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

Emerging Two-Dimensional Magnetism in Nonmagnetic Electrides Hf2X (X = S, Se, Te)

Shuyuan Liu1, Chongze Wang1, Hyunsoo Jeon1, Jia Yu2, and Jun-Hyung Cho1∗ 1Department of Physics and Research Institute for Natural Science, Hanyang University, 222 Wangsimni-ro, Seongdong-Ku, Seoul 04763, Republic of Korea
2Key Laboratory for Special Functional Materials of the Ministry of Education, Henan University, Kaifeng 475004, People’s Republic of China
Abstract

Recent experimental discoveries of two-dimensional (2D) magnets have triggered intense research activities to search for atomically thin magnetic systems. Using first-principles calculations, we predict the emergence of 2D magnetism in the monolayers (MLs), few layers, and surfaces of nonmagnetic layered electrides Hf2X (X = S, Se, Te) consisting of three-atom-thick Hf-X-Hf stacks. It is revealed that each bulk Hf2X hosts a novel quantum state of Dirac nodal lines with a high density of states arising from Hf-5dd cationic and interlayer anionic electrons around -0.9 eV below the Fermi level EFE_{F}. However, for the MLs, few layers, and surfaces of Hf2X, such hybridized states are shifted toward EFE_{F} to generate van Hove singularities, leading to a Stoner instability. The resulting surface ferromagnetism gives rise to strongly spin-polarized topological surface states at Hf2X(001), demonstrating that anionic electrons, 2D magnetism, and band topology are entangled with each other. Our findings will open new perspectives for the discovery of 2D magnets via exploiting surface effects in nonmagnetic layered electrides.

A wide variety of 2D materials have been explored to exhibit unique physical and chemical properties that are strikingly different from those of their 3D parent compounds [1, 2]. For examples, in contrast to graphite that is a semimetal with an overlap between the conduction and valence bands [3], graphene is a zero-gap semiconductor [4]; some transition metal dichalcogenides (TMDCs) exhibit a change in the electronic structure from indirect bandgap in their bulk form to direct bandgap in monolayers (MLs) [5]. By exploiting such dimensionality-driven novel electronic and optical properties, 2D materials including graphene, TMDCs, and the families of monoelemental (e.g., black phosphorus [6], arsenene [7], antimonene [8], and bismuthine [9]) and ternary (e.g., Bi2O2Se [10], BiOX [11], and CrOX (X = Cl, Br, I) [12]) compounds have been illustrated to provide many exciting new opportunities for diverse technical applications at the atomic and nanometer scales [13, 14, 15].

Nevertheless, 2D magnetism has been a long-standing elusive issue [16]. According to the Mermin-Wagner theorem [17], 2D magnetic systems described by the isotropic Heisenberg model cannot have a long-range magnetic ordering at any finite temperature due to thermal fluctuations. However, magnetocrystalline anisotropy enables the suppression of such thermal fluctuations, thereby allowing the stabilization of 2D magnetism [18]. Despite the early availability of magnetic van der Waals (vdW)-layered crystals, the discovery of 2D magnetism in their ML or few-layer form has only recently been made experimentally [19, 20, 21, 22, 23]. For the vdW-layered ferromagnets CrI3 [19] and Cr2Ge2Te6 [20], ferromagnetic (FM) order was observed to be maintained down to the ML and bilayer limits at low temperatures, respectively. Meanwhile, bulk FePS3 having an antiferromagnetic (AFM) order was experimentally observed to preserve its AFM property up to ML and few layer [21]. Interestingly, for VSe2, bulk does not have spontaneous magnetization, but ML VSe2 exhibits a FM order [22, 23]. These experimental evidences of 2D magnetism in vdW-layered materials have stimulated many searches for a variety of 2D magnetic candidate materials [24].

As the unconventional class of ionic compounds, 2D layered electrides A2X consisting of a three-atom-thick building block of A-X-A stacks [see Fig. 1(a)] have recently been discovered to offer promising opportunities for both fundamental research and technological applications [25, 26, 27, 28, 29, 30, 31, 32]. Depending on the cationic constituent atoms of A2X, anionic excess electrons confined in the interstitial spaces between positively charged A-X-A stacks are distributed in different degrees of localization, which in turn lead to nonmagnetic (NM) or magnetic electrides [33]. So far, NM Ca2[26], Hf2[34], Sr3CrN3 [35], Sr8P5 [36], and Sr5P3 [36], paramagnetic Y2[37, 38, 39], and FM Gd2[40] and YCl [41] have been synthesized experimentally. Here, we theoretically predict the emergence of 2D magnetism in the MLs, few layers, and surfaces of the bulk nonmagnetic layered electrides Hf2X (X = S, Se, Te). Unlike the above-mentioned 2D vdW-layered magnets, the MLs and few layers of Hf2X show strong surface effects that significantly modify their band structures and charge distributions at the outermost Hf layers, thereby inducing a Stoner instability, as will be demonstrated later.

