Next Article in Journal
Insulation Performance and Simulation Analysis of SiO2-Aramid Paper under High-Voltage Bushing
Next Article in Special Issue
Recent Progress in the Design, Characterisation and Application of LaAlO3- and LaGaO3-Based Solid Oxide Fuel Cell Electrolytes
Previous Article in Journal
Long-Term Fluorescence Behavior of CdSe/ZnS Quantum Dots on Various Planar Chromatographic Stationary Phases
Previous Article in Special Issue
Construction of Electrostatic Self-Assembled 2D/2D CdIn2S4/g-C3N4 Heterojunctions for Efficient Visible-Light-Responsive Molecular Oxygen Activation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetic Properties in Mn-Doped δ-MoN: A Systematic Density Functional Theory Study

1
College of Food and Pharmaceutical Engineering, Suihua University, Suihua 152061, China
2
Key Laboratory of Environmental Catalysis and Energy Storage Materials of Heilongjiang Province, Suihua 152061, China
3
College of Materials Science and Chemical Engineering, Harbin University of Science and Technology, Harbin 150080, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(5), 747; https://doi.org/10.3390/nano12050747
Submission received: 13 January 2022 / Revised: 14 February 2022 / Accepted: 18 February 2022 / Published: 23 February 2022
(This article belongs to the Special Issue Nanocomposite Design for Energy-Related Applications)

Abstract

:
Due to the potential applications of transition metal nitrides in modern electronic and spintronic devices, we have systematically studied the magnetic properties of δ-MoN induced by the Mn dopant, with the goal of identifying the origin of magnetism and figuring out the magnetic coupling mechanism between the Mn dopants. Based on the density functional theory, one Mn atom doped at different Mo sites (2a and 6c in the International Tables) in the unit cell of δ-MoN was firstly studied. It was found that the Mn dopant located at the 2a or 6c site leads to significant spin splitting of the density of states, suggesting that the Mn doping induces magnetism in δ-MoN. The calculations were then extended to a 2 × 1 × 2 supercell, which contains two impurity Mn atoms. Detailed analysis reveals that the different couplings of the Mn–Mn pair cannot be simply attributed to the different Mn–Mn distances but are closely related to the electronic processes that take place in the segment (–N– or –N–Mo–N–) that connects two Mn dopants. The mechanisms responsible for the FM/AFM coupling of the Mn–Mn pairs are the superexchange and the pd exchange mediated by the N atoms, and the dd coupling between the host Mo atom and the Mn dopant.

1. Introduction

Transition metal nitrides, by virtue of their unique electronic structure and exact nature of bonding (mixed metallic, ionic, and covalent bonding), possess excellent physical properties, such as high hardness [1,2], high melting point [3], and high electronic conductivity [4,5], which have made them attract tremendous attentions in recent years. In order to enhance the performance of this class of materials, much effort has been devoted to investigating the possibility of improving the energetic [6,7,8], mechanical [9,10], and electrical [11] properties by doping foreign atoms. Recent research has confirmed that introduction of foreign transition metal atoms (for instance, Fe, V, Mn, Cr, etc.) into the host material could effectively enhance the magnetism of transition metal nitrides, such as ScN [12,13], CrN [14,15], and ZnN [16,17]. It has been found that the V, Cr, Mn, Fe, Co, and Ni doping can induce magnetism in ScN [12], and the Mn-doped ScN is a dilute magnetic semiconductor with the Curie temperature above 400 K [18,19]. The CrN undergoes a relative increase in the magnetic order with the substitution of Mn atom [14], and V dopant in CrN could introduce holes into the host material and produce a series of inhomogeneous magnetic or electronic states [15]. The ZnN doped with Cr impurity is found to exhibit half-metallic ferromagnetism, which is a good candidate for spintronic applications [16]. It is known that the enhancement or the inducement of magnetism caused by the transition metal doping is closely related to the modification of electronic states; so, if we want to clearly understand the origin of magnetism, it is necessary to investigate the modified electronic structure of the doped transition metal nitrides.
As a typical type of transition metal nitrides, molybdenum nitrides are known to have a set of interesting properties, such as low compressibility [20], high melting point [21], and excellent catalytic activity [22,23], which are very attractive for a wide range of technological applications. Furthermore, their multiple transport properties (including the metallic and magnetic properties [24,25] and, especially, the high TC superconductivity [26,27]), also support the potential applications in electronic devices. Molybdenum nitrides can crystallize in different phases, including a stoichiometric structure, hexagonal δ-MoN, and two nonstoichiometric structures, cubic γ-Mo2N and tetragonal β-Mo2N [27,28,29,30]. In particular, hexagonal δ-MoN is considered the hardest superconducting metal nitride [1], whose bulk modulus is measured to be 345 GPa [20] and TC is up to 12–15 K [31,32]. A number of experimental works have been performed which mainly focus on the synthesis of pure δ-MoN with desired superconducting properties [27,31,33]. Only a few theoretical studies have been devoted to δ-MoN with the purpose of investigating its structural, electronic, and mechanical properties [34,35]. Up to now, research concerning the magnetism of δ-MoN has been really scarce. Our previous work has theoretically studied the Cr-doping effect on the magnetic and spin transport properties of δ-MoN [36]. However, to the best of our knowledge, no theoretical works have been related to the magnetism of δ-MoN induced by the Mn dopant.
In this paper, we systematically studied the electronic and magnetic properties of the Mn-doped δ-MoN, with the goal of identifying the origin of magnetism and figuring out the magnetic coupling mechanism between the Mn dopants. This paper is organized as follows: Firstly, one Mn atom, with d5 structure, was used as dopant to substitute one Mo atom located at two different nonequivalent positions (2a and 6c in the International Tables) in the unit cell, respectively, to investigate the Mn-doping effects on the structural, electronic, and magnetic properties of δ-MoN. Then, the calculations were extended to a large supercell with two substitutional Mn atoms to explore the favorable coupling between the Mn dopants. Based on the detailed analysis of magnetic ordering, electronic structure, and spin charge density distribution, the origin of magnetism and the coupling mechanism between the Mn dopants were revealed.

2. Computational Details

The calculations were performed using the projector augmented wave (PAW) pseudopotentials as implemented in Vienna ab initio simulation package (VASP) [37]. The exchange correlation functional was treated by Perdew–Burke–Ernzerhof form generalized gradient approximation (GGA-PBE) [38]. Considering the electron correlations in the transition metal d shell, we introduce the GGA + U method by a simplified approach of Dudarev et al. [39], where the effective Hubbard parameter Ueff = UJ. The Hubbard parameter U measures the increase in energy caused by placing an additional electron into a particular site, and J is a screened Stoner-like exchange parameter. Here, we choose Ueff = 4 eV for Mn-3d electrons [40,41] and Ueff = 2 eV for Mo-4d electrons [42], which are taken from the literature. The tetrahedron method with Blöchl corrections was used to determine the partial occupancies for setting each wave function [43]. The Kohn–Sham orbitals were expanded in a plane-wave basis with a cutoff energy of 500 eV. The Brillouin-zone integration was performed on well-converged 7 × 7 × 7 and 3 × 5 × 3 Monkhorst–Pack k-point meshes for 1 × 1 × 1 unit cell and 2 × 1 × 2 supercell, respectively. Both the atomic positions and the unit cell parameters were fully relaxed until the forces on each atom were less than 0.01 eV·Å−1. All the calculations were performed based on the optimized geometries.

