Matter and Radiation at Extremes, Volume. 9, Issue 3, 037402(2024)

Strong electron correlation-induced Mott-insulating electrides of Ae5X3 (Ae = Ca, Sr, and Ba; X = As and Sb)

Ya Xu1... Lu Zheng1, Yunkun Zhang2, Zhuangfei Zhang1, QianQian Wang1, Yuewen Zhang1, Liangchao Chen1, Chao Fang1, Biao Wan1,a) and Huiyang Gou3 |Show fewer author(s)
Author Affiliations
  • 1Key Laboratory of Material Physics of Ministry of Education, School of Physics and Laboratory of Zhongyuan Light, Zhengzhou University, Zhengzhou 450052, China
  • 2School of Mechanical and Equipment Engineering, Hebei University of Engineering, Handan 056038, China
  • 3Center for High Pressure Science and Technology Advanced Research, Beijing 100094, China
  • show less

    The presence of interstitial electrons in electrides endows them with interesting attributes, such as low work function, high carrier concentration, and unique magnetic properties. Thorough knowledge and understanding of electrides are thus of both scientific and technological significance. Here, we employ first-principles calculations to investigate Mott-insulating Ae5X3 (Ae = Ca, Sr, and Ba; X = As and Sb) electrides with Mn5Si3-type structure, in which half-filled interstitial electrons serve as ions and are spin-polarized. The Mott-insulating property is induced by strong electron correlation between the nearest interstitial electrons, resulting in spin splitting and a separation between occupied and unoccupied states. The half-filled antiferromagnetic configuration and localization of the interstitial electrons are critical for the Mott-insulating properties of these materials. Compared with that in intermetallic electrides, the orbital hybridization between the half-filled interstitial electrons and the surrounding atoms is weak, leading to highly localized magnetic centers and pronounced correlation effects. Therefore, the Mott-insulating electrides Ae5X3 have very large indirect bandgaps (∼0.30 eV). In addition, high pressure is found to strengthen the strong correlation effects and enlarge the bandgap. The present results provide a deeper understanding of the formation mechanism of Mott-insulating electrides and provide guidance for the search for new strongly correlated electrides.

    I. INTRODUCTION

    Novel functional materials have provided the basis for many recent developments in science and technology, and among these materials, electrides are an emerging group with rich physical and chemical properties. Electrides constitute a unique class of ionic compounds, with their excess electrons behaving as anions.1 Owing to their distinct crystal and electronic structures, electrides have a wide range of potential applications as catalysts,2 components of electronic devices,3 magnetic materials,4 and battery electrodes.5 In addition, recent research has predicted an abnormal splitting of longitudinal and transverse acoustic modes (LA–TA splitting) in electrides.6 Exploitation of this phenomenon could lead to further, unprecedented applications of electrides. The organic electrides such as Cs+(18-crown-6)2e, which were the first to be discovered, are usually characterized by poor thermal stability and sensitivity to air and water, which limits their practical applications.7 In 2003, Matsuishi et al.8 succeeded in synthesizing the first inorganic electride, [Ca24Al28O64]4+ (4e) (C12A7:e), which can be stabilized at room temperature and has highly localized anionic electrons in cage-like interstitial voids in its lattice.9 Subsequently, electrides with two-dimensional layered voids (e.g., Ca2N10 and Y2C11) and one-dimensional tubular voids (e.g., [La8Sr2(SiO4)6]4+:4e12 and Y5Si313) were discovered. According to the interstitial voids and the distribution of interstitial electrons, electrides are categorized as zero-dimensional (0D), one-dimensional (1D), and two-dimensional (2D).14 In addition, some alkali metals, alkaline earth metals, and compounds containing these elements can also exhibit electride properties under high pressure and are referred to as high-pressure electrides (HPEs). Examples include Na9B,15 Na-hP4,16,17 and Li6C.18

