Introduction

Materials combining semiconductivity and magnetism open up possibilities for novel electronic devices that utilise electron spin in addition to charge degrees of freedom.1,2 Ferromagnetic semiconductors (FMSs) are in particular valued for their potential in spintronics for spin-polarised transport. Compared to ferromagnetic metals, FMSs are more suited for injecting spin-polarised electrons into non-magnetic semiconductors.3,4,5,6,7,8 A closely related and technologically important phenomenon is spin filtering, which can be realised through the use of FMSs as the tunneling barrier for generating highly spin-polarised current.9,10,11,12,13,14

FMSs used in spintronics are primarily based on magnetic impurities embedded into conventional non-magnetic semiconductors.15 The robustness of carrier-induced ferromagnetism is extremely sensitive to the growth conditions and processing methods, and the origin of room-temperature ferromagnetism of such diluted magnetic semiconductors remains a subject of debate.2,16 In contrast, concentrated magnetic semiconductors exhibit long-range magnetism without resorting to extrinsic doping. A few concentrated FMSs have been reported, including Cr halides CrBr317,18 and CrI3,19,20,21,22,23 Cr spinel selenides,6 Mn pyrochlore oxides,24 and perovskites such as BiMnO3,25 CuSeO3,26 and YTiO3.27 Among the most studied FMSs for spintronics are the Eu chalcogenides EuX (X = O,S,Se).10,11,12,14,28 While providing very good performances in spin-filter devices, the EuX exhibit very low Curie temperature (e.g., TC = 69 K for EuO29), which is characteristic for most FMSs known to date.

In addition, the electronic structure of FMSs needs to be tailored in the context of spin transport. For a barrierless electrical spin injection depicted in Fig. 1a, the efficiency is determined by the exchange splitting of the conduction band while a low effective electron mass is appreciated for achieving high carrier mobility. Analogously, the exchange splitting is critical for spin filtering as it gives rise to spin-dependent potential barriers for the tunneling current (cf. Fig. 1b), resulting in spin-polarised current in favour of the spin with a lower potential barrier.13,30,31 As such, EuO is in particular attractive for spin injection32 and filtering12 because of its large exchange splitting of the conduction band (0.6 eV) and highly dispersive conduction band.33 Nevertheless, its poor air stability34,35,36,37 along with the low TC present major obstacles for practical applications.

Fig. 1
figure 1

Band diagram schematics of spin-polarised electron injection using a ferromagnetic (FM)/non-magnetic (NM) n–n heterojunction a and of spin filtering achieved with an FM tunneling barrier sandwiched between two NM metal contacts b. For the ferromagnetic semiconductor, the exchange splitting of the conduction band amounts to 2Δex and is depicted by dashed lines

Combining strong ferromagnetism and attractive semiconducting properties in one material is therefore desirable but remains an open problem. Here, we set out to identify systematically concentrated FMSs through a large-scale computational screening of known compounds. We report on the materials identified and especially their semiconducting properties, their Curie–Weiss temperatures, and their stabilities. In particular, we identify the Mn pyrochlore oxide In2Mn2O7 as a very promising material. We discuss its potential use for spin transport and the inherent structural and chemical reasons for its high performances.