3. Results and Discussions

We begin our discussion with the structural, electronic, and magnetic properties of δ-MoN, with a single substitutional Mn atom. Hexagonal δ-MoN crystallizes in a distorted NiAs-type structure with a space group of P63mc (186), which has been determined by experiment investigations [20,32,44], as well as theoretical calculations [34]. In δ-MoN, Mo atoms have two kinds of lattice points: one is labeled 2a and the other is labeled 6c in the International Tables. Each unit cell contains eight Mo atoms and eight N atoms. The unit cell of δ-MoN is presented in Figure 1a, in which the Mo atoms (green balls) are numbered to facilitate our discussion. Mo1 and Mo2 are in the 2a (0, 0, 0) sites, while Mo3–Mo8 are in the 6c (0.5082, 0.0165, −0.0064) sites [32]. To find the preferable location of Mn impurity in δ-MoN, we substitute one Mo atom with one Mn atom at two different nonequivalent Mo sites (2a and 6c), respectively. Here, the side Mo1 atom (representing for the 2a site) and the central Mo6 atom (representing for the 6c site) were substituted, respectively, by one Mn atom to carry out VASP calculations for investigating the Mn-doping effects on the electronic and magnetic properties of δ-MoN. Such a Mn-doped unit cell contains one Mn atom, seven Mo atoms, and eight N atoms, corresponding to the Mn doping concentration of 12.5%. Following that, the two doped systems with Mn doping at 2a and 6c sites are written as Mn-MoN(2a) and Mn-MoN(6c), respectively.
Firstly, we performed a geometry optimization for the pure δ-MoN. It is found that the calculated lattice constants of δ-MoN (a = 5.757 Å, c = 5.668 Å) are in good agreement with the experimental values [32] with deviations within 0.9%. On the basis of the equilibrium structure of δ-MoN, the cells of Mn-MoN(2a) and Mn-MoN(6c) were constructed and relaxed to the minimum energy configurations. The optimized lattice parameters and bond lengths are presented in Table 1 and Figure 1b,c, respectively. The presence of Mn has a marked effect upon the geometry of δ-MoN. After relaxation, the substitutional Mn1/Mn6 atom is found to pull some N atoms closer and push some N atoms farther, suggesting that the lattice structure at the Mn site is distorted and forms a new local structure. The bond lengths of Mn1–N and Mn6–N are in the range of 1.99–2.30 Å and 1.98–2.33 Å for Mn-MoN(2a) and Mn-MoN(6c), respectively, which are different from those of Mo1–N (2.14–2.23 Å) and Mo6–N (2.16–2.22 Å) in the pure δ-MoN. The shortest bond lengths occur at Mn1–N4,5 (1.99 Å) for Mn-MoN(2a) and Mn6–N1,4,5 (1.98 Å) for Mn-MoN(6c), which are very close to the sum of the covalent radius of Mn3+ and N3− (1.92 Å), indicating a covalent bonding feature between Mn and its nearest N atoms. The average bond lengths of Mn1–N and Mn6–N are 2.14 Å and 2.15 Å, respectively, which are about 0.05 Å and 0.04 Å shorter than those of Mo1–N and Mo6–N. The shortening of the average Mn1–N/Mn6–N bond length consequently leads to the shrinkage of the local structure, and thus results in the decrease of the lattice parameters of Mn-MoN(2a)/Mn-MoN(6c), compared to that of δ-MoN (cf. Table 1). This case is probably due to the fact that the radius of Mn3+ is smaller than that of Mo3+.
The charge transfer between Mn dopant and δ-MoN can be examined by the charge density difference, which can be determined by subtracting the charge densities of pristine δ-MoN and isolated Mn atom from the total charge density of the Mn-doped system. Figure 2 shows the charge density difference isosurfaces of Mn-MoN(2a) and Mn-MoN(6c), which describe redistribution of the valence charge density of atoms caused by chemistry bonding. The yellow and light blue regions refer to electron accumulation and depletion, respectively. Evident charge depletion around the Mn atom can be observed from the isosurfaces. Integrating the density of states up to the Fermi energy, the estimated charges transferred from Mn to δ-MoN are 1.22 and 1.12 electrons for Mn-MoN(2a) and Mn-MoN(6c), respectively. The charge accumulation between Mn and N suggests a strong covalent character of the Mn–N bonds, which is in line with the calculated bond lengths, as discussed above. This may be generated by the pd hybridization between Mn and N, which can be further confirmed by the partial density of states.
The preferable site for the Mn dopant in δ-MoN can be determined by estimating the formation energies of different doping configurations. The formation energies of Mn-MoN(2a) and Mn-MoN(6c), E f , are calculated by the following formula [45]:
E f = E tot doped     E tot pure   +   m μ Mo n μ Mn
where Etot(doped) and Etot(pure) are the total energies of the Mn-doped and undoped δ-MoN, respectively. μ Mo ( μ Mn ) is the chemical potential of the Mo (Mn) atom. The n, m are the numbers of the doped Mn atoms and the substituted Mo atoms, respectively. The calculated formation energies of the Mn-doped systems are listed in Table 1. Note that the formation energies of Mn-MoN(2a) and Mn-MoN(6c) decrease relative to the pure δ-MoN. A negative energy means that the formation of Mn-MoN(2a)/Mn-MoN(6c) is spontaneous, that is, the Mn dopant will easily occupy the 2a/6c Mo lattice site in δ-MoN. The relative stability of two Mn-doped configurations can be evaluated by comparing their formation energies. The smaller the formation energy, the more stable the structure. As seen from Table 1, the formation energy of Mn-MoN(2a) is slightly lower than that of Mn-MoN(6c). It seems that the Mn dopant prefers to substitute the 2a Mo lattice site in δ-MoN, and Mn-MoN(2a) is a more energetically favorable structure.
To investigate the effect of Mn dopant on the modification of electronic structure, the spin-polarized density of states (DOS) were calculated for both of the undoped and doped systems, as given in Figure 3. The characteristic of the density of states around the Femi level demonstrates the metallic nature of these systems. It is seen that the majority and minority spin carriers of the pure δ-MoN exhibit mirror symmetry, demonstrating its nonmagnetic characteristic. However, evident spin-polarization of the density of states is observed in the Mn-doped systems. This asymmetrical distribution of the wave functions of the spin-up and spin-down channels on the total DOS of Mn-MoN(2a) and Mn-MoN(6c) suggests that the Mn dopant induces magnetism in δ-MoN. The Mn substitutions at 2a and 6c sites result in 2.74 and 2.56 μB total magnetic moments for Mn-MoN(2a) and Mn-MoN(6c), respectively. The partial DOS analysis reveals that the magnetism of the Mn-doped systems is mainly due to the presence of Mn impurity in the δ-MoN host, as well as a partial contribution from the neighboring Mo and N atoms around the dopant. A local magnetic moment of about 4 μB (4.03 μB for Mn1 and 3.99 μB for Mn6) per Mn atom obtained from the calculations indicates that the Mn ion presents a Mn3+ valence state (3d4). It is notable from the partial DOS that significant hybridization occurs between the orbitals of Mn-3d, Mo-4d, and N-2p, which is a good proof of the much more delocalized Mn impurity bands in comparison to an isolated Mn atom. The Mn-3d states spread throughout the valence band and the conduction band, with almost completely filled majority d bands and nearly unoccupied minority d bands. The strong interactions between Mn and Mo induce evident spin polarization of the Mo atoms with large magnetization. The Mo atoms adjacent to the Mn dopant get as large as −0.51 and −0.47 μB-induced magnetic moments for Mn-MoN(2a) and Mn-MoN(6c), respectively. The neighboring N atoms to the Mn dopant are also slightly spin-polarized with small magnetic moments of −0.03 to 0.02 µB.
It is necessary to investigate the favorable coupling between the Mn dopants. Thus, we extend the analysis to a pair of Mn atoms in δ-MoN. Here, a 2 × 1 × 2 δ-MoN supercell, as shown in Figure 4, was used to investigate the magnetic coupling between two impurity Mn atoms. This supercell contains 32 formula units of MoN, in which two of the Mo atoms are replaced by two magnetic Mn atoms, to form a supercell containing 2 Mn atoms, 30 Mo atoms, and 32 N atoms, amounting to 6.25% Mn doping concentration. Such a large supercell adopted in the calculations allows us to simulate various distributions of two Mn dopants. In brief, we fixed one dopant Mn atom and varied the possible positions of the second Mn atom. As shown in Figure 4, the first Mn atom is fixed at a Mo lattice site (marked 0), and the second Mn atom changes its doping position from 1 to 11 to form the Mn–Mn pair in the supercell. The substituted Mo atoms numbered in 1–11 follow the sequence of the distance to Mo0. This generates eleven different doping configurations, in which the Mn–Mn separation within the supercell varies from 2.73 to 8.08 Å. Hereafter, these arrangements are referred to as the (0, n) configurations.
