Open Access Article
Jiajun Zhu
a,
Guangsheng Liu*b and
Duohui Huang*a
aSchool of Mathematics and Physics, Computational Physics Key Laboratory of Sichuan, Yibin University, Yibin, 644000, China. E-mail: hdhzhy912@163.com
bSchool of Materials and Energy, Yunnan University, Kunming, 650091, P. R. China
First published on 2nd October 2025
Two-dimensional non-metallic ferromagnetic states have attracted widespread attention due to their complete spin polarization, long spin relaxation times, and abundant availability, highlighting their potential applications in next-generation spintronic devices. Based on first-principles calculations, we predict the two-dimensional non-metallic ferromagnet T-XN2 (X = Sb, Bi), which has a structure similar to T-MoS2. Calculations of thermodynamic, kinetic, and mechanical properties confirm that they possess good stability. Spin polarization results indicate that the ground state of monolayer T-XN2 is ferromagnetic, with its magnetism arising from the direct p-orbital interactions between N atoms, unlike conventional d/f orbital magnetic materials. Furthermore, the ferromagnetism retains good stability under strain engineering and carrier doping. These results suggest that the magnetic properties of monolayer T-XN2 hold significant fundamental research implications and make it a potential candidate material for non-metallic ferromagnetic devices.
Recently, the layered magnetic material H-MoN2 has been successfully synthesized, attracting significant attention due to its structural similarity to H-MoS2.16 Researchers refer to this nitrogen-rich structure as transition metal dinitrides (TMN2). Theoretically, H-MoN2 is predicted to be a ferromagnetic metal with a Curie temperature above room temperature (420 K).17 Upon hydrogenation, it undergoes a phase transition to the T-MoN2 phase structure, which features a Dirac cone.18 Both T-YN2 and T-TaN2 exhibit characteristics of ferromagnetic semimetals. T-YN2 is a room-temperature p state Dirac semimetal, whereas T-TaN2 has a larger semimetal bandgap (0.72 eV) that prevents spin flipping.19 Additionally, Chen et al. systematically predicted H-phase and T-phase transition metal dinitrides using high-throughput calculations, identifying 87 potentially viable materials and studying their stability.20 In contrast to the aforementioned transition metal dinitrides, main group dinitrides are theoretically predicted to exist in T-PN2 and T-AsN2,21,22 which exhibit ferromagnetic semimetal characteristics, with magnetism derived from the p orbitals of nitrogen atoms. However, related compounds, SbN2 and BiN2, have not yet been reported. Therefore, we are interested in the magnetic and electronic structures of the remaining V-group dinitrides, specifically T-SbN2 and T-BiN2.
Based on this background, we predicted the main group magnetic nitrides T-XN2 (X = Sb, Bi), and investigated their stability and magnetic properties. Through phonon dispersion, first-principles molecular dynamics, and mechanical property analysis, we confirmed that they exhibit good stability. Spin-polarized calculations show that the ground state of monolayer T-SbN2 and T-BiN2 is ferromagnetic, with their magnetism originating from the p-orbitals of the N atoms, where each N atom contributes a magnetic moment of 0.5 μB. Additionally, calculations indicate that the effects of strain engineering and carrier doping on the magnetic properties are highly robust. These results suggest that monolayer T-SbN2 and T-BiN2 are potential candidates for two-dimensional non-metallic spintronic materials.
m1 (No. 164) with a nonpolar point group D3d. The optimized lattice constants a = b are 3.378 Å and 3.491 Å for 1T-SbN2 and 1T-BiN2. The X atom is located at the midpoint of the line connecting two layers of N atoms and is bonded to six equivalent N atoms. The bond length X–N are 2.19 Å (Sb–N) and 2.30 Å (Bi–N), and the vertical height between the two N atoms are 1.99 Å (SbN2) and 2.20 Å (BiN2). The bonding properties of the XN bond can be characterized by Bader charge and the Electron Localization Function (ELF). The calculation results of the Bader charge show that electrons were transferred from the X atom to the N atom, with a 1.05 e per N atom (SbN2) and 0.87 e per N atom (BiN2). We calculated the ELF along the (hkl) = (110) plane, which is displayed in Fig. S1 in the SI. The ELF map shows that electrons are primarily localized around each atom, with only a small amount of electrons existing between the X–N bonds, indicating that the X and N atoms are connected by a weak covalent bond. The above evidence suggest that the X–N bond in monolayer 1T-XN2 is a weak covalent bond. This result is similar to the bonding properties of V–V group compounds that we know.
The stability of the material can be demonstrated through mechanical stability, dynamical stability, and thermodynamic stability. In addition, we have taken into account the ferromagnetism of the material in all calculations related to stability proof. The elastic modulus are typically used to represent the ability of materials to resist deformation and can be used to characterize the mechanical stability of the materials. The elastic modulus of 2D materials can be expressed as:
![]() | (1) |
m1 can be expressed using three elastic constants, denoted as C11, C12, and C66. The calculated results are listed in Table 1.
| Materials | a | lTi–X | ΔQ | C11 | C12 | C66 |
|---|---|---|---|---|---|---|
| 1T-SbN2 | 3.378 | 2.19 | 1.05 | 97.2 | 48.8 | 29.3 |
| 1T-BiN2 | 3.491 | 2.30 | 0.87 | 66.4 | 32.0 | 15.9 |
The three elastic modulus values of T-SbN2 are all greater than those of T-BiN2, indicating that monolayer T-SbN2 is softer than monolayer T-BiN2 and more suitable for applying strain to modulate its physical properties. In addition, the mechanical stability of the 2D hexagonal lattice can be determined by the Born–Huang criterion,30 which must satisfy the C11C22 > C122 and C66 > 0. We further examine whether the monolayer T-XN2 material satisfies the above formula, and the results show that they are mechanically stable.
Phonon dispersion is often used to characterize dynamic stability, and the phonon dispersion of monolayer T-XN2 is presented in Fig. 1(b) and S2 in SI. The calculation results show that all acoustic and optical branches are greater than zero, demonstrating that monolayer T-XN2 is dynamically stable. Furthermore, increasing atomic mass lowers the total phonon energy and increases the band gap between the optical and acoustic branches.
The thermodynamic stability is confirmed by the AIMD simulation of the material's total energy variation over a finite time period.31,32 We simulated the total energy variation of monolayer T-XN2 over 7.5 ps, and the results are presented in Fig. 1(c) and S2 in SI. The calculation results show that the structure at the end of 300 K remains essentially unchanged except for a slight twist, and the total energy fluctuation of the supercell does not exceed 24 meV per atom. The degree of distortion in the structure is represented by the root mean square displacement
between the X–N bonds, where li and l0 represent the bond lengths during AIMD simulations and the bond lengths when not simulated, respectively. The calculated root mean square displacement
values are 0.05 Å and 0.09 Å, indicating that the monolayer T-XN2 are thermodynamically stable at room temperature.
The band structure of non-magnetic T-BiN2 is presented in Fig. 1(b), clearly showing that there are energy bands cross the Fermi level, indicating a metallic state. Two energy bands intersect at the K point, forming a Dirac cone similar to that of single-layer graphene. These two energy bands are primarily composed of the pz orbitals of nitrogen. The density of states and the shape of the energy bands exhibit characteristics reminiscent of a Mexican hat, suggesting that a magnetic phase transition may occur in monolayer T-BiN2
The electron distribution of N and Bi are 2s22p3 and 6s26p3, respectively, making them non-magnetic elements belonging to group V. However, when the Bi atom in the middle layer bonds with the N atoms in the upper and lower layers, the Bi atom will lose 3 electrons from its outer 6p orbital. The N atoms in the upper and lower layers each gain one electron, filling the px and py orbitals along with their original 2p orbitals.21,22 As a result, the net magnetic moment of the system is 1 μB, with each N sharing a magnetic moment of 0.5 μB. The spin-polarization calculation shows that monolayer T-BiN2 exhibits ferromagnetic state with a magnetic moment of 1 μB per unit cell. The spin-polarized density map in Fig. 2(b) indicates that the magnetism originates from the p orbitals of the N atoms. The spin-polarized energy bands are shown in Fig. 2(e), where the two degenerate bands near the Fermi level are split, resulting in channels for spin-up and spin-down, while retaining the Dirac cone at the K point. The projected density of states in Fig. 2(b) indicates that the magnetism primarily originates from the p orbitals of nitrogen.
To further investigate the magnetic properties of monolayer T-XN2, we constructed a 2 × 2 × 1 supercell and designed three different magnetic configurations, as shown in Fig. S3 in SI. Using the formula ΔE = (EAFM − EFM)/4, with the ferromagnetic state as the reference phase, we compared the total energies between different magnetic configurations. It was found that the ground state of monolayer T-XN2 is ferromagnetic, while the antiferromagnetic state is higher than the ferromagnetic state by 50–110 meV per unit cell, indicating that the ferromagnetic state has good stability.
The total magnetic anisotropy energy (MAE) includes contributions from the magnetocrystalline anisotropy (MCA) and the magnetic dipole–dipole anisotropy energy, and the dipole–dipole anisotropy energy contribution can be significant for large-scale systems. In this work we only consider MCA with SOC effect and have treated MCA as the MAE.33,34 We have calculated the MAE under the effect of SOC, which describes the magnetic properties of monolayer T-BiN2, as shown in Fig. 2(c). We define the MAE as the MAE = E100 − E001, where E100 and E001 represent the total energy of ferromagnetic states along [100] and [001] direction. It can be observed that, due to the equivalent energies along the x and y directions, the angular dependence in the xoz plane resembles that in the yoz plane. Numerically, the magnetization direction tends to favor the out-of-plane direction, with an energy lower that is 118.7 μeV lower than that of the in-plane orientation.
The key parameter of ferromagnetic spintronic devices is the Curie temperature (TC), which can be obtained using Monte Carlo simulations based on the Heisenberg model. Its Hamiltonian can be expressed as:35
![]() | (2) |
![]() | (3) |
![]() | (4) |
![]() | (5) |
![]() | (6) |
![]() | (7) |
All relevant magnetic parameters of monolayer T-XN2 are listed in Table 2. The results show that the Curie temperatures of T-SbN2 and T-BiN2 are 155 K and 92 K, respectively, which are lower than those of their counterparts T-PN2 at 385 K and T-AsN2 at 460 K.21,22
| Materials | ΔEN′eel | ΔEzigzag | ΔEstripe | J1 | J2 | MAE | TC |
|---|---|---|---|---|---|---|---|
| T-SbN2 | 57.7 | 71.2 | 91.8 | 136.4 | 0.095 | 97.5 | 155 |
| T-BiN2 | 64.5 | 83.1 | 107.4 | 171.7 | −0.064 | 118.7 | 92 |
Strain engineering is commonly used to manipulate the physical properties of materials, such as magnetism, band structure, topological properties and superconductivity of two-dimensional materials.36,37 In this study, we investigate the effect of in-plane biaxial strain on magnetism. Strain ε can be written as ε = (a/a0 − 1) × 100%, where a and a0 represent the lattice constants after and before stretching, respectively. Here, we focuses on the magnetic properties within the strain range of −5% to 5%. Fig. 3(a) and (e) illustrates the variation of the magnetic moment of monolayer T-SbN2 and T-BiN2 under different strains. It is evident that under the compressive strain, the magnetic moment remains unchanged at 1 μB, while it continuously increases with the increase in tensile strain, reaching 1.07 μB (SbN2) and 1.13 μB (BiN2) at 5%. The total energy changes with strain for different magnetic configurations are shown in Fig. 3(b) and (f). The calculated results indicate that for monolayer T-SbN2, the ferromagnetic state remains the ground state, regardless of whether it is under tensile or compressive strain. Similarly, monolayer T-BiN2 is also ferromagnetic within the strain range we calculated. According to the trend of total energy changes shown in Fig. 3(f), with the continuous increase of compressive strain, monolayer T-BiN2 may transition from a ferromagnetic state to a antiferromagnetic state.
The variation of the Curie temperature of the ferromagnet with in-plane biaxial strain is shown in Fig. 3(d) and (h). For the monolayer T-SbN2 and T-BiN2, the response curves of the Curie temperature to strain are nearly monotonic; monolayer T-SbN2 shows a monotonically decreasing trend, while monolayer T-BiN2 exhibits a monotonically increasing trend. Both are quite sensitive to strain, with a variation range reaching 100–200 K. This change is mainly due to the effect of strain on the magnetic nearest-neighbor coupling parameters, as illustrated in the Fig. 3(c) and (g). The trends in the magnetic nearest-neighbor coupling constant align with the variations in the Curie temperature, and this pattern also exists in the two-dimensional ferromagnetic materials MnN and XN.38 Considering the limited impact of strain on magnetic MAE, the energy difference between the antiferromagnetic and ferromagnetic spin ordered structures induced by strain represents the effective coupling interaction between adjacent magnetic atoms.39 For monolayers T-SbN2 and T-BiN2, this is almost the sole response to structural deformation (the bond length between adjacent magnetic atoms).
Compared to strain engineering, the use of an electrolyte gate can more easily achieve carrier (hole or electron) doping in experiments, while simulation calculations are conducted by removing or adding electrons within the system.40,41 Carrier doping can induce Stoner-type ferromagnetic phase transitions in nonmagnetic materials; therefore, we will investigate the influence of carriers on the ferromagnetic properties of monolayers T-SbN2 and T-BiN2. The range of carrier concentration we calculated is 0 − 5 × 1.01 × 1014 (9.47 × 1013) for T-SbN2 (T-BiN2), which can be extracted or injected experimentally through positive or negative gate voltages, and this concentration range is relatively easy to achieve in experiments. Fig. 4(a) and (e) shows that the magnetic moment of monolayer T-XN2 has an almost linear relationship with carrier concentration, with a slope of 1 μB per 1.01 × 1014 (9.47 × 1013) for T-SbN2 (T-BiN2) concentration, indicating that the doped carriers are fully spin-polarized. Similar to strain engineering, Fig. 4(a) and (e) shows that carrier doping still results in a ferromagnetic ground state for monolayer T-XN2. The difference is that as the electron doping concentration increases, the total energy difference between different magnetic configurations becomes larger, whereas the opposite is true for hole doping. In addition, the ferromagnetic Curie temperature of monolayer T-XN2 varies with carrier concentration as shown in Fig. 4(d) and (h). The trends are generally similar: as the hole concentration increases, the ferromagnetic Curie temperature increases, while an increase in electron concentration leads to a decrease in the ferromagnetic Curie temperature. We also examined the relationship between the magnetic nearest-neighbor coupling constant and carrier concentration. It can be observed that the trend of the magnetic nearest-neighbor coupling constant is almost consistent with that of the ferromagnetic Curie temperature, similar to the effects of strain engineering. Therefore, it can be concluded that carrier doping is one of the effective methods for tuning the magnetic Curie temperature. In summary, monolayer T-XN2 maintains good ferromagnetism and semimetallic properties under the effects of external biaxial strain and carrier doping.
| This journal is © The Royal Society of Chemistry 2025 |