    Semiconducting electrides have a wide range of applications in infrared (IR) photodetectors, hydrogen storage, and fluorine ion battery electrodes, and they have attracted much research interest in recent years.21 The interstitial electrons are dominantly distributed around the Fermi level, which can significantly influence the electronic structure of these electrides. Three representative classes of semiconducting electrides are shown in Fig. 1. Under pressure, the properties of most of the semiconducting electrides are induced by the loosely bound nature of the interstitial electrons, for example, the metal–semiconductor transitions in Ca2N,22 Li,23 Na,17 and Na2He.19 Under ambient conditions, some low-dimensional (0D or 1D) electrides with a high electronegativity difference Ediff, such as Y2Cl320 and C12A7:e,24,25 can also exhibit semiconducting properties. In these electrides, the typical ionic bonding character and low-dimensional interstitial electrons give rise to highly localized energy bands near the Fermi level, separating occupied and unoccupied states. In 2022, McRae et al.21 revealed a unique semiconducting mechanism in 2D electrides, namely, Sc2C and Al2C. A higher electronegativity of the cation can enhance the strength of hybridization between interstitial electrons and cation atoms, producing semiconducting properties.

    Representative semiconducting electrides.19–21

    Figure 1.Representative semiconducting electrides.19–21

    The interstitial electrons can act as ions and form magnetic centers.26 The extremely low work function and unique magnetic properties of magnetic electrides make them promising as spin injection materials and for spintronic device applications.4 Numerous magnetic electrides have been explored so far, such as Y2C,27 Gd2C,26 YCl,20 Ba3MnN3,4 and Nd5Pb3.28 Similar to the typical Mott insulators MnO29 and LaTiO3,30 magnetic electrides with unpaired interstitial electrons can also exhibit Mott-insulating properties. In 2018, Lu et al. reported that α- and β-Yb5Sb3 are Mott-insulating electrides, similar to Sr5P3.31,32 However, there are a few known examples of Mott insulator electrides. The interstitial electrons can hybridize with the d/f orbitals of the surrounding real atoms, but unfortunately the influence of this hybridization interaction on the strong correlation effect remains elusive. Among 1D electrides, the Mn5Si3 structure represents one of the most common structural types.33,34 To date, at least 87 electrides with this type of structure have been reported in the literature (see Fig. 1), and most of them have been experimentally synthesized.35–37 The low-dimensional nature of the Mn5Si3 structure and the abundant number of electrides that possess it provides a broad platform for the search for new Mott insulators.

    In the work reported in this article, we investigated a series of Mott-insulating electrides Ae5X3 (Ae = Ca, Sr, and Ba; X = As and Sb), all of which have been synthesized experimentally and characterized as potential 1D electrides.4 These materials crystallize in the Mn5Si3 configuration with a large interstitial space surrounded by Ae6 octahedrons, and the anionic electrons are localized in a 1D tubular lattice interstitial void. Considering the formal oxidation states of Ae2+ and X3−, one extra electron is trapped in each Ae6 octahedron and arranged in the antiferromagnetic (AFM) configuration. The strong correlation effect between the interstitial electrons induces spin splitting of the energy bands near the Fermi levels, resulting in a greater value of the bandgap (∼0.30 eV). Ca5As3, Sr5As3, and Sr5Sb3 are spin-gapped Mott insulators, similar to Zn(tmdt)2,38 oxide perovskites,39 and β-Yb5Sb3.40 The strong correlation effect between the interstitial electrons can be adjusted through strain–stress and high pressure. The results reveal a great potential of Ae5X3 for a wide range of practical applications.

