × 1 unit cell.

Table 1
Optimized equilibrium lattice constants of LnNiO3 obtained using the GGA+U method and their total and atomic magnetic moments at their optimized equilibrium lattice constants

LnNiO3 a (Å) c (Å) Mtotal (μB) MLn (μB) MNi (μB) MO (μB)
LaNiO3 5.499 13.078 6.236 0.008 1.326 −0.093 /−0.103/−0.093
5.49 13.14
CeNiO3 5.538 13.127 6.250 0.008 1.344 −0.098/−0.109/−0.098
DyNiO3 5.435 12.648 6.263 0.016 1.308 −0.094/−0.095/−0.094
ErNiO3 5.410 12.612 6.263 0.017 1.300 −0.089/−0.092/−0.091
GdNiO3 5.439 13.068 6.298 0.015 1.358 −0.105/−0.108/−0.110
HoNiO3 5.411 12.378 6.323 0.017 1.373 −0.110/−0.114/−0.112
LuNiO3 5.499 13.078 6.341 0.017 1.395 −0.117/−0.120/−0.119
NdNiO3 5.511 12.868 6.244 0.010 1.328 −0.096/−0.102/−0.099
PmNiO3 5.449 13.058 6.262 0.011 1.341 −0.103/−0.107/−0.099
TbNiO3 5.447 12.671 6.261 0.015 1.311 −0.092/−0.096/−0.094
†From our work.
‡Experimental parameters (Sreedhar et al., 1992[Sreedhar, K., Honig, J. M., Darwin, M., McElfresh, M., Shand, P. M., Xu, J., Crooker, B. C. & Spalek, J. (1992). Phys. Rev. B, 46, 6382-6386.]).
[Figure 1]
Figure 1
Crystal structures of LaNiO3; the different magnetic structures including (a) FM, (b) NM, (c) AFM-I and (d) AFM-II are taken into consideration.

Figs. 2[link](a) and 2(b) show the energy band structures of LaNiO3 under equilibrium lattice parameters and the most stable magnetic structure. The high-symmetry points were located in the Brillouin zone, and we selected M–Γ–K–H–A–H. Normally, for some transition-metal (TM)-based materials with strongly correlated d electrons, the GGA+PBE method may not describe the electronic structure very well. Therefore, the GGA+U method (Anisimov et al., 1991[Anisimov, V. I., Zaanen, J. & Andersen, O. K. (1991). Phys. Rev. B, 44, 943-954.]) should be selected in this work, namely, U = 6.4 eV (Balachandran et al., 2018[Balachandran, P. V., Emery, A. A., Gubernatis, J. E., Lookman, T., Wolverton, C. & Zunger, A. (2018). Phys. Rev. Mater. 2, 043802.]) was added to the Ni-d orbital during the DFT calculation. As shown in Fig. 2[link](a), LaNiO3 showed multiple linear energy band dispersions (Dirac cones) near the Fermi level in the spin-up direction. Specifically, one Dirac cone was located at the A high-symmetry point, four Dirac cones were distributed along the M–Γ interval, two Dirac cones were distributed along the Γ–K direction and one Dirac cone was distributed along the A–H direction. It should be noted that more Dirac cones would have been discovered if we had considered all the high-symmetry points in the Brillouin zone. In general, all of the Dirac cones in [R\bar 3 c]-type LaNiO3 are protected by the D3d symmetry. A detailed band symmetry analysis can be seen in Table S1 of the supporting information. Obviously, one can see that the bands near the Fermi level belong to two-dimensional irreducible representation E.

[Figure 2]
Figure 2
Calculated band structures of LaNiO3 using (a) GGA+U and (b) GGA methods at its optimized equilibrium lattice constants. For (c), the experimental lattice constants are selected during the electronic structure calculations.

Compared with DHMs with one Dirac cone, the multiple Dirac energy band dispersions are expected to result in a stronger nonlinear electromagnetic response and a higher efficiency of carrier transport (because multiple Dirac channels exist at the Fermi level). These results of spin-up channel are similar to those reported by Du and coworkers, who also demonstrated the existence of MDCs in LaCuO3 (Zhang et al., 2018[Zhang, C., Jiao, Y., Kou, L., Liao, T. & Du, A. (2018). J. Mater. Chem. C. 6, 6132-6137.]). However, unlike the non-spin polarization system LaCuO3, the spin-polarized Dirac behavior in LaNiO3 is intrinsic, and this system can obtain magnetism without the help of experimental techniques such as applied electric field and pressure. For LaNiO3, the band structures of both spin channels were also calculated according to its experimental lattice constants, as shown in Fig. 2[link](c); one can see that the electronic structure results are consistent with the theoretical ones except for very small band gap differences (∼0.06 eV).