After geometry optimization, we calculated the formation energies for all the eleven nonequivalent configurations based on Equation (1), and the results are presented in the eighth column of Table 2. The formation energy for the single-doped system, in which is one single Mn atom doped at 0 Mo lattice site, is also calculated for comparison. It is observed from Table 2 that the formation energies of these double-doped systems vary from −3.06 to −3.48 eV, which is much lower than that of the single-doped system (−1.73 eV). This indicates that the formation energy of Mn decreases with its doping concentration, and it is highly favorable to form Mn substitution in δ-MoN.
The interactions responsible for the magnetic ground state of the double Mn-doped systems can be theoretically estimated by calculating the interatomic-exchange constant J, by making use of the total energies in the ordered ferromagnetic (FM) and antiferromagnetic (AFM) structures, EFM and EAFM. Based on the nearest-neighbor Heisenberg model, the value of J can be approximately deduced by ΔE = 4JS2, where ΔE = EAFMEFM and S is the net spin per Mn dopant [46]. For a positive exchange parameter J, the ground state is FM, while the ground state is AFM at a negative exchange parameter J. Table 2 lists the main results of our work concerning the Mn doping for all the eleven (0, n) configurations, and the magnetic moment on each Mn atom is presented in the sixth column. It is seen that the magnetic moments of Mn atoms are almost independent of Mn distributions and the values are close to 4 µB, implying that each Mn atom doping in the supercell possesses about +3e charge with a spin S = 2. Accordingly, the interatomic-exchange constants J were calculated and the variation of J value with the Mn–Mn distance is plotted in Figure 5. Interestingly, a complex oscillatory behavior of the J value as a function of Mn–Mn distance is observed. Among the eleven configurations, five positions ((0, 1), (0, 5), (0, 6), (0, 8), (0, 11)) result in FM and six positions ((0, 2), (0, 3), (0, 4), (0, 7), (0, 9), (0, 10)) result in AFM coupling between two dopant Mn atoms. The lowest energy configuration is found to be an FM state with a J value of 4.8 meV and Mn–Mn distance of 6.44 Å (configuration (0, 8)). Other configurations are higher than the ground state in energy by 68 to 422 meV (cf. Table 2). The oscillation of the J values in Figure 5 indicates that the different magnetic couplings of the Mn–Mn pair cannot be simply attributed to the different Mn–Mn distances. As for the case of FM orientation, the maximum J coupling is reached at the (0, 6) configuration with the Mn0–Mn6 separation of 5.77 Å. Although this Mn–Mn distance is longer than the nearest Mn0–Mn1 distance by 3.04 Å, the coupling strength of Mn0–Mn6 (J = 23.9 meV) is about 8.5 times stronger than that of Mn0–Mn1 (J = 2.8 meV). Compared with the FM orientation, the oscillation magnitude of J value for the AFM orientation is relatively weaker, and the largest J value (−3.3 meV) is achieved at the third-nearest Mn0–Mn3 configuration in which the Mn atoms are separated by 3.88 Å. The above evidence suggests that long-range magnetic coupling exists between the Mn–Mn pair in the supercell. The delocalized character of the Mn-3d states may offer some clues for this long-range magnetic interaction between the Mn dopants. As can be seen from the partial DOS of the double-doped systems in Figure 6, the Mn-3d states are relatively delocalized and strongly hybridize with the electronic states of the host atoms in the supercell, which translates into a long-range effective interaction between the dopant states.
Ferromagnetic materials are characterized by their coercivity (Hc), the magnetic field necessary to reduce the magnetization to zero, and their Curie temperature (TC), the temperature at which random motions cause the magnetization to vanish [47]. Strong spin exchange coupling is a prerequisite for high coercivity [48]. This is because strong exchange interaction makes it hard to flip the spins under an applied magnetic field [48,49]. The strength of Mn−Mn spin exchange can be rationalized by the parameter J. A large positive J value (such as 23.9 meV of (0, 6) configuration) indicates a strong FM spin exchange coupling of the Mn–Mn pair, which may lead to a high coercivity. The Curie temperature TC can be roughly estimated by means of the mean-field approximation and calculated according to the following equation [50]:
T C = 2 Δ E 3 k B
where ΔE is the total energy difference between AFM and FM states and kB is the Boltzmann constant. Using Equation (2), the TC of (0, 1), (0, 6), (0, 8) and (0, 11) configurations are estimated to be higher than the room temperature; especially, the TC of (0, 6) configuration is up to 2951 K. Such high Curie temperatures make the material promising for practical applications.
To gain more insight into the magnetic nature of the double Mn-doped systems, we calculated the spin charge density distributions, as shown in Figure 7. For the (0, 1) configuration, the two Mn dopants are ferromagnetically coupled to each other with the nearest Mn–Mn distance of 2.73 Å considered in our calculations. We find that this Mn–Mn distance is comparable to that of the free Mn2 cluster (2.62 Å) [51]. Thus, a direct interaction of Mn–Mn d-orbital core spins is nonnegligible in the (0, 1) configuration, which gives rise to a bonding state between Mn0 and Mn1 (cf. Figure 7). Furthermore, the three N atoms, which simultaneously connect Mn0 and Mn1, contribute their 2p orbitals to form pd hybridization with two Mn dopants. Evident overlap of N-2p and Mn-3d states could be observed from the partial DOS in Figure 6, demonstrating an indirect Mn–Mn exchange interaction mediated by the N atoms. This pd exchange interaction between N and dopant Mn can also be confirmed by the magnetic moment of about −0.05 µB for the mediated N atoms. According to the Zener model, the direct interaction between d orbitals of adjacent Mn atoms tends to result in an AFM configuration of the d spins. Only when the indirect interaction dominates over the direct coupling between adjacent d orbitals is ferromagnetism present [52]. Therefore, the coupling of the Mn0–Mn1 pair is determined by a competition between the direct dd interaction and the indirect pd exchange. Due to the fact that Mn0 and Mn1 are ferromagnetically coupled, the indirect interaction of d states on different Mn atoms mediated by the p states of N atom is dominant in the (0, 1) configuration.
A different case is found for the configurations with larger Mn–Mn separations ((0, 2) to (0, 11)), in which the two Mn atoms are too far apart (more than 3.18 Å) to allow significant direct overlap of the orbitals and the direct dd coupling between the Mn–Mn pair should be ignored. In such case, the key role underlying the mediated Mn–Mn interaction (FM or AFM) in these configurations should be sought in the electronic processes that take place in the segments connecting two Mn dopants. Here, the mediating segments for the Mn–Mn pairs in the (0, 2) to (0, 11) configurations can be divided into the following two cases:
Mn–N–Mn, (1)
and
Mn–N–Mo–N–Mn (2)
That is, in segment (1) the mediation takes place only by N atom, while in segment (2) the mediated segment is N–Mo–N.
We first focus on the case of the Mn–Mn pair mediated only by the N atom (i.e., two Mn atoms sharing a common N neighbor, forming the segment of Mn–N–Mn), which refers to the configurations of (0, 2) and (0, 3). An indirect Mn–Mn exchange interaction mediated by the N atom is dominant in the two configurations. As seen from Figure 7, the spin polarizations of the mediating N atoms in the two configurations are antiparallel to those of the Mn atoms, and p character of the spin-polarized N orbitals is evident. The mediating N atom provides two different p-orbitals to hybridize with Mn0 and Mn2 (or Mn0 and Mn3), respectively. That is, the spin-up N p-orbital makes a bond with a spin-down Mn orbital, and the spin-down N p-orbital bonds to a spin-up Mn orbital. Superexchange is naturally developed between the Mn dopants along the segment of Mn–N–Mn, thus resulting in the AFM alignment of Mn0 and Mn2 (or Mn0 and Mn3). Superexchange is a mechanism which describes an interaction between moments on atoms too far apart to be connected by direct exchange but coupled over a relatively long distance through the mediating atoms [53]. The extent to which a mediating atom orbital contributes to the overall interaction is governed by the magnitude of its orbital overlap with the coupled metal orbitals, which is influenced by the factors such as the symmetry and the distance between the mediating atom and metal. The weaker magnetic coupling of Mn0–Mn2 (J = –2.9 meV) as compared to that of Mn0–Mn3 (J = –3.3 meV) is mainly attributed to the nearly 90° exchange path (∠Mn0–N–Mn2 ≈ 95°) and the longer Mn–N distances in the mediating segment (the bond lengths of Mn0–N and N–Mn2 are about 0.12 Å and 0.07 Å longer than those of Mn0–N and N–Mn3, respectively).