In this Letter, we focus on a recently synthesized [34] electride Hf2S to explore 2D magnetism in its ML, few layer, and surface using first-principles density-functional theory (DFT) calculations. We find dramatic changes in the electronic structure of ML and few layer: i.e., bulk Hf2S has a large density of states (DOS) around -0.9 eV below the Fermi level EFE_{F}, arising from hybridized Hf-5dd cationic and interlayer anionic states, while ML and few layer Hf2S exhibit a shift of such a van Hove singularity (vHs) toward EFE_{F}. As a result, the outermost Hf layers in the top and bottom surfaces of ML and few layer have an in-plane FM order with opposite spin polarizations. Moreover, we find that the Hf2S(001) surface hosts nontrivial topological surface states associated with the bulk Dirac nodal lines, which are strongly spin-polarized to exhibit a surface ferromagnetism. The present findings can also be applicable to other isoelectronic NM electrides Hf2Se and Hf2Te, where emerging 2D magnetism is enhanced compared to Hf2S. Therefore, we propose a new family of layered electrides Hf2X (X = S, Se, Te) showing an intriguing surface-driven transformation from bulk nonmagnetic to 2D magnetic order.

We begin by examining the ground state of bulk Hf2S using DFT calculations [42]. Our spin-polarized calculations for bulk Hf2S show that any initial FM or AFM configuration converges to the NM state. Therefore, bulk Hf2S has a NM ground state, consistent with a combined experimental and DFT study of Kang etet alal[34]. Figures 1(a) and 1(b) show the optimized ground structure with the lattice parameters a1a_{1} = a2a_{2} = 3.375 Å and a3a_{3} = 11.764 Å, where S atoms locating in a triangular lattice are surrounded by six Hf atoms in an trigonal prismatic geometry with the space group P63/mmcmmc (No. 194). The calculated band structure of bulk Hf2S shows that Hf-5dd cationic and interstitial anionic states are strongly hybridized around -0.9 eV below EFE_{F} [48], giving rise to a large peak in their partial density of states (PDOS) [see Fig. 1(c)] [49]. Here, the anionic electrons localized at the positions marked as X1X_{1} and X2X_{2} in the interlayer space are well represented by the electron localization function (ELF) [see Fig. 1(d)]. These local maxima positions of ELF agree well with those of a previous DFT calculation [34]. Interestingly, for the ML and few layer of Hf2S, such hybridized Hf-5dd cationic and interstitial anionic states are shifted toward EFE_{F}, thereby inducing a magnetic instability, as discussed below.

Refer to caption
Figure 1: (a) Optimized structure of bulk Hf2S and (b) its top view. The primitive unit cell contains two Hf-S-Hf stacks with an alternative AB stacking sequence. In (b), the blue and green circles represent Hf atoms in two different stacks, while the dashed circles represent X1X_{1} and X2X_{2} anions locating at hollow sites. The calculated band structure of bulk Hf2S is displayed in (c) together with the PDOS, where the projected bands onto Hf-5dd and X1X_{1}-, X2X_{2}-ss-like orbitals are represented by circles whose radii are proportional to the weights of the corresponding orbitals. The inset of (c) shows the BZ of the primitive unit cell. In (d), the calculated ELF of Hf2S is drawn on the (110) plane with a contour spacing of 0.05, where the MT radius of X1X_{1} and X2X_{2} anions is chosen as 1 Å. In (e), the SOC gap of 2D nodal surfaces is displayed throughout the kzk_{z}=π{\pi}/cc plane, where the black lines represent DNLs.

It is noticeable that there is a fourfold degenerate band crossing EFE_{F} along the high-symmetry ALHAA-L-H-A paths, indicated by the arrow in Fig. 1(c). Using the tight-binding Hamiltonian with a basis of maximally localized Wannier functions [50, 51], we find the existence of 2D nodal surface crossing EFE_{F} throughout the kzk_{z}=π{\pi}/cc plane on the boundary of Brillouin zone (BZ) [52]. This nodal surface formed by a touching of two doubly-degenerate bands is respected 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 (see symmetry analysis in the Supplemental Material [53]). The inclusion of SOC lifts the fourfold degeneracy of nodal surface except along the high-symmetry paths kxk_{x}=0 and kxk_{x}=±3ky{\pm}\sqrt{3}k_{y}, preserving 1D nodal lines [see Fig. 1(e)]. These Dirac nodal lines (DNLs) showing C3zC_{3z} rotation symmetry are protected by additional mirror symmetries [54]. Thus, bulk Hf2S is characterized as a topological semimetal having DNLs crossing EFE_{F}.