    II. METHODS

    The structures of Ae5X3 were obtained from the Inorganic Crystal Structure Database (ICSD).41 First-principles calculations were performed using density functional theory (DFT) in the Vienna Ab initio Simulation Package (VASP)42 for structural optimization43 and electronic structure calculations of Ae5X3. Electron exchange-correlation interactions were taken into account using the generalized gradient approximation (GGA) with the Perdew–Burke–Ernzerhof (PBE)44 exchange-correlation functional and the projected-augmented plane-wave (PAW) method.45 The plane-wave cutoff energy was set to 500 eV, and the Monkhorst–Pack method46 was used to delineate the Brillouin zone with a K-point grid spacing of 2π × 0.03 Å−1. Spin-polarized calculations were performed by applying an initial magnetic bias on Ae atoms around the voids (4d site). We separately calculated the different AFM and ferromagnetic (FM) settings of Ae5X3 and found that the lowest energy was achieved with AFM. The calculated spin polarization energy ΔE of Ae5X3 with AFM or FM configuration can be expressed as follows:ΔE=EAFM/FMENM,where EAFM/FM is the energy of Ae5X3 with AFM or FM configuration, and ENM is the energy of Ae5X3 in the nonmagnetic state.

    Owing to the more complex space of d/f electron orbitals and the susceptibility to spin-polarized leaps, strongly correlated electronic systems are corrected by the DFT + U method to account for the Coulomb repulsion between the spin electrons.47,48 Although the Hubbard U cannot act directly on the anionic electrons, it can be used to correct the system effectively by acting on the surrounding Ae atoms.49 Lu et al.32 investigated the bandgap of α-Yb5Sb3 by applying Coulomb repulsion (Hubbard U) to the Yb_d orbital and found that the system was well described at U = 5 eV. In the present work, different Hubbard U values (3 and 5 eV) were applied to the Ae_d orbital (4d site). The hybrid density functional HSE06 was also tested to correct the structure and electronic structure with a mixing parameter of 25% and a screening parameter of 0.2.31

    III. RESULTS AND DISCUSSION

    A. Structural properties

    Figure 2(a) shows the crystal structures of Ae5X3 (Ae = Ca, Sr, and Ba; X = As and Sb) with the hexagonal Mn5Si3 configuration (P63/mcm), with metal Ae atoms occupying the Wyckoff 4d and 6g sites.13 Each Ae atom at the 4d site is coordinated by six X atoms in a linear arrangement along the c direction, and the Ae atoms at the 6g site form the face-sharing Ae6 octahedron, resulting in a 1D tubular interstitial space along the c axis. It should be noted that this structural feature is common in 1D electrides, such as Sc5P3,50 Ca5Pb3,36 and Na3S.51 Two prototypes are usually reported for A5B3 electrides, namely, the Mn5Si3 configuration mentioned above and the β-Yb5Sb3 configuration.32 The β-Yb5Sb3 phase crystallizes in an orthorhombic structure (Pnma), with cationic atoms forming 0D voids in a tetrahedral cage; examples include Sr5Bi3,52 Yb5Bi3,53 and Ca5Sb3.54

    (a) Crystal structure of Ae5X3. (b) Structural prototypes of A5B3 compounds as a function of the radius ratio rM/rX and the electronegativity difference Ediff. The red, brown, and purple spheres represent compounds that adopt the β-Yb5Sb3 prototype, coexistence of β-Yb5Sb3 and Mn5Si3 prototypes, and the Mn5Si3 prototype, respectively. The gray spheres represent phases that have not yet been reported. (c) Volume of Ae6 octahedron as a function of Ae ionic radius. (d) Spin polarization energies ΔE of Ae5X3 with AFM configuration.

    Figure 2.(a) Crystal structure of Ae5X3. (b) Structural prototypes of A5B3 compounds as a function of the radius ratio rM/rX and the electronegativity difference Ediff. The red, brown, and purple spheres represent compounds that adopt the β-Yb5Sb3 prototype, coexistence of β-Yb5Sb3 and Mn5Si3 prototypes, and the Mn5Si3 prototype, respectively. The gray spheres represent phases that have not yet been reported. (c) Volume of Ae6 octahedron as a function of Ae ionic radius. (d) Spin polarization energies ΔE of Ae5X3 with AFM configuration.