The situation is more complicated in the case of the mediating segment (2), i.e., Mn–N–Mo–N–Mn, which corresponds to the configurations of (0, 4) to (0, 11). In this case, the mediated Mn–Mn interaction is mainly due to the electronic process that takes place in the segment of N–Mo–N connecting two Mn atoms. Detailed analysis reveals that the presence of Mo atom in the segment plays a critical role in mediating the Mn–Mn magnetic interaction. In fact, owing to the anisotropy of electronic structure and the directional nature of chemical bonding, the magnetic coupling of the Mn–Mn pair differs from one configuration to another. Here, we take (0, 4), (0, 5), and (0, 6) configurations as examples to illustrate the magnetic coupling mechanism of the Mn–Mn pair mediated by the segment of N–Mo–N.
We first pay our attention to the (0, 6) configuration, in which the two Mn atoms are strongly coupled with each other with a relatively large J value of 23.9 meV, indicating a strong FM interaction of Mn0–Mn6 pair. From Figure 4, we notice that the two Mn dopants in the (0, 6) configuration are both located at the ab plane of the supercell, and each substitutional Mn atom is surrounded by six Mo atoms, which forms a slightly deformed hexagon structure. It is evident from Figure 7 that the six surrounding Mo neighbors are all antiferromagnetically coupled to the Mn dopant, and notable magnetic moments locate at the Mo sites, giving rise to an AFM order in the neighborhood of the Mn atom. This AFM environment around the Mn dopant is the origin of the reduction of total magnetic moment, as reflected from Table 2. In particular, the induced magnetic moment of the mediating Mo atom (connecting Mn0 and Mn6) is up to −0.54 µB. The large magnetization of the mediating Mo atom is partially due to the indirect superexchange interaction mediated by the N atom that connects Mn and Mo, i.e., two different N p-orbitals antiferromagnetically hybridize with Mn and Mo, respectively, leading to the induced spin of Mo that is antiparallel to that of Mn. This indirect Mn–N–Mo superexchange interaction resembles the case of Mn–N–Mn, as discussed above. Most importantly, the significant spin polarization of the Mo atom is arising from the direct interaction between the Mo-4d and Mn-3d orbitals, which can be clearly seen from the spin charge density distributions in Figure 7. The three transition metal d-orbitals (one Mo-4d orbital and two Mn-3d orbitals) along the chain of Mn–Mo–Mn couple well with each other to reduce the total energy. The large J value of the (0, 6) configuration is good proof that the direct dd interaction between the dopant Mn and the mediating Mo atom is beneficial to enhance the superexchange interaction of the Mn–Mn pair.
Although the Mn0–Mn4 pair in the (0, 4) configuration both locate at the ab plane as well, the local symmetry around the doping site and the spin polarization of the neighboring Mo atoms to the Mn dopants are different from those of the (0, 6) configuration. Because of the loss of symmetry at the doping site upon geometry optimization, the two mediating Mo atoms move away from their original positions and connect Mn0 and Mn4 with nonequivalent Mn–Mo distances (i.e., the Mn0–Mo distance is more than 3.18 Å, while the Mn4–Mo distance is less than 2.72 Å). It is seen from Figure 7 that the magnetic moments of two mediating Mo atoms are antiparallel to the nearer Mn dopant (Mn4), and, at the same time, are parallel to the farther Mn dopant (Mn0). For the two mediating Mo atoms, the direct dd interaction to Mn4 is significant, but it is really weak to Mn0. The interaction between Mn0 and the mediating Mo atom is mainly mediated by the N atoms. Such nonequilibrium interactions along the mediating segment consequently lead to the AFM coupling of the Mn0–Mn4 pair. By examining the spin charge density distributions of (0, 4) and (0, 6) configurations, we find that the exhibited magnetism of the Mo atoms parallel/antiparallel to that of the Mn dopants is not only correlated with the longer/shorter Mn–Mo distances, but also the coordination orientation of Mo d-orbital that participates in bonding.
As for the (0, 5) configuration, the two Mn dopants are FM coupled to each other along the c direction of the supercell. It is visible from Figure 7 that this long-ranged FM coupling of the Mn0–Mn5 pair is established by the indirect pd exchange between the N p-orbital and the transition metal d-orbitals along the chain of Mn(↑)–N(↓)–Mo(↑)–N(↓)–Mn(↑). One Mn dopant (Mn0) develops a strong bonding with its nearest N atom and induces AFM polarization of the N atom with a magnetic moment of about −0.05 µB, which further leads to a 0.13 µB spin of the mediating Mo atom, reflecting the effect of through-bond spin polarization. The other Mn dopant (Mn5) in turn couples to the mediating atoms (N and Mo) in the same way for an energy gain, resulting in an indirect FM coupling among the two Mn dopants. However, due to the weak hybridization between Mo d-orbital and N p-orbital along the c axis, the mediated interaction through the segment of N–Mo–N in (0, 5) configuration is rather feeble, which leads to a relatively small J value of 0.7 meV.
As stated in the above, the mediated Mn–Mn interaction should be sought in the electronic processes that take place in the segment that connects two Mn dopants. From the observations of (0, 4), (0, 5), and (0, 6) configurations, we find that the spin of the mediating N atom (which makes a bond to the Mn dopant) is always antiparallel to that of Mn regardless of the Mn doping position. The mediating Mo atom can couple to the Mn dopant either ferromagnetically or antiferromagnetically, which mainly depends on the different coupling mechanisms (p–d exchange, superexchange, and dd coupling). The p–d exchange mediated by the N atom could induce the Mo–Mn FM coupling, while the superexchange through the N atom can result in the Mo–Mn AFM coupling. The substantial d–d coupling between Mo and Mn could effectively enhance the magnetic coupling of the Mn–Mn pair, which plays a critical role in mediating the Mn–Mn indirect interaction. The above analysis provides some insight into the origin of magnetic coupling between the Mn dopants. For the configurations in which the two Mn dopants are connected by the segment of N–Mo–N (configurations of (0, 4) to (0, 11)), the indirect FM/AFM Mn–Mn interaction can be described as follows: One Mn dopant induces AFM polarization of its neighboring N atom (one of the N atoms belongs to Mn–N–Mo–N–Mn), and this spin-polarized N atom may polarize the mediating Mo atom in a parallel way or an antiparallel way (through p–d exchange or superexchange). Then, by the similar mechanism, the mediating Mo atom interacts with the other N atom connected to it, and, through which, ferromagnetically or antiferromagnetically couples to the other Mn dopant. That is, one of the Mn dopants dictates the spin polarization of its neighboring N atom, then this N atom affects the spin direction of the mediating Mo atom, and, through which, interacts with the other N atom, and eventually determines the FM or AFM coupling of the Mn–Mn pair. The different coupling mechanisms, i.e., the p–d exchange and the superexchange mediated by the N atom, and the dd coupling between Mo and Mn, are associated with the geometry structure along the segment (such as bond length and bond angle) and the coordination orientation of the bonding orbital. Additionally, it should be emphasized that, except for the mediating atoms, other host atoms located next to the dopants also play important roles: first they become spin-polarized and then they can polarize their nearest neighbors, which facilitates the development of delocalized spin-polarized orbitals over a more extended range.