In Fig. 2[link](b), the band structures without the effect of on-site Coulomb interaction U have been given. Compared with GGA+U results, the energy levels of the MDCs in the spin-up direction rose by approximately 0.5 eV. Also, we want to point out that the Dirac cones of LaNiO3 are observed in both the spin-up and spin-down directions [yellow area shown in Figs. 2[link](a) and 2(c)]. As shown in Fig. 2[link](b), for the case of U = 0 eV, the energy levels of the MDCs in the spin-down direction were approximately 0.5 eV higher than those in the spin-up direction. Spin-polarized Dirac cones have more interesting properties than conventional cones (e.g. those in graphene) because of the high spin polarization. For the spin-degenerate Dirac cones in graphene, the Dirac state is destroyed, although artificial modification by introducing border structures, dopants, defects and adsorbed atoms can be used to obtain magnetism. To further clarify the effect of U on the band structures of [R\bar 3 c] LaNiO3, different U values (U = 1, 3, 5 and 7 eV) were taken into consideration in these electronic structure calculations and the results are shown in Fig. S2. From this figure we can see that the MDCs in the spin-up channel are almost unchanged; however, for the half-metallic property, the effects are greater. For U = 0 eV and U = 1 eV, the electronic states and the Fermi level overlap with each other in both spin channels; therefore, the complete spin-polarization of LaNiO3 was lost. However, for U = 3, 5 and 7 eV, the 100% spin-polarization was maintained in this material. With the increase of U values, the band gap in the spin-down channel increases, reflecting that the half-metallic properties of this material become more and more robust.

In Fig. S3, the orbital-resolved band structure for LaNiO3 is plotted with the aim of furthering our understanding of the electronic structure. From it, we can see that the p orbitals of the O atoms and the d orbitals of Ni atoms mainly contribute to the total electronic structures between energies of −0.8 and 1.6 eV. In this region, the La-d orbital made little contribution to the band structures. That is to say, Dirac cones near the Fermi level mainly come from the hybridization between the O-p and Ni-d orbitals. We should point out that these types of MDCs arising from d orbitals are very rare. As shown in Fig. S4, the density-of-states results show that the states in the range −2–0 eV mainly arise from the O-p orbitals, and the states in the range 0–2 eV are derived from the hybridization between the O-p and Ni-d orbitals. The states ranging from 2 to 4 eV in the spin-down channel were contributed by the Ni-d orbital. The spin polarization (P) of LaNiO3 around the Fermi level can be obtained according to the formula

[P = {{| {N \uparrow ({{E_{\rm f}}}) - N \downarrow ({{E_{\rm f}}} )} |} \over {| {N \uparrow ({{E_{\rm f}}} ) + N \downarrow ({{E_{\rm f}}} )} |}},]

where [N \uparrow ({{E_{\rm f}}})] and [N \downarrow ({{E_{\rm f}}})] are the number of spin-up and spin down states, respectively. As shown in Fig. S4, one can see the P of this material is 100%, indicating LaNiO3 could be useful for spin injection (He et al., 2019[He, J., Ding, G., Zhong, C., Li, S., Lib, D. & Zhang, G. (2019). Nanoscale, 11, 356-364.]).

Fig. 3[link] presents a structural diagram of the energy bands, showing the effects of spin-orbit coupling (SOC) on the Dirac cones. Dirac cones still present under the influence of SOC, and the conduction and valence bands are still degenerate with respect to each other. That is to say, for the SOC effect, the MDCs near the Fermi level show strong resistance, reflecting that this material has long spin coherence which is favorable for spin transport. As shown in Fig. S5, the effects of uniform strain have been taken into consideration because the experimental lattice constants of materials always deviate from the ideal calculated equilibrium lattice constants; −8 GPa represents the compression stress of the 8 GPa imposed along the x, y and z directions. In contrast, the positive value represents the imposed tensile stress, from which, one can see this effect will have almost no impact on Dirac cones. The materials with robust MDCs are able to withstand external factors more easily.