Results

We consider a material to be a good FMS candidate if it offers a high ferromagnetic transition temperature and good semiconducting properties. Because electrons have much longer spin lifetimes than holes,2 we focus on spin transport based on electrons as illustrated in Fig. 1, and hence look for FMSs with a large exchange splitting of the conduction band and a low electron effective mass. Starting from the materials project (MP) database comprising over 40,000 density-functional theory (DFT) calculations using the semilocal Perdew–Burke–Ernzerhof (PBE) functional38 and the Hubbard U correction (PBE + U)39 (for transition-metal oxides), we first screen the materials based on their thermodynamic stability (energy above convex hull at 0 K lower than 50 meV per atom) and electronic band gap (>100 meV). This step leads to about 15,300 semiconductors, out of which 3100 compounds show a finite magnetic moment (>0.5 μB) in the ground state (when the computation is initialised in a ferromagnetic state). Among these magnetic materials, only about 1000 compounds exhibit an electron effective mass (\(m_e^ \ast\)) smaller than 1.5 m0. In comparison, typical semiconductors (e.g. GaAs, Si, and ZnO) present \(m_e^ \ast\) ranging from 0.05 to 0.5 m0.40 Figure 2a shows the distribution of \(m_e^ \ast\) for materials exhibiting a finite magnetisation compared to non-magnetic materials. It is clear that low \(m_e^ \ast\) is more easily achieved in non-magnetic compounds. The need for magnetism often implies partially filled d bands. When the conduction-band character is dominated by these orbitals, their localised nature leads to a high effective mass.41 Figure 2b confirms that low \(m_e^ \ast\) materials are mainly of s character. The poor effective mass and strong ferromagnetism are, for instance, present in CrBr3 and certain manganites such as LaMnO3, where a predominant 3d character in the lowest conduction band leads to a high \(m_e^ \ast\) of over 10 m0. At variance, the low \(m_e^ \ast\) of EuO (0.4 m0) is remarkable in that the ferromagnetism arises from an indirect exchange between the localised Eu–4f electrons in the valence and the delocalised 5d/6s electrons in the conduction band.33,42

Fig. 2
figure 2

a Violin plot of electron effective masses for semiconductors in the FM and non-FM configurations. The vertical lines refer to the median value along with the first and third quartiles. b Probability distribution of the orbital characters in the conduction band minimum for FM semiconductors with an electron effective mass smaller than 1.5 m0

The presence of a non-zero total magnetisation in the 0 K DFT computation with an initial ferromagnetic ordering does not imply that the ground state is necessarily ferromagnetic and that this ferromagnetic configuration is sustained at high temperature. We thereby estimate the magnetic ordering of the ~1000 compounds by comparing the total energies of the ferromagnetic ground state to the antiferromagnetic (AFM) or ferrimagnetic (FiM) one. The difference serves as an indicator of whether the compound in question is dominated by ferromagnetic exchange interactions. To determine the magnetic ground state, we use supercells that contain at least four atoms for each distinct magnetic species. An exhaustive search of the lowest-energy AFM (or FiM) configuration is carried out by enumerating all possible configurations in which half of the magnetic sites are initialised with a positive magnetic moment whereas the other half with a negative magnetic moment. The absolute value of the initial magnetic moment follows the calculated magnetic moment in the FM configuration. We consider only the collinear magnetic configurations as non-collinear calculations would be computationally prohibitive at this stage of screening. We find that less than 30 compounds favour an FM ground state by over 10 meV per formula unit compared to the AFM or FiM configurations (see Table S1 of Supplementary Information), manifesting already the difficulty of finding semiconductors with robust ferromagnetism.

Our computational screening thus far relies on the PBE( + U) calculations. While instrumental in determining the energetic stability among various magnetic orderings, PBE and PBE + U with U values calibrated for formation enthalpies do not warrant a faithful description of the underlying electronic structure. For a higher accuracy and a better treatment, particularly of localised d and f electrons, hybrid functionals such as the Heyd–Scuseria–Ernzerhof (HSE) functional43,44 should be used.45 We have thus performed HSE calculations on the candidates exhibiting the most favourable ferromagnetic ordering (The HSE calculations exclude the pyrochlore oxides containing the lanthanide elements with partially filled f electrons due to convergence issues. Nevertheless, these materials are expected to exhibit more exotic magnetic properties than the simple ferromagnetic ordering24). We report in Table 1 the electron effective mass as well as the Curie–Weiss temperature θCW obtained from HSE calculations. The latter is defined from the paramagnetic response at high temperature, and is estimated with the random-phase approximation46 as described in Supplementary Information. When known, we also report on their experimental Curie temperature TC. The difference between θCW and TC indicates the degree of geometrical frustration in a magnetic system.24 Notably, the FMSs listed in Table 1 can be classified into five categories: Eu chalcogenides, Cr spinel chalcogenides, Bi manganites, Mn pyrochlore oxides, and Mn double perovskites. Figure 3 shows the HSE band structure for a representative compound in each category. Our screening recovers the well-known FMSs in the context of spintronics such as EuO, CdCr2Se4, and BiMnO3. Less traditionally associated to spintronics are the Mn pyrochlores (e.g., In2Mn2O7).