    To investigate the structural preferences of A5B3-type compounds, we analyze in detail the structures of A5B3-type compounds from the literature. Figure 2(b) shows structural prototypes of A5B3 compounds as a function of the radius ratio rM/rX and the electronegativity difference Ediff. Here, rM is the ionic radius of a cation atom with +2 state and Ⅵ coordination, and rX is the ionic radius of an anion atom with −3 state and Ⅵ coordination. It is found that when rM/rX < 1.1, A5B3 crystallizes in the β-Yb5Sb3 structure (e.g., Ca5Bi355 and Yb5Bi353); in the range 1.1 < rM/rX < 1.64, β-Yb5Sb3 and Mn5Si3 structures coexist (e.g., Yb5Sb332 and Ca5Sb354); when rM/rX > 1.64, A5B3 prefers the Mn5Si3 configuration (e.g., Ba5Sb352 and Eu5As353). Despite the fact that only the Mn5Si3 configurations of Ba5Bi3 and Dy5Sb3 have been reported experimentally, their rM/rX values suggest that they may also have a β-Yb5Sb3 phase, which might be obtained through high-temperature treatment of Mn5Si3 type phases.32 In 2017, through a combination of structural search and experimental validation, Wang et al.31 successfully synthesized a new 1D electride, Sr5P3, which crystallizes in the Mn5Si3 configuration (rM/rX = 2.68) and was shown to be a Mott insulator. The Sc5P3 was also theoretically reported to adopt the Mn5Si3 structure with rM/rX > 1.70 (rSc in the +3 state was used here).50 Furthermore, we find that the crystallization type of A5B3 is unrelated to Ediff. Considering the stability of Ae5X3 in the Mn5Si3 configuration, we only discuss Mn5Si3-type Ae5X3 in this work, although the results can also apply to β-Yb5Sb3-type Ae5X3.

    The formation of anionic electrons is closely related to the volume of the lattice voids. Larger voids usually induce less orbital hybridization between anionic electrons and d orbitals of neighboring atoms.20Figure 2(c) shows the volume of the Ae6 octahedron along with the Ae ionic radius, and we find that all Ae6 octahedrons in Ae5X3 possess much larger voids (>25 Å3) than those in Sc-based and Y-based electrides,20,50 and thus can hold excess electrons in Ae5X3. The volume of the Ae6 octahedron increases linearly with increasing radius of the Ae ion. In addition, the volume of the Ae6 octahedron decreases as the electronegativity of the X anion increases. This latter phenomenon is due mainly to the higher electronegativity of X endowing Ae with a lower ionic radius, resulting in a larger lattice void. The volume of the Ba6 octahedron in Ba5Sb3 is the largest octahedron volume among the various Ae5X3-type compounds.

    To determine the magnetic ground states of these materials, we calculate different magnetic arrangements by applying an initial magnetic bias to different Ae atoms surrounding the voids. The energy differences between the AFM and FM configurations of Ae5X3 at U = 0, 3, and 5 eV (Fig. S1, supplementary material) indicate that the most stable magnetic configuration of Ae5X3 is the AFM one. Figure 2(d) shows the spin polarization energy ΔE of Ae5X3 with AFM configuration, corrected by the Coulomb repulsion between the electrons using U = 0, 3, and 5 eV, respectively. The negative spin polarization energies of Ae5X3 with the AFM configuration (U = 5 eV) suggest an energy preference for this configuration. We illustrate the calculations of Ba5Sb3, which has been experimentally reported to be an AFM with a bandgap of 0.3 eV.56 When U = 0 or 3 eV, the nonmagnetic (NM) properties of Ba5Sb3 are more stable and the material exhibits a metallic character. The calculated result at U = 5 eV reveals a large negative value of ΔE, suggesting an AFM state, and agreeing well with the experimental results. It has been reported that strong orbital hybridization exists in intermetallic electrides Yb5Sb3 (in both α and β phases), and the interstitial electrons project onto the other energy bands.32 As shown in Fig. 2(d), all the Ae5X3-type compounds discussed here possess lower spin polarization energies at U = 5 eV compared with the intermetallic electride α-Yb5Sb3. This suggests that the bonding characteristics of interstitial electrons have a significant influence on the strong correlation effect.