4. Conclusions

Based on the density functional theory, the structural, electronic, and magnetic properties of δ-MoN with one Mn atom substituted at different Mo sites (2a and 6c in the International Tables) were firstly studied. The substitution of Mn atom results in the local distortion of the structure and the shrink of the unit cell. Formation energy calculations confirm that it is energetically favorable to form Mn substitution in δ-MoN. It is noteworthy that the Mn dopants located at 2a and 6c sites lead to significant spin splitting of the density of states and induce 2.74 and 2.56 μB total magnetic moments, respectively, suggesting that the Mn dopant induces magnetism in δ-MoN. The calculations were then extended to a 2 × 1 × 2 supercell, which contains two impurity Mn atoms, to investigate the favorable coupling between the Mn dopants. Among the eleven different magnetic configurations, five positions result in FM and six positions result in AFM coupling between the Mn–Mn pair. Strong evidence indicate that long-range magnetic coupling exists between the Mn dopants in the supercell. The different couplings of the Mn–Mn pair cannot be simply attributed to the different Mn–Mn distances, but closely related to the electronic processes that take place in the mediating segment (–N– or –N–Mo–N–) that connects two Mn dopants. The mechanisms responsible for the FM/AFM coupling of the Mn–Mn pairs are the superexchange and the pd exchange mediated by the N atoms, and the dd coupling between the Mo and Mn atoms.