To reveal the effect of reduced dimensionality on electronic structure, we first consider the NM phase of ML Hf2S. We find that the lattice constants become a1a_{1} = a2a_{2} = 3.230 Å, slightly smaller than those (3.375 Å) of bulk Hf2S. As shown in Fig. 2(a), ML Hf2S has a large peak in the PDOS of Hf-5dd cationic and interstitial anionic states around EFE_{F}. We increase the interlayer spacing dsd_{s} [see Fig. 1(a)] in bulk Hf2S to examine the change of band structure. As dsd_{s} increases, the hybridized Hf-5dd cationic and interstitial anionic states locating around -0.9 eV are shifted toward EFE_{F} (see Fig. S4 in the Supplemental Material [53]), converging to the band structure of ML Hf2S. We also find that the distribution of interstitial anionic electrons changes with respect to dsd_{s}: i.e., for bulk Hf2S, the number of electrons nX1n_{X_{1}} (nX2n_{X_{2}}) within the muffin-tin (MT) sphere of the X1X_{1} (X2X_{2}) anion is 1.120 (0.405) electrons per ML [see Fig. 1(d)], which decreases (increases) to 0.904 (0.510) electrons in ML Hf2S. Such rearrangements of nX1n_{X_{1}} and nX2n_{X_{2}} in ML Hf2S [see Fig. 2(b)] together with the shift of hybridized Hf-5dd cationic and interstitial anionic states reflect strong surface effects, as will be demonstrated at the Hf2S(001) surface. As shown in Fig. S5, the partial charge distributions with respect to the energy ranges exhibit a strong variation of anionic electrons between bulk and ML Hf2S, due to breaking bonds at surfaces [55]. In these senses, surface formation significantly changes the distribution of interstitial anionic electrons as well as their electronic band structure, both of which are the peculiar features of Hf2X. These dramatic changes of interstitial anionic states between bulk and surface induce the emergence of 2D ferromagnetism in Hf2X, as discussed below.

Refer to caption
Figure 2: (a) Calculated band structure with PDOS and (b) ELF (with a contour spacing of 0.05) of the NM phase of ML Hf2S. The band structure and spin-polarized LDOS (projected onto Hft and Hfb) of the AFM phase of ML Hf2S are displayed in (c) and (d), respectively. The AFM spin density is also given in (d) with an isosurface of 0.01 ee3.

Since the band structure of the NM phase of ML Hf2S has a vHs at EFE_{F} [see Fig. 2(a)], the Stoner criterion may be fulfilled to lead to a FM instability. Indeed, we find that the FM phase is favored over the NM phase by 1.55 meV per Hf atom. However, the AFM coupling of the ferromagnetically ordered magnetic moments on the outermost Hf layers is further stabilized compared to the FM phase by 11.41 meV per Hf atom. In such an AFM ground state, the calculated spin magnetic moments integrated within the MT spheres around Hf [56], X1X_{1}, and X2X_{2} are 0.498, 0.104, and 0.105 μB{\mu}_{B}, respectively. Here, the AFM coupling of Hf 5dd spins is likely driven by superexchange interactions [57, 58, 59] through the occupied S 3pp states (see Fig. S6 in the Supplemental Material [53]). Consequently, this AFM ordering opens a pseudogap for the electronic states around EFE_{F} [see Fig. 2(c)]. As shown in Fig. 2(d), the spin-up and spin-down local DOS (LDOS) projected onto Hft and Hfb residing at the top and bottom layers exhibit the separation between occupied and unoccupied states [60]. By including SOC, the easy axis points along the zz direction with a magnetic anisotropy energy of 0.76 meV per Hf atom, indicating that ML Hf2S has the Ising anisotropy with a strong AFM interlayer interaction. Since this AFM ML structure is revealed to be thermodynamically stable [61] [see Figs. S7(a) and S7(b)], we anticipate that it would be experimentally synthesized in the future by either mechanical exfoliation such as graphene [62] and MoS2 [63] or epitaxial growth on proper substrates such as silicene [64], stanene [65], and tellurene [66].