Table 1 Properties of identified ferromagnetic semiconductors evaluated with HSE hybrid functional, including the exchange splitting of the conduction band (2Δex), electron effective mass (me), and Curie–Weiss temperature (θCW)
Fig. 3
figure 3

Spin-polarised band structures of the five representative ferromagnetic semiconductors obtained with HSE hybrid-functional calculations. The majority and minority spin channels are shaded by the blue and light red colours, respectively

To compare the performances of these different compounds, we plot in Fig. 4 \(m_e^ \ast\) vs θCW obtained from HSE calculations. We further indicate the stability of the materials against oxidation by showing the maximum oxygen chemical potential reachable while keeping the material thermodynamically stable. This provides a measure of air sensitivity: the higher oxygen chemical potential, the greater stability. The highest θCW is clearly obtained among the double perovskites LaBMnO6 (B = Ni, Co). In particular, La2NiMnO6 shows near room-temperature ferromagnetism arising from the strong ferromagnetic superexchange interactions between the Mn4+ and Ni2+.47 However, the large \(m_e^ \ast\) of over 1.1 m0 could be a limiting factor for high mobility applications. Following La2NiMnO6, the sulfide and selenide spinels ACr2X4 (A = Hg, Cd, Zn, and Mg; X = S, Se) show θCW up to 200 K. The prevalence of Cr3+ can be related to the high magnetic moment of its d3 configuration. The strongest ferromagnetism is observed in CdCr2Se4 in accordance with experiment.48 MgCr2Se4, which has been overlooked as a ferromagnetic spinel in literature, shows comparable ferromagnetism as CdCr2Se4 according to our computational screening. In any case, all these spinel chalcogenides show poor stability in air due to their sulfide or selenide chemistry. The air stability is also an issue for Eu chalcogenides. In fact, EuO is known for the difficulty in growing high-quality thin films since Eu2+ is easily oxidised to Eu3+.34,35,36,37 The remaining oxides are the pyrochlores and BiMnO3. Among the pyrochlores, In2Mn2O7 is especially noteworthy as it shows the highest θCW and the lowest \(m_e^ \ast\). While showing similar electronic and magnetic properties as BiMnO3, In2Mn2O7 exhibits a higher air stability thanks to their high stability of the oxidation states of its cations: In3+ and Mn4+. In comparison with EuO, it offers an even lower \(m_e^ \ast\) (0.29 m0), better air stability, and a significantly higher θCW (155 K vs 76 K).

Fig. 4
figure 4

Figure of merit of the ferromagnetic semiconductors identified through the computational screening. \(\Delta \mu _{{\rm{O}}_2}\) indicates the oxygen chemical potential (referred to the isolated molecule) below which the compound is stable in an oxidising atmosphere. Oxidisation is more likely for compounds with a more negative \(\Delta \mu _{{\rm{O}}_2}\)

The calculated exchange splitting of the conduction band shown in Table 1 for the candidates confirms the good performance in spin filtering with EuO12 and BiMnO3.25 Table 1 implies that In2Mn2O7 should also present an excellent spin-filter effect. But as uncertainty remains in the exchange splitting with the HSE calculations and little is known from experiment, we resort to the self-consistent quasiparticle GW calculations (QSGW) with vertex corrections49 to calculate the electronic structure of In2Mn2O7. The QSGW method does not depend on any adjustable parameter and starting point, and it has been shown to provide a reasonable description of the electronic structure for correlated transition-metal oxides.50 As shown in the QSGW band structure in Fig. 5a, the exchange splitting further opens up to 1.1 eV, in support of using In2Mn2O7 for efficient spin filtering.