Author Contributions

Conceptualization, K.W. and J.Y.; methodology, K.W.; software, K.W., J.Y. and C.C.; validation, K.W. and J.Y.; formal analysis, K.W.; investigation, K.W.; resources, J.Y.; data curation, K.W. and J.Y.; writing—original draft preparation, K.W.; writing—review and editing, K.W., G.Z. and J.Y.; visualization, K.W.; supervision, J.Y. and G.Z.; project administration, J.Y.; funding acquisition, J.Y. and C.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Heilongjiang Province (LH2021B034), the PhD Research Startup Foundation of Suihua University, and the Suihua University Scientific Research Innovation Team Project Funding (SIT05001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, S.; Antonio, D.; Yu, X.; Zhang, J.; Cornelius, A.L.; He, D.; Zhao, Y. The hardest superconducting metal nitride. Sci. Rep. 2015, 5, 1–8. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, M.; Cheng, K.; Yan, H.; Wei, Q.; Zheng, B. Electronic bonding analyses and mechanical strengths of incompressible tetragonal transition metal dinitrides TMN2 (TM = Ti, Zr, and Hf). Sci. Rep. 2016, 6, 1–10. [Google Scholar] [CrossRef] [PubMed]
  3. Andrievski, R.A. High-melting-point compounds: New approaches and new results. Phys. Uspekhi 2017, 60, 276–289. [Google Scholar] [CrossRef]
  4. Jauberteau, I.; Bessaudou, A.; Mayet, R.; Cornette, J.; Jauberteau, J.L.; Carles, P.; Merle-Méjean, T. Molybdenum nitride films: Crystal structures, synthesis, mechanical, electrical and some other properties. Coatings 2015, 5, 656–687. [Google Scholar] [CrossRef] [Green Version]
  5. Simmendinger, J.; Pracht, U.S.; Daschke, L.; Proslier, T.; Klug, J.A.; Dressel, M.; Scheffler, M. Superconducting energy scales and anomalous dissipative conductivity in thin films of molybdenum nitride. Phys. Rev. B 2016, 94, 064506. [Google Scholar] [CrossRef] [Green Version]
  6. Saha, B.; Garbrecht, M.; Perez-Taborda, J.A.; Fawy, M.H.; Koh, Y.R.; Shakouri, A.; Martin-Conzalez, M.; Hultman, L.; Sands, T.D. Compensation of native donor doping in ScN: Carrier concentration control and p-type ScN. Appl. Phys. Lett. 2017, 110, 252104. [Google Scholar] [CrossRef]
  7. Tang, H.; Luo, J.; Tian, X.L.; Dong, Y.; Li, J.; Liu, M.; Liu, L.; Song, H.; Liao, S. Template-free preparation of 3D porous Co-doped VN nanosheet-assembled microflowers with enhanced oxygen reduction activity. ACS Appl. Mater. Interfaces 2018, 10, 11604–11612. [Google Scholar] [CrossRef]
  8. Luo, J.; Tian, X.; Zeng, J.; Li, Y.; Song, H.; Liao, S. Limitations and improvement strategies for early-transition-metal nitrides as competitive catalysts toward the oxygen reduction reaction. ACS Catal. 2016, 6, 6165–6174. [Google Scholar] [CrossRef]
  9. Mu, Y.; Liu, M.; Zhao, Y. Carbon doping to improve the high temperature tribological properties of VN coating. Tribol. Int. 2016, 97, 327–336. [Google Scholar] [CrossRef]
  10. Ye, Y.; Wang, Y.; Chen, H.; Li, J.; Yao, Y.; Wang, C. Doping carbon to improve the tribological performance of CrN coatings in seawater. Tribol. Int. 2015, 90, 362–371. [Google Scholar] [CrossRef]
  11. Ningthoujam, R.S.; Gajbhiye, N.S. Synthesis, electron transport properties of transition metal nitrides and applications. Prog. Mater. Sci. 2015, 70, 50–154. [Google Scholar] [CrossRef]
  12. Sukkabot, W. Structural and magnetic properties of transition-metal doped scandium nitride (ScN): Spin density functional theory. Phys. B 2019, 570, 236–240. [Google Scholar] [CrossRef]
  13. Benissad, F.; Houari, A. Electronic and magnetic properties of transition-metal-doped ScN for spintronics applications. Phys. Status Solidi B 2021, 258, 2000241. [Google Scholar] [CrossRef]
  14. Lone, I.U.N.; Mohamed, S.S.M.; Banu, I.S.; Basha, S.S. Structural, elastic and magnetic properties of Mn and Sb doped chromium nitride—An ab initio study. Mater. Chem. Phys. 2017, 192, 291–298. [Google Scholar] [CrossRef]
  15. Quintela, C.X.; Rivadulla, F.; Rivas, J. Electronic and magnetic phase diagram of Cr1xVxN. Phys. Rev. B 2010, 82, 245201. [Google Scholar] [CrossRef]
  16. Sirajuddeen, M.M.S.; Banu, I.S. An ab initio study on electronic and magnetic properties of Cr, V doped Cd and Zn nitrides for spintronic applications. J. Magn. Magn. Mater. 2016, 406, 48–59. [Google Scholar] [CrossRef]
  17. Sirajuddeen, M.M.S.; Banu, I.S. Electronic and magnetic properties of Fe (Mn)-doped Cd and Zn nitrides for spintronic applications: A first-principles study. J. Mater. Sci. 2015, 50, 1446–1456. [Google Scholar] [CrossRef]
  18. Herwadkar, A.; Lambrecht, W.R. Mn-doped ScN: A dilute ferromagnetic semiconductor with local exchange coupling. Phys. Rev. B 2005, 72, 235207. [Google Scholar] [CrossRef]
  19. Herwadkar, A.; Lambrecht, W.R.; van Schilfgaarde, M. Linear response theoretical study of the exchange interactions in Mn-doped ScN: Effects of disorder, band gap, and doping. Phys. Rev. B 2008, 77, 134433. [Google Scholar] [CrossRef]
  20. Soignard, E.; McMillan, P.F.; Chaplin, T.D.; Farag, S.M.; Bull, C.L.; Somayazulu, M.S.; Leinenweber, K. High-pressure synthesis and study of low-compressibility molybdenum nitride (MoN and MoN1−x) phases. Phys. Rev. B 2003, 68, 132101. [Google Scholar] [CrossRef]
  21. Lévy, F.; Hones, P.; Schmid, P.E.; Sanjinés, R.; Diserens, M.; Wiemer, C. Electronic states and mechanical properties in transition metal nitrides. Surf. Coat. Technol. 1999, 120, 284–290. [Google Scholar] [CrossRef]
  22. Tagliazucca, V.; Leoni, M.; Weidenthaler, C. Crystal structure and microstructural changes of molybdenum nitrides traced during catalytic reaction by in situ X-ray diffraction studies. Phys. Chem. Chem. Phys. 2014, 16, 6182–6188. [Google Scholar] [CrossRef] [PubMed]
  23. Zhong, Y.; Xia, X.; Shi, F.; Zhan, J.; Tu, J.; Fan, H.J. Transition metal carbides and nitrides in energy storage and conversion. Adv. Sci. 2016, 3, 1500286. [Google Scholar] [CrossRef] [PubMed]