    The electron localization function (ELF) has been widely used to investigate bonding characters and lone pairs, and it is able to effectively visualize interstitial electrons.57 On the basis of formal charge states (e.g., Ba2+ and Sb3−), the presence of one excess electron and large lattice voids in Ae5X3 satisfies the criterion for an electride. The 3D ELF map of Ba5Sb3 is shown in Fig. 3(a), from which it can be observed that the anionic electrons are distributed in the tubular lattice interstitial voids, exhibiting a quasi-1D electride character. The 2D ELF maps calculated with and without spin polarization for Ba5Sb3 compared with those for the intermetallic electride α-Yb5Sb3 are shown in Fig. 3(b). When the spin polarization is not included, the half-filled interstitial electrons are linearly distributed along the channel voids, forming a 1D electron gas (left). The strong orbital hybridization in the intermetallic electride α-Yb5Sb3 induces delocalization of the interstitial electrons. In the Mn5Si3-type structure, the centers of different Ae6 octahedrons are equivalent and share the same energy level. According to the Pauli exclusion principle and the Hubbard model, the nearest half-filled electrons prefer an AFM configuration.47,48 As in a traditional strongly correlated system, an energy penalty (U > 0 eV) needs to be imposed when the interstitial electrons transfer to neighboring lattice voids. Consequently, the nearest spin-up and spin-down electrons will avoid being at the same lattice site. This property causes adjacent interstitial electrons to form highly localized magnetic centers [Fig. 3(b), right side]. Compared with α-Yb5Sb3, the interstitial electrons are more localized in Ba5Sb3, suggesting a weaker orbital hybridization between the half-filled interstitial electrons and the surrounding atoms, and inducing a pronounced strong correlation effect. Given that the interstitial electrons are dominantly distributed around the Fermi level, the electron localization in Ae5X3 will have a significant influence on the electronic structure, resulting in a wider bandgap.

    (b) ELF of Ba5Sb3. (b) ELF maps of interstitial electrons in Ba5Sb3 and the intermetallic electride α-Yb5Sb3, calculated with and without spin polarization, respectively. (c) Spin charge density of Ba5Sb3. (d) Maximum spin charge density in the voids of Ae5X3 compared with α-Yb5Sb3.

    Figure 3.(b) ELF of Ba5Sb3. (b) ELF maps of interstitial electrons in Ba5Sb3 and the intermetallic electride α-Yb5Sb3, calculated with and without spin polarization, respectively. (c) Spin charge density of Ba5Sb3. (d) Maximum spin charge density in the voids of Ae5X3 compared with α-Yb5Sb3.

    The spin-charge density of Ba5Sb3 is displayed in Fig. 3(c). Although the initial magnetic bias is applied to the Ae 4d site, the spin-charge density around the Ae 4d site is negligible. The magnetic moment originates mainly from the interstitial electrons with an AFM spin arrangement, and the highest value of the spin-charge density is 0.0043 bohr3 in Ba5Sb3. The distribution of spin-polarized interstitial electrons suggests that Ae5X3 should be viewed as potential 0D electrides. These results further confirm the validity of altering the basis sets of surrounding atoms to provide an appropriate description of the electronic structures of interstitial electrons. Figure 3(d) shows the maximum values of spin-charge density in the voids of Ae5X3 compared with α-Yb5Sb3. The Ae5As3 possess relatively larger values of spin-charge density [Fig. 3(d)] and lower spin polarization energies [Fig. 2(d)] than the Ae5Sb3 and α-Yb5Sb3. These results are consistent with the ELF analysis, which implies that high ionic bonding strength can enhance the strong correlation effect.

    B. Electronic structure

    Both the Coulomb interaction correction and hybrid density functionals have been widely used to calculate the electronic structure of Mott insulator.31,48 To further investigate the effect of the calculation method on the electronic structure of Ae5X3, we take Ba5Sb3 as an illustrative example in the following discussion. Figures 4(a)4(d) show the spin-polarized band structures of Ba5Sb3 calculated with the HSE06 hybrid functional and with Hubbard U = 0, 3, and 5 eV, respectively. The weighted band structure and partial density of states (Fig. S2, supplementary material) of Ba5Sb3 (U = 0 eV) reveal that the interstitial electrons form separate bands and that only a minimal number of electrons are projected to other bands, suggesting that the orbital hybridization between the half-filled interstitial electrons and the surrounding atoms is extremely weak. Since the dominant contribution to the energy bands crossing the Fermi level is from interstitial electrons, these can be regarded as “interstitial bands.”20 The band structures of Ba5Sb3 calculated with HSE06, U = 0 eV, and U = 3 eV exhibit a similar character, with half-filled and spin-degenerate “interstitial bands” crossing the Fermi level. In the case of the calculation with U = 5 eV, the energy bands of core electrons remain roughly the same, but the degenerate “interstitial bands” are split and an indirect bandgap of 0.27 eV is opened at the Fermi level, which coincides well with experimental results. Therefore, the semiconducting behavior should be primarily attributed to the Coulomb interactions of the interstitial electrons. Other band structures of Ae5X3 are plotted in Figs. S3 and S4 (supplementary material). It is noteworthy that Ca5As3, Sr5As3, and Sr5Sb3 are spin-gapped Mott insulators, and similar phenomena can also be observed in the molecular material Zn(tmdt)2,38 oxide perovskites,39 NiO,29 and β-Yb5Sb3.40 The experimental and calculated bandgaps of Ae5X3 are shown in Table I in comparison with Sr5P3 and Yb5Sb3. The bandgaps of Ae5X3 are close to 0.30 eV, except for Ca5Sb3 (0.07 eV), about two or three times those of α-Yb5Sb3 (0.14 eV)32 and Sr5P3 (0.10 eV)31 reported previously. On the other hand, the relatively strong correction effects in Ae5As3 endow them with higher bandgaps.

    Spin-polarized band structures of Ba5Sb3 calculated (a) with the HSE06 functional, (b) without Coulomb interaction correction (U = 0 eV), (c) with U = 3 eV, and (d) with U = 5 eV. The calculations were performed for an AFM configuration.

    Figure 4.Spin-polarized band structures of Ba5Sb3 calculated (a) with the HSE06 functional, (b) without Coulomb interaction correction (U = 0 eV), (c) with U = 3 eV, and (d) with U = 5 eV. The calculations were performed for an AFM configuration.

    • Table 1. Calculated and experimental bandgaps of Ae5X3 compared with Yb5Sb3 and Sr5P3.

      Table 1. Calculated and experimental bandgaps of Ae5X3 compared with Yb5Sb3 and Sr5P3.

      PhaseBandgap (eV)
      ExperimentalCalculated
      Ca5As30.29
      Sr5As30.24
      Ba5As30.29
      Ca5Sb30.07
      Sr5Sb30.24
      Ba5Sb30.30560.27
      α-Yb5Sb30.140.01 (U = 5)32
      β-Yb5Sb30.200.07 (U = 5)32
      Sr5P30.100.10 (HSE)31

    To further uncover the formation mechanism of Mott insulators in Ae5X3, we calculated the weighted band structure of interstitial electrons in the nearest Ae6 octahedron sites A [Figs. 5(a) and 5(c)] and B [Figs. 5(b) and 5(c)] in Ba5Sb3. As discussed above [Fig. 4(b)], the energy bands near the Fermi level of Ba5Sb3 are spin-degenerate at A–H and L–H when the Coulomb U is not considered When the Coulomb interactions of interstitial electrons are applied with U = 5 eV, the interstitial electrons in site A [Figs. 5(a) and 5(c)] are spin-split, with the spin-up electrons (unoccupied states) lying above the Fermi level and the spin-down (occupied states) electrons lying below the Fermi level. In comparison, the interstitial electrons in site B [Figs. 5(b) and 5(c)] display the opposite behavior. This spin splitting of occupied and unoccupied states opens the bandgap. The partial charge densities of the occupied states are shown in Fig. 5(d), where it is observed that the anionic electrons have a tubular distribution in the interstitial positions, further confirming that Ba5Sb3 is a potential 1D electride. The maximum value of the partial charge density is 0.0032 bohr3, which is comparable to the maximum value of 0.0043 bohr3 for the spin-charge density calculated in Fig. 3(c), which suggests that the interstitial electrons, serving as ions and essentially in the spin state, form a local magnetic center.

    (a) and (b) Weighted band structures of Ba5Sb3 in sites A and B, respectively. (c) Schematic illustrations of the interaction between anionic electrons distributed in sites A and B. (d) Partial charge density of interstitial states.

    Figure 5.(a) and (b) Weighted band structures of Ba5Sb3 in sites A and B, respectively. (c) Schematic illustrations of the interaction between anionic electrons distributed in sites A and B. (d) Partial charge density of interstitial states.

    C. Bandgap modification

    The bandgap is one of the most critical fundamental parameters for semiconductor-based electronic and optoelectronic devices, and modifying the semiconductor bandgap plays a crucial role in improving the performance of these devices. Common methods of bandgap modification include electric field modulation,58 application of strain–stress,59 atomic doping,37 and the construction of heterostructures.60 Bilayer LaBr2 can change to become semimetallic under a vertical electric field, and spin polarization conversion can be achieved by reversing the electric field.58 The bandgap of Mott-insulating TiPO4 can be adjusted through the application of high pressure.61 In experiments on Y2C weakly magnetic anionic electrons have been observed and it has been found that the magnetic properties of two-dimensional anionic electrons can be modulated by strain–stress and hole doping.27Figure 6(a) shows the bandgap and spin polarization energy of Ba5Sb3 under isotropic strain–stress. The bandgap of Ba5Sb3 is strongly correlated with the spin polarization energy. The bandgap increases and the spin polarization energy decreases under compression (negative lattice deviation). By contrast, when the spin polarization energy increases under tension (positive lattice deviation), the bandgap decreases, and finally disappears, i.e., the material is in a metallic state.

    (a) Bandgap and spin polarization energy of Ba5Sb3 under isotropic strain. (b) ELF maps of Ba5Sb3 with (−5% and +5%) and without (0%) lattice deviation. (c) Bandgaps of Ba5Sb3 under pressure; the inserts are ELF maps of Ba5Sb3 at 2 and 6 GPa. (d) Band structure of Ba5Sb3H.

    Figure 6.(a) Bandgap and spin polarization energy of Ba5Sb3 under isotropic strain. (b) ELF maps of Ba5Sb3 with (−5% and +5%) and without (0%) lattice deviation. (c) Bandgaps of Ba5Sb3 under pressure; the inserts are ELF maps of Ba5Sb3 at 2 and 6 GPa. (d) Band structure of Ba5Sb3H.

    ELF maps of Ba5Sb3 [Fig. 6(b)] show strong localization of anionic electrons at −5% (compression) lattice deviation, while the distribution of anionic electrons is more discrete at 5% (stretch) than at 0% lattice deviation. This suggests that stronger interactions between anionic electrons in electrides with localized magnetic centers are more pronounced and more likely to lead to the formation of Mott insulators. Pressure is an effective way to change the atomic distances between atoms and has essential effects on the behavior of the anionic electrons, such as in the pressure-induced formation of an electron-deficient-type Ca5Pb3 electride36 and the metal-to-semiconductor transition in compressed Ca2N.22 Therefore, we calculated the bandgaps of Ba5Sb3 under different pressures [Fig. 6(c)]. As the pressure increases, the bandgap first increases, reaching a maximum value (0.41 eV) at around 6 GPa, and then decreases. The ELF shows that the electron localization becomes more significant with elevation of pressure, indicating that pressure can effectively enhance the strong correlation effect between the interstitial electrons. However, as the pressure continues to increase, the interatomic distance is reduced, and the corresponding energy bands become more dispersed. Thus, the bandgap decreases when the pressure is increased beyond a certain value. Such results can guide the search for other Mott insulators with highly localized magnetic centers induced by pressure (chemical or physical) in electrides.

    Elemental doping can change the electronic structure and the mechanical and optical properties of materials. On the other hand, electrides usually possess a strong hydrogen affinity, and hydrogen atoms are generally adsorbed at the sites of the anionic electrons in electrides, thereby changing their electronic structures.20,62 For example, C12A7: H is converted from an electride to a conventional ionic compound by hydrogen adsorption.63 The addition of H significantly reduces the work functions of YClH and Y2Cl3H, and the ferromagnetism of YCl disappears.20 Li et al.37 discovered a series of new ternary electrides by appropriate elemental substitution in Mn5Si3-type non-electride materials. As shown in Fig. 6(d), for Ba5Sb3, when H atoms are introduced into the Sr6 octahedron, the electride characteristic and magnetism disappear, but the semiconducting properties remain. The interstitial electrons are transferred to H_1s orbitals and distributed at the deep energy levels.

    IV. CONCLUSIONS

    We have investigated the Mott-insulating properties of Ae5X3 electrides with the Mn5Si3 configuration. Our results show that the interstitial electrons in Ae5X3 adopt an antiferromagnetic configuration and have a quasi-one-dimensional distribution in the face-sharing Ae6 octahedron. Analysis of the spin-charge density and partial charge density reveals that the half-filled interstitial electrons are essentially in a spin state, forming a localized magnetic center. The strong electron correlations between nearest interstitial electrons induce spin splitting and a separation between occupied and unoccupied states, which is responsible for the semiconducting properties. The calculated indirect bandgap of Ae5X3 is about 0.30 eV, which is contributed by the weaker orbital hybridization of the half-filled interstitial electrons with the surrounding atoms. Additionally, Ca5As3, Sr5As3, and Sr5Sb3 are spin-gapped Mott insulators. The bandgaps of Ae5X3 are tunable by application of strain–stress or pressure. The unique semiconducting and magnetic properties of these electrides provide new possibilities for their application in spintronic devices, thermoelectric materials, and electron-emitting devices.

    SUPPLEMENTARY MATERIAL

    ACKNOWLEDGMENTS

    Acknowledgment. This work was supported by the National Natural Science Foundation of China (Grant Nos. 12204419 and 12074013) and the China Postdoctoral Science Foundation (Grant No. 2021M702956).

    Tools

    Get Citation

    Copy Citation Text

    Ya Xu, Lu Zheng, Yunkun Zhang, Zhuangfei Zhang, QianQian Wang, Yuewen Zhang, Liangchao Chen, Chao Fang, Biao Wan, Huiyang Gou. Strong electron correlation-induced Mott-insulating electrides of Ae5X3 (Ae = Ca, Sr, and Ba; X = As and Sb)[J]. Matter and Radiation at Extremes, 2024, 9(3): 037402

    Download Citation

    EndNote(RIS)BibTexPlain Text
    Save article for my favorites
    Paper Information

    Category:

    Received: Nov. 13, 2023

    Accepted: Feb. 23, 2024

    Published Online: Jul. 2, 2024

    The Author Email: Wan Biao (biaowan@zzu.edu.cn)

    DOI:10.1063/5.0187372

    Topics