Fig. 5
figure 5

Electron density distributions of the lowest conduction band at the Γ point and QSGW band structures for In2Mn2O7 a and Y2Mn2O7 b in the ferromagnetic configuration. The electron density is plotted on the (111) plane centered at an In (Y) atom, whereas the characters associated with the states are resolved by the fat bands mapped onto the atoms. The significant s character in the lowest conduction band of In2Mn2O7 is absent in Y2Mn2O7

Discussion

Our large-scale computational screening shows that the viable routes toward ferromagnetism in semiconducting materials involve either the partially filled Eu–4f electrons or the partially filled 3d electrons of transition metals such as Cr, Mn, and to some extent, V. Indeed, the identified FMSs are mostly Cr spinels and Mn pyrochlores. They are commonly characterised by the high-spin S = 3/2 state in the 3d3 configuration, which in the (pseudo)cubic crystal field results in an occupied t2g and an unoccupied eg manifold of states. For Cr spinels, the strength of ferromagnetism reduces from selenides to sulfides, and eventually inverts to antiferromagnetism for oxides as the ferromagnetic t2geg exchange interaction is outweighed by the AFM t2gt2g interaction.51 While the same competing mechanism is also at play for the pyrochlores, the larger lattice constant stabilises the ferromagnetic configuration for a series of Mn and V pyrochlore oxides. The double perovskites, on the other hand, offer significantly higher TC than the simple perovskite counterparts such as BiMnO3 and LaMnO3. The anomalously strong ferromagnetism of La2NiMnO6 stems from the fully occupied eg state of Ni2+, which is unique to this type of material. In comparison, the eg state is either partially occupied for the Mn3+ in BiMnO3, or simply empty for the Mn4+ and Cr3+ in the case of pyrochlores and spinels.

Our results confirm the challenge in combining adequate air stability, effective mass, and Curie temperature. In2Mn2O7 offers an exceptional compromise between these three metrics. Among the ferromagnetic pyrochlore materials, In2Mn2O7 shows a very low \(m_e^ \ast\) of 0.29 m0, which is among the lowest for all identified FMSs. Such a low effective mass is the result of the prominent In–5s character of the conduction band minimum (CBM) in the minority spin channel, as clearly shown by the element-resolved band structure in Fig. 5a. In contrast, most pyrochlore oxides, such as Y2Mn2O7, exhibit much less dispersive CBM in both spin channels (cf. Fig. 5b) as the Y–5s states do not mix in the lower conduction band. In–5s states are known to lead to dispersive conduction band in binary and ternary oxides:41 one of the highest electron mobility oxide being doped In2O3.

The s character in the conduction band is also at the origin of the strong ferromagnetism present in In2Mn2O7, leading to the highest TC among all pyrochlore oxides. Apparently, the semi-empirical Goodenough–Kanamori rules of superexchange52,53 do not fully account for such strong ferromagnetism as all the pyrochlore oxides considered in Table 1 show Mn–O–Mn bond angles between 130° and 133°. Longer Mn–O bond lengths reduce the AFM t2gt2g interactions among neighbouring Mn atoms, yet this does not explain the higher TC of In2Mn2O7 (dMn−O = 1.89 Å) compared to Y2Mn2O7 (dMn−O = 1.91 Å). Indeed, the hybridisation among Mn(t2g)–O(p)–In(s) states is key to the strong ferromagnetism of In2Mn2O7. Specifically, the In–O covalency mixes with the Mn-t2g–O-p states, stabilising the ferromagnetic configuration by shifting the In–O states upward (downward) in the majority (minority) channel.54 This is supported by the band-resolved crystal orbital Hamilton population (COHP) analysis,55,56,57,58,59 showing the antibonding nature of In(5s)–O and Mn–O interactions at the CBM of the minority spin channel (see Table S2 and Fig. S2 of Supplementary Information). More intuitively, the enhanced ferromagnetism can be understood by the indirect-exchange mechanism60 involving virtual electron hopping from the O–p to the In–s states in the conduction band. This leaves the O–p state effectively spin polarised and enhances the ferromagnetic superexchange through the O atom. For this mechanism to take effect, the atomic valence s state needs to be in a reasonable proximity to the O–p state, which is exactly the case of the group 13 elements such as In and Tl, although Tl2Mn2O7 is a half-metal.54,60,61,62,63 While pyrochlore oxides comprising other group 13 elements (such as B, Al, and Ga) do not appear as a candidate because of their instability, they indeed exhibit a highly dispersive s-like CBM from the minority channel and a high θCW comparable to In2Mn2O7 (see Table S3 of Supplementary Information for the properties of these hypothetical pyrochlore oxides).

Finally, the FMSs need to be n-type to facilitate the transport of spin-polarised electrons. To this end, we assess several dopants in In2Mn2O7, among which Sn and Mo are found to incorporate on the In site while acting as shallow donors, analogous to that in In2O3.64,65 The computational details of defect calculations are described in Supplementary Information, whereas the formation energies of the dopants in various charge states are given in Fig. S1. We additionally find no evidence of favourable self-trapping of electrons as small polarons in this material and a general unfavourability of native compensating centers like cation vacancies, which suggests that In2Mn2O7 can be effectively n-type doped.

In conclusion, we have carried out a large-scale computational screening in quest of concentrated FMSs. Among the very few identified materials, the pyrochlore oxide In2Mn2O7 emerges as a particularly interesting candidate that exhibits robust ferromagnetism, good air stability, and a low electron effective mass, an uncommon combination that is of great promise for high mobility spin transport. While In2Mn2O7 does not yet fulfill the requirement of room-temperature ferromagnetism, its Curie temperature could be potentially increased with epitaxial strain.66,67,68 Indeed, as shown in Supplementary Information, we find that tensile stress due to the lattice mismatch to some semiconductor substrates (such as Si and GaAs) can effectively increase the Curie temperature of In2Mn2O7, but it needs to be practiced with caution as it has adverse effects on the effective mass (see Fig. S3). Other routes, such as doping, can also be explored to enhance the Curie temperature as it has been demonstrated for EuO and BiMnO3.68,69,70,71,72

Methods

First-principles calculations

Collinear spin-polarised semilocal DFT–PBE and hybrid functional HSE calculations are performed with the Vienna ab initio simulation package (VASP).73,74 Electron–ion interactions are described by the projector-augmented-wave (PAW) method.75,76 We use the Pymatgen package77 to generate VASP input files based on the structures retrieved from the MP database. Throughout the calculations, the kinetic energy cut-off is set to 520 eV, and a regular Γ-centered k-point mesh is used with a grid density of 1600 k points per atom. For transition-metal oxides, the PBE calculation is carried out with the Hubbard U correction (PBE + U), for which the U parameters take the values adopted by the MP following the approach described by Wang et al.78

Quasiparticle self-consistent GW calculations are performed with the ABINIT code79,80 using the PseudoDojo optimised norm-conserving pseudopotentials.81,82 Vertex corrections in the dielectric screening are accounted for through the use of the bootstrap exchange-correlation kernel.49,83 The dielectric function is evaluated through the contour deformation method84 including unoccupied states up to 150 eV above the Fermi level in the summations. The dielectric matrix is represented by a plane-wave basis set with an energy cut-off of 160 eV. The self-consistent iteration of the wavefunctions is restricted to the lowest 2Nv states where Nv is the number of the valence bands.

Band-resolved COHP calculations are carried out with a development version of the LOBSTER package.55,56,57,58,59 The pbeVaspFit2015 basis is used with the following basis functions: O: 2s, 2p; In: 5s, 5p, and 4d; Mn: 4s, 3p, and 3d. The wavefunctions are obtained using the PBE + U functional.

Effective mass calculation

The reported effective mass is defined as the conductivity effective mass

$$(m^ \ast )^{ - 1} = \frac{{\sigma (T,\mu )}}{{n(T,\mu )e^2\tau }},$$
(1)

where the electrical conductivity σ and the charge carrier concentration n are computed directly from the Boltztrap calculations85 with T = 300 K and a chemical potential μ leading to n = 1018 cm−3. The relaxation time τ is assumed to be independent of T and μ following previous high-throughput works.41,86