  24. Papaconstantopoulos, D.A.; Pickett, W.E.; Klein, B.M.; Boyer, L.L. Electronic properties of transition-metal nitrides: The group-V and group-VI nitrides VN, NbN, TaN, CrN, MoN, and WN. Phys. Rev. B 1985, 31, 752. [Google Scholar] [CrossRef] [PubMed]
  25. Gajbhiye, N.S.; Ningthoujam, R.S. Structural, electrical and magnetic studies of nanocrystalline δ-MoN and γ-Mo2N. Phys. Status Solidi C 2004, 1, 3449–3454. [Google Scholar] [CrossRef]
  26. Ihara, H.; Hirabayashi, M.; Senzaki, K.; Kimura, Y.; Kezuka, H. Superconductivity of B1-MoN films annealed under high pressure. Phys. Rev. B 1985, 32, 1816. [Google Scholar] [CrossRef]
  27. Inumaru, K.; Nishikawa, T.; Nakamura, K.; Yamanaka, S. High-pressure synthesis of superconducting molybdenum nitride δ-MoN by in situ nitridation. Chem. Mater. 2008, 20, 4756–4761. [Google Scholar] [CrossRef]
  28. Maoujoud, M.; Binst, L.; Delcambe, P.; Offergeld-Jardinier, M.; Bouillon, F. Deposition parameter effects on the composition and the crystalline state of reactively sputtered molybdenum nitride. Surf. Coat. Technol. 1992, 52, 179–185. [Google Scholar] [CrossRef]
  29. Inumaru, K.; Baba, K.; Yamanaka, S. Superconducting molybdenum nitride epitaxial thin films deposited on MgO and α-Al2O3 substrates by molecular beam epitaxy. Appl. Surf. Sci. 2006, 253, 2863–2869. [Google Scholar] [CrossRef]
  30. Inumaru, K.; Baba, K.; Yamanaka, S. Synthesis and characterization of superconducting β-Mo2N crystalline phase on a Si substrate: An application of pulsed laser deposition to nitride chemistry. Chem. Mater. 2005, 17, 5935–5940. [Google Scholar] [CrossRef]
  31. Bezinge, A.; Yvon, K.; Muller, J.; Lengaeur, W.; Ettmayer, P. High-pressure high-temperature experiments on δ-MoN. Solid State Commun. 1987, 63, 141–145. [Google Scholar] [CrossRef]
  32. Bull, C.L.; McMillan, P.F.; Soignard, E.; Leinenweber, K. Determination of the crystal structure of δ-MoN by neutron diffraction. J. Solid State Chem. 2004, 177, 1488–1492. [Google Scholar] [CrossRef]
  33. Zhang, Y.; Haberkorn, N.; Ronning, F.; Wang, H.; Mara, N.A.; Zhuo, M.; Chen, L.; Lee, J.H.; Blackmore, K.J.; Bauer, E.; et al. Epitaxial superconducting δ-MoN films grown by a chemical solution method. J. Am. Chem. Soc. 2011, 133, 20735–20737. [Google Scholar] [CrossRef] [PubMed]
  34. Sahu, B.R.; Kleinman, L. Theoretical study of structural and electronic properties of δ-MoN. Phys. Rev. B 2004, 70, 073103. [Google Scholar] [CrossRef]
  35. Zhao, E.; Wang, J.; Wu, Z. Displacive phase transition, structural stability, and mechanical properties of the ultra-incompressible and hard MoN by first principles. Phys. Status Solidi B 2010, 247, 1207–1213. [Google Scholar] [CrossRef]
  36. Yu, J.; Wang, K.; Qiao, X.; Tian, J.; Zhang, G.; Guo, Q. Effects of Cr doping in δ-MoN: Structural, magnetic and spin transport properties. Theor. Chem. Acc. 2020, 139, 1–9. [Google Scholar] [CrossRef]
  37. Madsen, G.K.; Blaha, P.; Schwarz, K.; Sjöstedt, E.; Nordström, L. Efficient linearization of the augmented plane-wave method. Phys. Rev. B 2001, 64, 195134. [Google Scholar] [CrossRef]
  38. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [Green Version]
  39. Dudarev, S.L.; Botton, G.A.; Savrasov, S.Y.; Humphreys, C.J.; Sutton, A.P. Electron-energy-loss spectra and the structural stability of nickel oxide: An LSDA+ U study. Phys. Rev. B 1998, 57, 1505. [Google Scholar] [CrossRef]
  40. Wang, L.; Maxisch, T.; Ceder, G. Oxidation energies of transition metal oxides within the GGA+ U framework. Phys. Rev. B 2006, 73, 195107. [Google Scholar] [CrossRef] [Green Version]
  41. Krcha, M.D.; Janik, M.J. Examination of oxygen vacancy formation in Mn-doped CeO2 using DFT+ U and the hybrid functional HSE06. Langmuir 2013, 29, 10120–10131. [Google Scholar] [CrossRef] [PubMed]
  42. Wang, Y.; Puggioni, D.; Rondinelli, J.M. Assessing exchange-correlation functional performance in the chalcogenide lacunar spinels GaM4Q8 (M = Mo, V, Nb, Ta; Q = S, Se). Phys. Rev. B 2019, 100, 115149. [Google Scholar] [CrossRef] [Green Version]
  43. Blöchl, P.E.; Jepsen, O.; Andersen, O.K. Improved tetrahedron method for Brillouin-zone integrations. Phys. Rev. B 1994, 49, 16223. [Google Scholar] [CrossRef] [PubMed]
  44. Cendlewska, B.; Morawski, A.; Misiuk, A. Superconducting MoNx prepared by isostatic direct nitriding at high pressure and high temperature. J. Phys. F Met. Phys. 1987, 17, L71. [Google Scholar] [CrossRef]
  45. Cantele, G.; Degoli, E.; Luppi, E.; Magri, R.; Ninno, D.; Iadonisi, G.; Ossicini, S. First-principles study of n-and p-doped silicon nanoclusters. Phys. Rev. B 2005, 72, 113303. [Google Scholar] [CrossRef]
  46. Yang, K.; Dai, Y.; Huang, B.; Whangbo, M.H. Density functional studies of the magnetic properties in nitrogen doped TiO2. Chem. Phys. Lett. 2009, 481, 99–102. [Google Scholar] [CrossRef]
  47. Oyama, S.T. Introduction to the chemistry of transition metal carbides and nitrides. In The Chemistry of Transition Metal Carbides and Nitrides; Oyama, S.T., Ed.; Springer: Dordrecht, The Netherlands, 1996; pp. 1–27. [Google Scholar]
  48. Zhang, Y.; Miller, G.J.; Fokwa, B.P. Computational design of rare-earth-free magnets with the Ti3Co5B2-type structure. Chem. Mater. 2017, 29, 2535–2541. [Google Scholar] [CrossRef]
  49. Ruiz-Díaz, P.; Stepanyuk, O.V.; Stepanyuk, V.S. Effects of interatomic coupling on magnetic anisotropy and order of spins on metallic surfaces. J. Phys. Chem. C 2015, 119, 26237–26241. [Google Scholar] [CrossRef] [Green Version]
  50. Seña, N.; Dussan, A.; Mesa, F.; Castaño, E.; González-Hernández, R. Electronic structure and magnetism of Mn-doped GaSb for spintronic applications: A DFT study. J. Appl. Phys. 2016, 120, 051704. [Google Scholar] [CrossRef]