[Figure 3]
Figure 3
Calculated band structures of LaNiO3 obtained using the GGA+U and SOC methods at its optimized equilibrium lattice constants.

In this work, we also selected LaNiO3 as an example to study thermodynamic properties including the thermal expansivity α, the heat capacity CV, the Gruneisen constant γ and the Debye temperature ΘD. Such a study was necessary. From this investigation, we were able to figure out the special properties of LaNiO3 under high pressures and high temperatures. The results are shown in Figs. S6–S10 and more details can be found in the supporting information.

Then, we examined the thermal stability of LaNiO3 at room temperature. To achieve this goal, ab initio molecular dynamics simulations (AIMD) were performed and a 2 × 2 × 1 superstructure of LaNiO3 was built. As shown in Fig. 4[link](c), this 2 × 2 × 1 superstructure contains 120 atoms (i.e. 24 La atoms, 24 Ni atoms and 72 O atoms) and the experiment was performed using a Nosé–Hoover thermostat at 300 K. Fig. 4[link](a) presents the fluctuations of the potential energy as a function of the simulation time (i.e. 2 ps) at 300 K. After 2 ps with a time step of 1 fs, we found no structural destruction of LaNiO3, except for some thermal-induced fluctuations [see Fig. 4[link](d)], indicating that this material is thermally stable at room temperature. Also, as shown in Fig. 4[link](b), the magnetic moment of this superstructure keeps a fixed value of 24 µB, showing that the total Mt can survive at room temperature.

[Figure 4]
Figure 4
(a) Calculated total energy fluctuation of the superstructure and (b) the total magnetic moments of the superstructure for LaNiO3 during AIMD simulations at 300 K. The superstructures show a snapshot at the end of the simulation of (c) 0 ps and (d) 2 ps.

The electronic structures of La0.833NiO3 and LaNi0.833O3 were shown in Figs. S11(a) and S11(b), respectively. From these figures, we can further confirm that the Ni-d orbital contributes more to the Dirac cones near the Fermi level than La-p orbital. For La0.833NiO3, the Dirac cones in the spin-up channel are still recognizable even if one La atom is lost. However, for the case of LaNi0.833O3, the Dirac cones in the spin-up channel are badly damaged when one Ni atom is missing. Moreover, we should point out that, for La0.833NiO3, complete spin-polarization of system has been broken.

Finally, the crystal structure of a material is strongly correlated with its physical properties. As demonstrated in our previous study (Wang et al., 2017b[Wang, X., Cheng, Z., Yuan, H. & Khenata, R. (2017b). J. Mater. Chem. C. 5, 11559-11564.]), Ti-based Heusler alloys include many Hg2CuTi-structured materials with high spin polarization. In contrast, the spin polarization of materials with the corresponding Cu2MnAl structure is much lower. That is to say, this specific space group allows for the three-dimensional Dirac point to be used as symmetric protection for degeneracy. Therefore, using a series of examples, we attempted to show that rhombohedral structures with the space group [R\bar 3c] comprise an important system that includes many DHMs with MDCs with strong potential for theoretical prediction and experimental synthesis. The structural diagrams of the energy bands of a series of materials with the general formula LnNiO3 (Ln = Ce, Nd, Pm, Gd, Tb, Dy, Ho, Er, Lu) and space group [R\bar 3c] are plotted in Fig. S12. From this figure, we can see that they are all DHMs with MDCs and 100% spin-polarization. The total and atomic magnetic moments as well as the obtained equilibrium lattice constants are given in Table 1[link]. We should point out here that the magnetic ordering of rare-earth atom Ln only occurs at very low temperature (Muñoz et al., 2009[Muñoz, A., Alonso, J. A., Martínez-Lope, M. J. & Fernández-Díaz, M. T. (2009). J. Solid State Chem. 182, 1982-1989.]; Fernández-Díaz et al., 2001[Fernández-Díaz, M. T., Alonso, J. A., Martínez-Lope, M. J., Casais, M. T. & García-Muñoz, J. L. (2001). Phys. Rev. B, 64, 144417.]) and therefore, in this work, we neglect the magnetic ordering of the Ln, namely, the f electrons of the Ln atom are frozen into the core and the corresponding p and s states are included as valence electrons. For these LnNiO3 (Ln = Ce, Nd, Pm, Gd, Tb, Dy, Ho, Er, Lu) materials, many methods can be used to prepare them. For example, (i) we can synthesize bulk by solid state reaction at elevated temperatures and high oxygen pressure; (ii) we can try to prepare thin films by PLD or sol-gel methods; (iii) we can try to synthesize nano-materials by hydro­thermal methods. Hence, we hope that our current work can give some inspiration to the subsequent experimental and theoretical work, and that more [R\bar 3c]-based Dirac materials will receive attention.

In summary, a series of [R\bar 3 c]-based DHMs with MDCs has been theoretically predicted: (1) LaNiO3, which was synthesized more than 20 years ago; and (2) LnNiO3 (Ln = Ce, Nd, Pm, Gd, Tb, Dy, Ho, Er, Lu), which have been predicted in terms of theory and need attention in experiment. Moreover, for experimentally synthesized material LaNiO3, we perform an all-round first-principle study on its electronic, magnetism and thermodynamic properties. The effects of U, uniform stain, SOC, vacancies and thermal stability have also been investigated via first principles, AIMD and quasi-harmonic Debye approximation. Considering the results of both Du et al. (Jiao et al., 2017[Jiao, Y., Ma, F., Zhang, C., Bell, J., Sanvito, S. & Du, A. (2017). Phys. Rev. Lett. 119, 016403.]) and this study, we believe that many materials in the space group [R\bar 3c] are worthy of theoretical development and experimental preparation. We hope that this letter stimulates theoretical research on DHMs with MDCs in the space group [R\bar 3c]. Furthermore, we believe that experimental preparation and the confirmation of predicted physical properties are imminent. This letter provides a theoretical basis for such experimental studies and is expected to help realize DHMs with MDCs in future spintronic applications.

In the supporting information we provide extensive details for the computational methods as well as the quasi-harmonic Debye model calculations, analysis of the thermodynamic properties and figures.

Supporting information


Footnotes

These authors contributed equally to this work.

§These authors contributed equally to this work

Funding information

The authors declare no competing financial interest and this research was funded by the National Key R&D Program of China (grant No. 2017YFA0303202), National Natural Science Foundation of China (grant Nos. 51801163, 11874306), the Research Funds from the Australia Research Council (DP190100150) and the Natural Science Foundation of Chongqing (grant No. CSTC-2017jcyjBX0035).

References

First citationAnisimov, V. I., Zaanen, J. & Andersen, O. K. (1991). Phys. Rev. B, 44, 943–954.  CrossRef CAS Web of Science Google Scholar
First citationBainsla, L., Mallick, A. I., Raja, M. M., Nigam, A. K., Varaprasad, B. S. D. C. S., Takahashi, Y. K., Alam, A., Suresh, K. G. & Hono, K. (2015). Phys. Rev. B, 91, 104408.  Web of Science CrossRef Google Scholar
First citationBalachandran, P. V., Emery, A. A., Gubernatis, J. E., Lookman, T., Wolverton, C. & Zunger, A. (2018). Phys. Rev. Mater. 2, 043802.  Web of Science CrossRef Google Scholar
First citationCastro Neto, A. H., Guinea, F., Peres, N. M. R., Novoselov, K. S. & Geim, A. K. (2009). Rev. Mod. Phys. 81, 109–162.  Web of Science CrossRef CAS Google Scholar
First citationFernández-Díaz, M. T., Alonso, J. A., Martínez-Lope, M. J., Casais, M. T. & García-Muñoz, J. L. (2001). Phys. Rev. B, 64, 144417.  Google Scholar
First citationHafner, J. (2007). Comput. Phys. Commun. 177, 6–13.  Web of Science CrossRef CAS Google Scholar
First citationHe, J., Ding, G., Zhong, C., Li, S., Lib, D. & Zhang, G. (2019). Nanoscale, 11, 356–364.  Web of Science CrossRef CAS Google Scholar
First citationHe, J., Li, X., Lyu, P. & Nachtigall, P. (2017). Nanoscale, 9, 2246–2252.  Web of Science CrossRef CAS PubMed Google Scholar
First citationHe, J., Ma, S., Lyu, P. & Nachtigall, P. (2016). J. Mater. Chem. C. 4, 2518–2526.  Web of Science CrossRef CAS Google Scholar
First citationIshizuka, H. & Motome, Y. (2012). Phys. Rev. Lett. 109, 237207.  Web of Science CrossRef PubMed Google Scholar
First citationJiao, Y., Ma, F., Zhang, C., Bell, J., Sanvito, S. & Du, A. (2017). Phys. Rev. Lett. 119, 016403.  Web of Science CrossRef PubMed Google Scholar
First citationKan, E., Wu, F., Xiang, H., Yang, J. & Whangbo, M. (2011). J. Phys. Chem. C, 115, 17252–17254.  Web of Science CrossRef CAS Google Scholar
First citationLi, X. & Yang, J. (2017). WIREs Comput. Mol. Sci. 7, e1314.  Google Scholar
First citationLiu, Z., Liu, J. & Zhao, J. (2017). Nano Res. 10, 1972–1979.  Web of Science CrossRef CAS Google Scholar
First citationMa, F., Jiao, Y., Jiang, Z. & Du, A. (2018). Appl. Mater. Interfaces, 10, 36088–36093.  Web of Science CrossRef CAS Google Scholar
First citationMuñoz, A., Alonso, J. A., Martínez-Lope, M. J. & Fernández-Díaz, M. T. (2009). J. Solid State Chem. 182, 1982–1989.  Google Scholar
First citationOuardi, S., Fecher, G. H., Felser, C. & Kübler, J. (2013). Phys. Rev. Lett. 110, 100401.  Web of Science CrossRef PubMed Google Scholar
First citationPeverati, R. & Truhlar, D. G. (2011). J. Phys. Chem. Lett. 2, 2810–2817.  Web of Science CrossRef CAS Google Scholar
First citationQin, G., Wu, W., Hu, S., Tao, Y., Yan, X., Jing, C., Li, X., Gu, H., Cao, S. & Ren, W. (2017). IUCrJ, 4, 506–511.  Web of Science CrossRef CAS PubMed IUCr Journals Google Scholar
First citationSreedhar, K., Honig, J. M., Darwin, M., McElfresh, M., Shand, P. M., Xu, J., Crooker, B. C. & Spalek, J. (1992). Phys. Rev. B, 46, 6382–6386.  CrossRef CAS Web of Science Google Scholar
First citationWang, X., Cheng, Z., Liu, G., Dai, X., Khenata, R., Wang, L. & Bouhemadou, A. (2017a). IUCrJ, 4, 758–768.  Web of Science CrossRef CAS PubMed IUCr Journals Google Scholar
First citationWang, X., Cheng, Z., Wang, J., Wang, X. & Liu, G. (2016). J. Mater. Chem. C. 4, 7176–7192.  Web of Science CrossRef CAS Google Scholar
First citationWang, X., Cheng, Z., Yuan, H. & Khenata, R. (2017b). J. Mater. Chem. C. 5, 11559–11564.  Web of Science CrossRef CAS Google Scholar
First citationWang, X., Li, T., Cheng, Z., Wang, X. & Chen, H. (2018). Appl. Phys. Rev. 5, 041103.  Web of Science CrossRef Google Scholar
First citationWang, X. L. (2008). Phys. Rev. Lett. 100, 156404.  Web of Science CrossRef PubMed Google Scholar
First citationWang, X. L. (2017). Natl Sci. Rev. 4, 252–257.  CAS Google Scholar
First citationWang, X. L., Dou, S. X. & Zhang, C. (2010). NPG Asia Mater. 2, 31.  Google Scholar
First citationZhang, C., Jiao, Y., Kou, L., Liao, T. & Du, A. (2018). J. Mater. Chem. C. 6, 6132–6137.  Web of Science CrossRef CAS Google Scholar
First citationZhang, X., Jin, L., Dai, X. & Liu, G. (2017). J. Phys. Chem. Lett. 8, 4814–4819.  Web of Science CrossRef CAS PubMed Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

IUCrJ
Volume 6| Part 6| November 2019| Pages 990-995
ISSN: 2052-2525