Next, we investigate the stability of the FM and AFM phases of few-layer Hf2S with increasing the number NN of Hf-S-Hf stacks. Figure 3(a) shows that the energy difference ΔE{\Delta}E between the FM and AFM phases decreases sharply even at NN = 2, indicating that the top and bottom surfaces of few-layer Hf2S can form an isolated FM order with their suppressed AFM coupling. Figure 3(b) shows the spin density at the (001) surface, obtained using a periodic slab of NN = 12 with {\sim}25 Å of vacuum in-between adjacent slabs. We find that the spin magnetic moment exists mostly at the topmost Hf layer, while it is significantly reduced at the second and third Hf layers (see Table I). The projected LDOS demonstrates that the spin-up and -down states arising from Hf-1 atom are separated by {\sim}0.47 eV [67], while those from Hf-2 atom exhibit a little separation. By dividing this exchange splitting of Hf-1 by the corresponding magnetic moment, we can estimate the Stoner parameter II, which satisfies the Stoner criterion D(EF)I>1D(E_{F})I>1 [68] [see Fig. S8(a) in the Supplemental Material [53]]. Here, D(EF)D(E_{F}) is the DOS at EFE_{F} from the NM phase. Thus, we can say that surface ferromagnetism emerging at the Hf2S(001) surface is driven by the Stoner instability due to a vHs at EFE_{F} arising from hybridized Hf-5dd cationic and surface anionic states [see Fig. S8(b)]. In order to estimate the Curie temperature TcT_{\rm c}, we perform spin-polarized calculations for various AFM surface configurations (see Fig. S9). We find that the lowest AFM configuration is less stable than the FM one by 15.35 meV per Hf atom. Using the mean field approximation [69], we estimate a TcT_{\rm c} of {\sim}118 K.

Refer to caption
Figure 3: (a) Calculated total energies of the FM and AFM phases (relative to that of the NM phase) of Hf2S, Hf2Se, and Hf2Te as a function of NN. Here, ”SS” represents the data of the (001) surface, obtained using a periodic slab of NN = 12. The spin density (with a contour spacing of 0.005 ee3) and the LDOS of Hf-1 and Hf-2 atoms at the Hf2S(001) surface are given in (b) and (c), respectively.

According to the bulk-boundary correspondence of topological nodal line semimetals [70], the presence of DNLs in bulk leads to the formation of topologically protected surface states [71, 72]. Using the Green’s function method based on the tight-binding Hamiltonian with maximally localized Wannier functions [50, 51], we obtain the projected surface spectrum of the NM Hf2S(001) surface [see Fig. 4(a)]. There are three drumhead surface states SS1SS_{1}, SS2SS_{2}, and SS3SS_{3} along the Γ¯M¯K¯Γ¯{\overline{\Gamma}}-{\overline{M}}-{\overline{K}}-{\overline{\Gamma}} path near EFE_{F}, which represents the hallmark of DNLs [73]. Note that SS1SS_{1}, SS2SS_{2}, and SS3SS_{3} are split by the SOC-driven gap openings (see Fig. S10 in the Supplemental Material [53]). In Fig. 4(b), the Fermi surface of Hf2S(001) at a chemical potential of -0.2 eV exhibits the closed Fermi arcs around the Γ¯{\overline{\Gamma}} and K¯\overline{K} point with helical spin textures. Such nontrivial topological surface states with a unique spin-momentum locking property [74] can be more distinguishable by subtracting the (001) projected bulk states from the surface spectrum (see Fig. S11). For the experimental measurements of these topological surface states, we propose the H-passivation of the Hf2S(001) surface where surface ferromagnetism can be removed [75]. Our DFT band structure of the H-passivated Hf2S(001) surface reproduces the dispersion of the SS1SS_{1}, SS2SS_{2}, and SS3SS_{3} states around EFE_{F} [see Figs. 4(c) and S10]. Meanwhile, the DFT band structure of a clean Hf2S(001) surface shows that the surface-induced ferromagnetism gives rise to strong spin polarizations for the SS1SS_{1}, SS2SS_{2}, and SS3SS_{3} states at Hf2S(001) [see Fig. 4(d)], indicating that the bulk DNLs are split into two spin-polarized nondegenerate bands due to time-reversal symmetry breaking at surface. These results reflect a strong correlation between spin degree of freedom and topological properties. It is thus likely that anionic electrons, 2D magnetism, and band topology are entangled with each other in a new class of electrides Hf2X.

Refer to caption
Figure 4: (a) Projected surface spectrum of the NM Hf2S(001) surface and (b) its isoenergy surface at -0.2 eV with the in-plane spin textures of the SS1SS_{1}, SS2SS_{2}, and SS3SS_{3} surface states around the Γ¯{\overline{\Gamma}} and K¯{\overline{K}} points. The DFT band structures of the H-passivated Hf2S(001) and clean Hf2S(001) surfaces are given in (c) and (d), respectively. The former (latter) bands are obtained with (without) including SOC. In (d), the radii of circles are proportional to the weights of the projection onto Hf-1 atom.

Finally, we examine 2D magnetism in other isoelectronic NM electrides Hf2Se and Hf2Te. Similar to the case of Hf2S, the MLs and few layers of Hf2Se and Hf2Te shift hybridized Hf-5dd cationic and interlayer anionic states toward EFE_{F} (see Fig. S13 in the Supplemental Material [53]), thereby inducing a 2D magnetism at their outermost Hf layers. As shown in Fig. 3(a), the magnetic stabilities in the MLs, few layers, and surfaces of Hf2Se and Hf2Te are enhanced compared to the corresponding ones of Hf2S. Accordingly, the spin magnetic moments of Hf-1 atom in Hf2Se and Hf2Te are larger than that in Hf2S (see Table I). For the Hf2Se(001) and Hf2Te(001) surfaces, the topological surface states associated with the bulk DNLs also exhibit the large spin splits around EFE_{F} (see Fig. S14), leading to a strong surface ferromagnetism.

Table 1: Calculated spin magnetic moments (in unit of μB{\mu}_{B} per Hf atom) of Hf-1, Hf-2, and Hf-3 at the Hf2S(001), Hf2Se(001), and Hf2Te(001) surfaces. The values of X1X_{1} and X2X_{2} at surface are also given.
Hf-1 Hf-2 Hf-3 X1X_{1} X2X_{2}
Hf2S(001) 0.490 -0.011 0.001 0.086 0.100
Hf2Se(001) 0.511 -0.011 0.001 0.084 0.092
Hf2Te(001) 0.536 -0.011 0.001 0.083 0.083

In summary, our first-principles DFT calculations have demonstrated the importance of surface effects that invokes the emergence of 2D magnetism in the MLs, few layers, and surfaces of nonmagnetic layered electrides Hf2X (X = S, Se, Te). Specifically, we revealed that nontrivial topological surface states associated with the bulk DNLs are largely spin-polarized to form a strong ferromagnetism at the Hf2X(001) surfaces. Our findings provide a novel platform to investigate the intriguing interplay between electride properties, nontrivial band topology, and surface ferromagnetism, which will be promising for future spintronics technologies [76, 77]. We found that Zr2X (X = S, Se, Te) can also exhibit similar electronic structures, 2D magnetism, and topological properties [see Figs. S7(c), S7(d), S15, and Table SI], as predicted for Hf2X. Thus, the present prediction of surface ferromagnetism via the strong variation of hybridized transition-metal dd orbitals and interstitial anionic states between bulks and surfaces of Hf2X and Zr2X are rather generic, thereby providing a more general physical picture to support our results.

Corresponding author: [email protected]

References

  • [1] A. Gupta, T. Sakthivel, and S. Seal, Prog. Mater. Sci. 73, 44 (2015).
  • [2] K. S. Novoselov, A. Mishchenko, A. Carvalho, and A. H. Castro Neto, Science 353, 6298 (2016).
  • [3] J. C. Slonczewski and P. R. Weiss, Phys. Rev. 109, 272 (1958).
  • [4] M. Sprinkle, D. Siegel, Y. Hu, J. Hicks, A. Tejeda, A. Taleb-Ibrahimi, P. Le Fevre, F. Bertran, S. Vizzini, H. Enriquez, S. Chiang, P. Soukiassian, C. Berger, W. A. de Heer, A. Lanzara, and E. H. Conrad, Phys. Rev. Lett. 103, 226803 (2009).
  • [5] A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, C.-Y. Chim, G. Galli, and F. Wang, Nano Lett. 10, 1271 (2010).
  • [6] L. Li, Y. Yu, G. Jun Ye, Q. Ge, X. Ou, H. Wu, D. Feng, X. Hui Chen, and Y. Zhang, Nat. Nanotechnol. 9, 372 (2014).
  • [7] S. Zhang, Z. Yan, Y. Li, Z. Chen, and H. Zeng, Angew. Chem. 127, 3155 (2015).
  • [8] J. Ji, X. Song, J. Liu, Z. Yan, C. Huo, S. Zhang, M. Su, L. Liao, W. Wang, Z. Ni, Y. Hao, and H. Zeng, Nat. Commun. 7, 13352 (2016).
  • [9] S. Zhang, M. Xie, F. Li, Z. Yan, Y. Li, E. Kan, W. Liu, Z. Chen, and H. Zeng, Angew. Chem. 128, 1698 (2016).
  • [10] J. Wu, H. Yuan, M. M. Meng, C. Chen, Y. Sun, Z. Chen, W. Dang, C. Tan, Y. Liu, J. Yin et al., Nat. Nanotechnol. 12, 530 (2017).
  • [11] X. Zhang, Z. Ai, F. Jia, and L. Zhang, J. Phys. Chem. C 112, 747 (2008).
  • [12] N. Miao, B. Xu, L. Zhu, J. Zhou, and Z. Sun, J. Am. Chem. Soc. 140, 2417 (2018).
  • [13] J. R. Schaibley, H. Yu, G. Clark, P. Rivera, J. S. Ross, K. L. Seyler, W. Yao, and X. Xu, Nat. Rev. Mater. 1, 16055 (2016).
  • [14] M. Zeng, Y. Xiao, J. Liu, K. Yang, and L. Fu, Chem. Rev. 118, 6236 (2018).
  • [15] N. R. Glavin, R. Rao, V. Varshney, E. Bianco, A. Apte, A. Roy, E. Ringe, and P. M. Ajayan, Adv. Mater. 32, 1904302 (2020).
  • [16] M. Gibertini, M. Koperski, A. Morpurgo, and K. Novoselov, Nat. Nanotechnol. 14, 408 (2019).
  • [17] N. D. Mermin, and H. Wagner, Phys. Rev. Lett. 17, 1133 (1966).
  • [18] S. Hope, B.-Ch. Choi, P. J. Bode, and J. A. C. Bland, Phys. Rev. B 61, 5876 (2000).
  • [19] B. Huang, G. Clark, E. Navarro-Moratalla, D. R. Klein, R. Cheng, K. L. Seyler, D. Zhong, E. Schmidgall, M. A. McGuire, D. H. Cobden et al., Nature (London) 546, 270 (2017).
  • [20] C. Gong, L. Li, Z. Li, H. Ji, A. Stern, Y. Xia, T. Cao, W. Bao, C. Wang, Y. Wang et al., Nature (London) 546, 265 (2017).
  • [21] X. Wang, K. Du, Y. Y. F. Liu, P. Hu, J. Zhang, Q. Zhang, M. H. S. Owen, X. Lu, C. K. Gan, P. Sengupta et al., 2D Mater. 3, 031009 (2016).
  • [22] M. Bonilla, S. Kolekar, Y. Ma, H. C. Diaz, V. Kalappattil, R. Das, T. Eggers, H. R. Gutierrez, M.-H. Phan, and M. Batzill, Nat. Nanotechnol. 13, 289 (2018).
  • [23] X. Wang, D. Li, Z. Li, C. Wu, C.-M. Che, G. Chen, and X. Cui, ACS Nano 15, 16236 (2021).
  • [24] X. Jiang, Q. Liu, J. Xing, N. Liu, Y. Guo, Z. Liu, and J. Zhao, Appl. Phys. Rev. 8, 031305 (2021).
  • [25] M. Kitano, Y. Inoue, Y. Yamazaki, F. Hayashi, S. Kanbara, S. Matsuishi, T. Yokoyama, S.-W. Kim, M. Hara, and H. Hosono, Nat. Chem. 4, 934 (2012).
  • [26] K. Lee, S. W. Kim, Y. Toda, S. Matsuishi, and H. Hosono, Nature (London) 494, 336 (2013).
  • [27] T. Inoshita, S. Jeong, N. Hamada, and H. Hosono, Phys. Rev. X 4, 031023 (2014).
  • [28] S. Yi, J.-H. Choi, K. Lee, S. W. Kim, C. H. Park, and J.-H. Cho, Phys. Rev. B 94, 235428 (2016).
  • [29] M. Hirayama, S. Matsuishi, H. Hosono, and S. Murakami, Phys. Rev. X 8, 031067 (2018).
  • [30] L. Liu, C. Wang, S. Yi, D. K. Kim, C. H. Park, and J.-H. Cho, Phys. Rev. B 99, 220401(R) (2019).
  • [31] S. Liu, C. Wang, L. Liu, J.-H. Choi, H.-J. Kim, Y. Jia, C. H. Park, and J.-H. Cho, Phys. Rev. Lett. 125, 187203 (2020).
  • [32] S. Liu, C. Wang, H. Jeon, J. Kim, and J.-H. Cho, Phys. Rev. B 105, L041406 (2022).
  • [33] S. Liu, W. Li, S. W. Kim, and J.-H. Choi, J. Phys. Chem. C 124, 1398 (2020).
  • [34] S. H. Kang, J. Bang, K. Chung, C. N. Nandadasa, G. Han, S. Lee, K. H. Lee, K. Lee, Y. Ma, S. H. Oh, S.-G. Kim, Y.-M. Kim, and S. W. Kim, Sci. Adv. 6, eaba7416 (2020).
  • [35] P. Chanhom, K. E. Fritz, L. A. Burton, J. Kloppenburg, Y. Filinchuk, A. Senyshyn, M. Wang, Z. Feng, N. Insin, J. Suntivich et al., J. Am. Chem. Soc. 141, 10595 (2019).
  • [36] J. Wang, K. Hanzawa, H. Hiramatsu, J. Kim, N. Umezawa, K. Iwanaka, T. Tada, and H. Hosono, J. Am. Chem. Soc. 139, 15668 (2017).
  • [37] X. Zhang, Z. Xiao, H. Lei, Y. Toda, S. Matsuishi, T. Kamiya, S. Ueda, and H. Hosono, Chem. Mater. 26, 6638 (2014).
  • [38] J. Park, K. Lee, S. Y. Lee, C. N. Nandadasa, S. Kim, K. H. Lee, Y. H. Lee, H. Hosono, S.-G. Kim, and S. W. Kim, J. Am. Chem. Soc. 139, 615 (2017).
  • [39] M. Hiraishi, K. M. Kojima, I. Yamauchi, H. Okabe, S. Takeshita, A. Koda, R. Kadono, X. Zhang, S. Matsuishi, H. Hosono, K. Hirata, S. Otani, and N. Ohashi, Phys. Rev. B 98, 041104(R) (2018).
  • [40] S. Y. Lee, J.-Y. Hwang, J. Park, C. N. Nandadasa, Y. Kim, J. Bang, K. Lee, K. H. Lee, Y. Zhang, Y. Ma, H. Hosono, Y. H. Lee, S.-G. Kim, and S. W. Kim, Nat. Commun. 11, 1526 (2020).
  • [41] H. Y. Song, B. I. Yoo, J.-H. Choi, S.-H. Kang, J. Bang, W. Li, C. N. Nandadasa, D. Thapa, D. Yoon, M. J. Han, K. H. Lee, S. G. Kim, K. Lee, and S. W. Kim, Materials Today Physics 20, 100473 (2021).
  • [42] Our first-principles DFT calculations were performed using the Vienna abab initioinitio simulation package with the projector-augmented wave method [43, 44, 45]. The exchange-correlation energy was treated with the generalized-gradient approximation functional of Perdew-Burke-Ernzerhof [46]. The plane wave basis was employed with a kinetic energy cutoff of 550 eV, and the kk-space integration was done with 18×{\times}18×{\times}1 and 18×{\times}18×{\times}6 meshes for ML (few layers) and bulk, respectively. All atoms were allowed to relax along the calculated forces until all the residual force components were less than 0.005 eV/Å. The phonon spectrum calculations of ML Hf2S were carried out by using the QUANTUM ESPRESSO package [47], with the 6×{\times}6×{\times}1 qq points.
  • [43] G. Kresse and J. Hafner, Phys. Rev. B 48, 13115 (1993).
  • [44] G. Kresse and J. Furthmüller, Comput. Mater. Sci. 6, 15 (1996).
  • [45] P. E. Blöchl, Phys. Rev. B 50, 17953 (1994).
  • [46] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996); 78, 1396(E) (1997).
  • [47] 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).
  • [48] This unique feature of anionic electrons in bulk Hf2X can be clearly seen by plotting their charge density around -0.9 eV (see Fig. S1). It is found that the integrated charge density of bulk Hf2S between -0.7 and -1.4 eV describes a charge distribution of hybridized Hf-5dd cationic and interstitial anionic states, while that between EFE_{F} and -0.7 eV represents Hf-5dd states.
  • [49] It is noted that for bulk Hf2S, the PDOS for S-3ss and -3pp orbitals are located below -3 eV (see Fig. S2). Thus, S atoms hardly participate in hybridization with interstitial anionic states.
  • [50] A. A. Mostofi, J. R. Yates, Y.-S. Lee, I. Souza, D. Vanderbilt, and N Marzari, Comput. Phys. Commun. 178, 685 (2008).
  • [51] Q. S. Wu, S. N. Zhang, H.-F. Song, M. Troyer, and A. A. Soluyanov, Comput. Phys. Commun. 224, 405 (2018).
  • [52] As shown in Fig. S3 of the Supplemental Material [53], the Wannier bands of bulk Hf2S agree well with the DFT bands obtained using the VASP code.
  • [53] See Supplemental Material at http://link.aps.org/supplemental/xxxx for the symmetry and topology analyses, the partial charge density of bulk Hf2S, the PDOS of bulk Hf2S, the comparison of the electronic bands obtained using the DFT and tight-binding Hamiltonian calculations, the band structures of bulk Hf2S as a function of interlayer spacing dsd_{s}, the integrated partial charge densities of bulk Hf2S and ML Hf2S, the PDOS of ML Hf2S, the thermodynamical stability the AFM phase of ML Hf2S and Zr2S, the Stoner instability at the Hf2S(001) surface, the Curie temperature at Hf2S(001), the adsorption sites of H atoms on the Hf2S(001) surface, the band structures for the MLs, few layers, and surface of Hf2Se and Hf2Te, the band structure and PDOS of bulk Zr2S and the NM phase of ML Zr2S, the spin magnetic moments of Zr2X surfaces.
  • [54] We note that the mirror symmetry MxM_{x} having (x,y,z)(x,y,z) \rightarrow (x,y,z)(-x,y,z) anticommutes with PS2zPS_{2z} (see symmetry analysis in the Supplemental Material [53]), which 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, 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 the lattice deformations conserving PS2zPS_{2z} and mirror symmetries.
  • [55] It is noted that the redistribution of anionic electrons at surfaces causes the upward shift of interstitial anionic states upon dimensionality reduction. Such intriguing surface effects occurring in Hf2X are absent in conventional layered electrides [28, 30], the bulks of which have largely delocalized interstitial anionic states near EFE_{F}.
  • [56] The radius of the MT sphere around Hf atom is chosen as 1.6 Å.
  • [57] P. W. Anderson, Phys. Rev. 79, 350 (1950).
  • [58] J. B. Goodenough, Phys. Rev. 100, 564 (1955).
  • [59] J. Kanamori, J. Phys. Chem. Solids 10, 87 (1959).
  • [60] Since the electronic states with the same spin direction hybridize with each other, the highest occupied (lowest unoccupied) states are shifted to lower (higher) energies, corresponding to the superexchange mechanism [57, 58, 59].
  • [61] To examine the dynamical stability of ML Hf2S, we investigate its phonon spectrum. It is found that the ML structure is dynamically stable without imaginary-frequency phonon modes [see Fig. S7(a) in the Supplemental Material [53]]. Furthermore, we ensure the thermodynamic stability of ML Hf2S by using abab initioinitio molecular dynamics simulations. Figure S7(b) shows that ML Hf2S preserves its structure up to {\sim}1000K without any structural transformation.
  • [62] K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva, and A. A. Firsov, Science 306, 666 (2004).
  • [63] K. S. Novoselov, D. Jiang, F. Schedin, T. J. Booth, V. V. Khotkevich, S. V. Morozov, and A. K. Geim, Proc. Natl. Acad. Sci. USA 102, 10451 (2005).
  • [64] B. Lalmi, H. Oughaddou, H. Enriquez, A. Kara, S. Vizzini, B. Ealet, and B. Aufray, Appl. Phys. Lett. 97, 223109 (2010).
  • [65] F.-F. Zhu, W.-J. Chen, Y. Xu, C.-L. Gao, D.-D. Guan, C.-H. Liu, D. Qian, S.-C. Zhang, and J.-F. Jia, Nat. Mater. 14, 1020 (2015).
  • [66] Z. Zhu, X. Cai, S. Yi, J. Chen, Y. Dai, C. Niu, Z. Guo, M. Xie, F. Liu, J.-H. Cho, Y. Jia, and Z. Zhang, Phys. Rev. Lett. 119, 106101 (2017).
  • [67] We estimate the exchange splitting of spin-up and -down states by calculating the average difference of their Kohn-Sham eigenvalues. Here, we considered the spin-up and -down states separately occupying 1.67 electrons, which are the number of spin-up electrons occupied below EFE_{F}.
  • [68] H. L. Zhuang, P. R. C. Kent, and R. G. Hennig, Phys. Rev. B 93, 134407 (2016).
  • [69] G. S. Rushbrooke and P. J.Wood, Mol. Phys. 1, 257 (1958); 6, 409 (1963).
  • [70] N. P. Armitage, E. J. Mele, and A. Vishwanath, Rev. Mod. Phys. 90, 015001 (2018).
  • [71] 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).
  • [72] R. Yu, H. Weng, Z. Fang, X. Dai, and X. Hu, Phys. Rev. Lett. 115, 036807 (2015).
  • [73] I. Belopolski, K. Manna, D. S. Sanchez, G. Chang, B. Ernst, J. Yin, S. S. Zhang, T. Cochran, N. Shumiya, H. Zheng et al., Science 365, 1278 (2019).
  • [74] X.-Q. Sun, S.-C. Zhang, and Z. Wang, Phys. Rev. Lett. 115, 076802 (2015).
  • [75] We considered three possible adsorption sites of H atoms on the Hf2S(001) surface (see Fig. S12 in the Supplemental Material [53]). Among them, the adsorption of H atoms on the hollow site is found to be the energetically most favorable configuration.
  • [76] A. R. Mellnik, J. S. Lee, A. Richardella, J. L. Grab, P. J. Mintun, M. H. Fischer, A. Vaezi, A. Manchon, E.-A. Kim, N. Samarth, and D. C. Ralph, Nature (London) 511, 449 (2014).
  • [77] A. Dankert, J. Geurs, M. V. Kamalakar, S. Charpentier, and S. P. Dash, Nano Lett. 15, 7976 (2015).