  51. Guo, Y.; Chen, M.; Guo, Z.; Yan, X. First-principles calculations for magnetic properties of Mn-doped GaN nanotubes. Phys. Lett. A 2008, 372, 2688–2691. [Google Scholar] [CrossRef]
  52. Pearton, S.J.; Norton, D.P.; Ivill, M.P.; Hebard, A.F.; Zavada, J.M.; Chen, W.M.; Buyanova, I.A. Ferromagnetism in transition-metal doped ZnO. J. Electron. Mater. 2007, 36, 462–471. [Google Scholar] [CrossRef]
  53. Hurd, C.M. Varieties of magnetic order in solids. Contemp. Phys. 1982, 23, 469–493. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic structure of the unit cell of δ-MoN, and optimized bond lengths of (b) Mn1–N and (c) Mn6–N in Mn-MoN(2a) and Mn-MoN(6c), respectively. For comparison, the relaxed bond lengths of Mo1–N and Mo6–N in the pure δ-MoN are also presented in (b,c), respectively. The pink, green, and blue balls denote the Mn, Mo, and N atoms, respectively.
Figure 1. (a) Schematic structure of the unit cell of δ-MoN, and optimized bond lengths of (b) Mn1–N and (c) Mn6–N in Mn-MoN(2a) and Mn-MoN(6c), respectively. For comparison, the relaxed bond lengths of Mo1–N and Mo6–N in the pure δ-MoN are also presented in (b,c), respectively. The pink, green, and blue balls denote the Mn, Mo, and N atoms, respectively.
Nanomaterials 12 00747 g001
Figure 2. Charge density difference isosurfaces for (a) Mn-MoN(2a) and (b) Mn-MoN(6c). The pink, green, and blue balls denote the Mn, Mo, and N atoms, respectively.
Figure 2. Charge density difference isosurfaces for (a) Mn-MoN(2a) and (b) Mn-MoN(6c). The pink, green, and blue balls denote the Mn, Mo, and N atoms, respectively.
Nanomaterials 12 00747 g002
Figure 3. Total and partial DOS of δ-MoN, Mn-MoN(2a), and Mn-MoN(6c). The Fermi level is set as zero.
Figure 3. Total and partial DOS of δ-MoN, Mn-MoN(2a), and Mn-MoN(6c). The Fermi level is set as zero.
Nanomaterials 12 00747 g003
Figure 4. A 2 × 1 × 2 supercell structure of δ-MoN. With respect to one Mn atom (labeled as 0), we select eleven nonequivalent positions for the other Mn atom (as numbered from 1 to 11) based on the symmetry of the supercell. The green and blue balls denote the Mo and N atoms, respectively.
Figure 4. A 2 × 1 × 2 supercell structure of δ-MoN. With respect to one Mn atom (labeled as 0), we select eleven nonequivalent positions for the other Mn atom (as numbered from 1 to 11) based on the symmetry of the supercell. The green and blue balls denote the Mo and N atoms, respectively.
Nanomaterials 12 00747 g004
Figure 5. Variation of J value as a function of Mn–Mn distance. The positive value of J represents the FM state, while the negative value represents the AFM state.
Figure 5. Variation of J value as a function of Mn–Mn distance. The positive value of J represents the FM state, while the negative value represents the AFM state.
Nanomaterials 12 00747 g005
Figure 6. Total and partial DOS of the selected double-doped systems. The Fermi level is set as zero.
Figure 6. Total and partial DOS of the selected double-doped systems. The Fermi level is set as zero.
Nanomaterials 12 00747 g006
Figure 7. Spin charge density distribution on the planes containing the Mn–Mn pair and their mediating Mo and N atoms. Positive spin density is represented by solid lines, while negative spin density is represented by dashed lines and marked by colors.
Figure 7. Spin charge density distribution on the planes containing the Mn–Mn pair and their mediating Mo and N atoms. Positive spin density is represented by solid lines, while negative spin density is represented by dashed lines and marked by colors.
Nanomaterials 12 00747 g007
Table 1. Optimized lattice parameters of δ-MoN before and after Mn substitution, and formation energies of Mn-MoN(2a) and Mn-MoN(6c).
Table 1. Optimized lattice parameters of δ-MoN before and after Mn substitution, and formation energies of Mn-MoN(2a) and Mn-MoN(6c).
Configurationsa (Å)c (Å)c/aVolume (Å3)Ef (eV)
δ-MoN5.7575.6680.985162.7
Mn-MoN(2a)5.7265.6260.983159.5−1.71
Mn-MoN(6c)5.7145.6440.988159.6−1.68
Table 2. The relaxed Mn–Mn distance (d), the energy difference (ΔE) between FM and AFM states, the energy relative to the ground state (ΔEground), the total magnetic moment of the supercell (Mtotal), the magnetic moments of two Mn dopants (M1 and M2), and the formation energies (Ef) for various double Mn-doped configurations.
Table 2. The relaxed Mn–Mn distance (d), the energy difference (ΔE) between FM and AFM states, the energy relative to the ground state (ΔEground), the total magnetic moment of the supercell (Mtotal), the magnetic moments of two Mn dopants (M1 and M2), and the formation energies (Ef) for various double Mn-doped configurations.
Configurationsd
(Å)
ΔE
(meV)
ΔEground
(meV)
Mtotal
(μB)
M1/M2
(μB)
CouplingEf
(eV)
(0, 1)2.73442834.493.88/4.01FM−3.20
(0, 2)3.18−46422−0.344.03/−4.05AFM−3.06
(0, 3)3.88−52265−0.414.01/−4.05AFM−3.22
(0, 4)5.01−392360.074.08/−4.06AFM−3.25
(0, 5)5.65112103.933.98/3.98FM−3.27
(0, 6)5.77382683.673.99/3.99FM−3.41
(0, 7)6.07−193810.054.05/−4.06AFM−3.10
(0, 8)6.447703.384.00/4.00FM−3.48
(0, 9)6.47−15376−0.354.03/−4.06AFM−3.11
(0, 10)7.25−36220−0.964.00/−4.03AFM−3.26
(0, 11)8.08541463.503.99/3.99FM−3.34
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, K.; Yu, J.; Chi, C.; Zhang, G. Magnetic Properties in Mn-Doped δ-MoN: A Systematic Density Functional Theory Study. Nanomaterials 2022, 12, 747. https://doi.org/10.3390/nano12050747

AMA Style

Wang K, Yu J, Chi C, Zhang G. Magnetic Properties in Mn-Doped δ-MoN: A Systematic Density Functional Theory Study. Nanomaterials. 2022; 12(5):747. https://doi.org/10.3390/nano12050747

Chicago/Turabian Style

Wang, Keda, Jing Yu, Caixia Chi, and Guiling Zhang. 2022. "Magnetic Properties in Mn-Doped δ-MoN: A Systematic Density Functional Theory Study" Nanomaterials 12, no. 5: 747. https://doi.org/10.3390/nano12050747

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop