Next Article in Journal
Algae-Enhanced Electrospun Polyacrylonitrile Nanofibrous Membrane for High-Performance Short-Chain PFAS Remediation from Water
Previous Article in Journal
Metal-Promoted Higher-Order Assembly of Disulfide-Stapled Helical Barrels
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetoelastic and Magnetoelectric Coupling in Two-Dimensional Nitride MXenes: A Density Functional Theory Study

1
Department of Physics, National Tsing Hua University, Hsinchu 30013, Taiwan
2
Institute of Physics, Academia Sinica, Taipei 11529, Taiwan
3
Physics Division, National Center for Theoretical Sciences, Taipei 10617, Taiwan
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(19), 2644; https://doi.org/10.3390/nano13192644
Submission received: 31 August 2023 / Revised: 22 September 2023 / Accepted: 23 September 2023 / Published: 26 September 2023

Abstract

:
Two-dimensional multiferroic (2D) materials have garnered significant attention due to their potential in high-density, low-power multistate storage and spintronics applications. MXenes, a class of 2D transition metal carbides and nitrides, were first discovered in 2011, and have become the focus of research in various disciplines. Our study, utilizing first-principles calculations, examines the lattice structures, and electronic and magnetic properties of nitride MXenes with intrinsic band gaps, including V2NF2, V2NO2, Cr2NF2, Mo2NO2, Mo2NF2, and Mn2NO2. These nitride MXenes exhibit orbital ordering, and in some cases the orbital ordering induces magnetoelastic coupling or magnetoelectric coupling. Most notably, Cr2NF2 is a ferroelastic material with a spiral magnetic ordered phase, and the spiral magnetization propagation vector is coupled with the direction of ferroelastic strain. The ferroelectric phase can exist as an excited state in V2NO2, Cr2NF2, and Mo2NF2, with their magnetic order being coupled with polar displacements through orbital ordering. Our results also suggest that similar magnetoelectric coupling effects persist in the Janus MXenes V8N4O7F, Cr8N4F7O, and Mo8N4F7O. Remarkably, different phases of Mo8N4F7O, characterized by orbital ordering rearrangements, can be switched by applying external strain or an external electric field. Overall, our theoretical findings suggest that nitride MXenes hold promise as 2D multiferroic materials.

1. Introduction

Intrinsic two-dimensional (2D) multiferroics have garnered significant attention in the research community due to their potential for downsizing and integration in various applications, allowing for the manipulation of ferroic orders through magnetoelectric, piezoelectric, or magnetoelastic interactions [1]. The experimental discovery of ferromagnetism in 2D CrI3 has spurred a rapid growth in fabricating 2D magnetic materials, including ferromagnets such as VI3 [2], CrBr3 [3], VSe2 [4], VTe2 [5], NbTe2 [6], MnSe2 [7], and Cr2Ge2Te6 [8,9]; and antiferromagnets such as CrCl3 [10], FePS3 [11], and MnPS3 [12], and exotic magnets such as α -RuCl3 [13], which exhibit Kitaev quantum spin liquid behavior. Additionally, numerous methods for controlling magnetization in 2D materials have been demonstrated, including electrostatic gate control of spin waves [14] and magnetic order [15,16] in bilayer CrI3, electrostatic gate control of the easy-magnetic axis in multilayer Cr2Ge2Te6 [17], and spin-current control of the magnetic order in CrSBr [18].
Considering their rich diversity in structure and composition, MXenes are promising 2D materials to realize multiferroicity. Tahir et al. [19] reported the first observation of ferroelectric and multiferroic properties in a delaminated Ti3C2Tx MXene film. Moreover, several theoretical studies have predicted intrinsic ferroelectricity or magnetism in some MXenes, such as i-MXene (Ta2/3Fe1/3)2CO2 as a type-I multiferroic material [20], Hf2VC2F2 as a type-II multiferroic material [21], and magnetic Janus MXenes, that can control their magnetic order via an out-of-plane electric field [22]. The recent advances in the synthesis of mono- or few-layer MXenes such as molten salt etching methods [23], halogen-based etching methods [24], the intercalation–alloying–expansion–microexplosion mechanism [25], and delamination based on the water freezing and thawing (FAT) method [26] have promised the production of large-lateral-size and high-quality MXenes that go beyond the well-studied Ti3C2Tx. These new techniques give us the flexibility to modulate the composition of MXenes, including the control of functional group termination, and thus, examine the theoretical prediction of multiferroic properties in some MXenes. Nevertheless, there are only a limited number of reports concerning the fabrication of nitride MXenes, with the examples of V2NTx [27], Mo2NTx [27], Ti2NTx [28], and Ti4N3Tx [29]. These reports primarily emphasize the fabrication processes of nitride MXenes and highlight their enhanced electrical conductivity compared to their carbide MXene counterparts. However, there is a notable absence of experimental studies regarding their magnetic properties and the possibility of multiferroicity within them.
Ab initio density functional theory (DFT) simulations can facilitate further exploration of 2D multiferroics for studying the electronic, magnetic, and mechanical properties of materials at a quantum mechanical level. Kumar et al. [30] carried out a systematic study on the magnetic properties of the nitride MXene M2NT2 and predicted that high-Curie-temperature ferromagnetism exists in multiple nitride MXenes, where M is an early transition metal (Ti, V, Cr, or Mn), N is Nitrogen, and T is a surface terminal group (O, OH, or F). Their results also hinted at the presence of orbital ordering in magnetic nitride MXenes, but did not fully elucidate its nature. Orbital ordering [31] is a common occurrence in oxides and plays important roles in many fascinating physical phenomena such as superconductivity [32,33,34,35], colossal magnetoresistance [36,37,38], and ferroelectric polarization [39]. Furthermore, it has been shown that electric fields can be used to control orbital ordering, Jahn–Tellar distortions [40], and the magnetic anisotropy of multiferroic PTO/LTO superlattices [41].
In this work, DFT simulations were used to investigate the structural, electronic, and magnetic properties of nitride MXenes M2NT2, with transition metals (M = V, Cr, Mn, or Mo), selected for their magnetic properties, and different surface terminations (T = O or F) chosen as functional groups to passivate the reactive transition metal surface. Our results show that the nitride MXenes V2NF2, V2NO2, Cr2NF2, Mo2NO2, Mo2NF2, and Mn2NO2 are semiconductors, with band gaps ranging from 0.20 eV to 2.1 eV, and exhibit intrinsic magnetism and orbital ordering. We found out that orbital ordering plays an important role in mediating magnetoelastic coupling in Cr2NF2 and magnetoelectric coupling in V2NO2 and the Janus MXene Mo8N4F7O.

2. Methods (Computational Approach)

2.1. Density Functional Theory

Density functional theory (DFT) calculations were performed using the Vienna ab initio simulation package (VASP) [42] with projector-augmented wave (PAW) pseudopotentials [43] and the Perdew–Burke–Ernzerhof (PBE) exchange-correlation functional [44]. Electron correlations in 3d and 4d electrons were treated using the rotational invariant on-site Coulomb parameters in Dudarev’s formulation [45]. Herein, the effective U values for V [46,47], Cr [46,48], Mn [46], and Mo [49,50,51] atoms are set to be 3 eV. A vacuum space of 20Å was used to prevent the interactions between the adjacent images along the vertical direction. Γ -centered 12 × 12 × 1 Monkhorst–Park k-point meshes were employed for structural optimizations and calculations of electronic properties, with the cutoff energy of the plane-wave basis set to 600 eV. To determine the ground state configuration, we performed structural optimization on the in-plane lattice constants and internal atomic coordinates of a 2 × 2 supercell with 5 different initial magnetic configurations, including four antiferromagnetic phases (AFM1, AFM2, AFM3, and AFM4 in Figure 1) and a ferromagnetic phase (FM), until the Hellman–Feynman force on each atom was less than 5 × 10 3 eV/Å. The nudged elastic band method [52] was used to determine the paths and energy barriers of ferroelectric switching, and ferroelectric polarization was calculated using the Berry phase method [53]. The ISOTROPY tool [54] was used to aid with the symmetry and space group analysis.

2.2. LKAG Formalism and Monte Carlo Simulations

To calculate the isotropic exchange parameters, J i j , of two-dimensional nitride MXenes, we first derive the tight binding Hamiltonian based on the maximally localized Wannier function (MLWF) with the Wannier90 package [55,56], and then calculate the J i j by treating local spin rotation as a perturbation, using LKAG formalism [57] as implemented in the TB2J package [58]. These values of J i j were later used as the input parameters for the Monte Carlo (MC) simulations to verify the magnetic ground states and estimate the transition temperatures, using the Metropolis algorithm as implemented in the code SPIRIT [59]. The simulations adopted the Heisenberg spin model with a magnetic anisotropy term:
H = J i j i j S i · S j A i ( S i z ) 2
with classical spin S i , j of length one at atomic sites i, j; single-ion magnetic anisotropy constant, A; component of the spin along the easy-magnetization axis, S i z . A 2D supercell of 50 × 100 (20,000 magnetic atoms) with a periodic boundary condition was adopted for the simulations, and for each temperature 2 × 10 5 steps were taken for thermalization and another 2 × 10 5 steps were performed for sampling.

3. Results and Discussion

3.1. Orbital Ordering, Electronic, and Magnetic Properties

The crystal symmetry analysis using ISOTROPY revealed that the space groups of the nitride MXenes V2NO2, CrNF2, MoNF2, and MnNO2 are monoclinic P 2 1 / m (Figure 2b), while the space groups of V2NF2 and MoNO2 are triclinic P 1 ¯ (Figure 2c). Group theory analysis [60] using ISODISTORT shows that the distortion of the high-symmetry parent structure of space group P 3 ¯ m 1 (Figure 2a) into the P 2 1 / m phase can be decomposed into three modes: the interplane antipolar displacements of T along the c-axis, with irreducible representation Γ 1 + ( a , 0 , 0 ) (Figure 2d); the strain and interplane antipolar displacements of M and T along the a-axis, with irreducible representation Γ 3 + ( a , 3 a , 0 ) (Figure 2e); and the intraplane antipolar displacements of M, N, and T along either the a- or c-axis, with irreducible representation M 2 ( 0 , a , 0 ) (Figure 2g). In the P 3 ¯ m 1 parent phase, the M ions form an ideal octahedral complex with the X and T ions. The Γ 1 + ( a , 0 , 0 ) displacements move the T ions closer to the M ions, thus shortening the M-T bonds; while Γ 3 + ( a , 3 a , 0 ) displacements move one T vertex closer to the central M ion and another two T vertices further from the central M ion, resulting in two longer M-T bonds and one shorter M-T bond in every octahedral complex. Finally, the M 2 ( 0 , a , 0 ) intraplane displacements shorten both M-X and M-T bonds in one metal complex while lengthening the M-X and M-T bonds in another, thus regrouping the M ions into two sublattices: sublattice 1 and sublattice 2 (Figure 2b). The M ions in sublattice 1 have longer average bond lengths and, thus, lower oxidation numbers, while sublattice 2 has shorter average bond lengths and higher oxidation numbers. Using Cr2NF2 as an example of structure P 2 1 / m , (Figure 3a,b) show that for both sublattice 1 and sublattice 2, the metal complexes have Jahn–Teller distortions, where the two axial bonds have different lengths when compared with the corresponding equatorial bonds.
The distortions of the P 1 ¯ structure can be decomposed into four modes: Γ 1 + ( a , 0 , 0 ) mode (Figure 2d), Γ 3 + ( 0 , a , 0 ) mode (Figure 2f), M 2 ( 0 , a , 0 ) mode (Figure 2g), and M 1 ( 0 , a , 0 ) mode (Figure 2h). The Γ 1 + ( a , 0 , 0 ) mode and M 2 ( 0 , a , 0 ) mode are identical to their counterparts in the distortion of the P 1 ¯ structure, while the Γ 3 + ( 0 , a , 0 ) mode’s antipolar displacements and strain are along the a -axis and this strain along the a -axis deforms the structure to a non-orthogonal cell. Other than that, there are additional intraplane antipolar displacements of M, N, and T along the b-axis with irreducible representation M 1 ( 0 , a , 0 ) . The M 1 ( 0 , a , 0 ) intraplane displacements cause every metal–ligand bond in the octahedral complex to have different lengths.
In order to better understand the electronic properties of the nitride MXenes, the ligand’s p-orbitals and the transition metal’s d-orbitals’ resolved band structures were plotted along high-symmetry paths in the Brillouin zone (Figure 4a–f). The results revealed that V2NF2, V2NO2, Cr2NF2, Mo2NO2, and Mo2NF2 exhibited the characteristics of a Mott transition, i.e., the valence band maximum (VBM) and conduction band minimum (CBM) were found to be primarily comprised of d electron bands. The band gap of these Mott insulating nitride MXenes is approximately 1 eV if the t 2 g orbitals are partially occupied, and approximately 2 eV for those with fully occupied t 2 g orbitals. In contrast, the band structure of Mn2NO2 shows a p d charge-transfer type of band gap, with the VBM being a mixture of e g orbitals from Mn and p orbitals from O, and the CBM dominated by e g and t 2 g orbitals.
Next, we analyze the nature of the orbital ordering by computing the magnetic properties. Table 1 summarizes the magnetic moments of metal ions (in nitride MXenes) at sublattice 1 and sublattice 2 sites, along with their ground state magnetic orders based on DFT calculations. The magnetic moments of the metal ions are calculated using Wannier functions based on the Wannier90 results of their respective magnetic ground states; as expected, the magnitudes of the magnetic moments of the metal ions are larger (smaller) at sublattice 1 (sublattice 2) as the number of occupied d orbitals is higher (lower). Since these transition metal complexes are in distorted octahedral or distorted tetrahedral symmetry, the d orbitals cannot be classified as pure e g or t 2 g orbitals, but rather a mixture of the two. For instance, the d-orbital occupancy of Cr2+ at the sublattice 1 of Cr2NF2 (Figure 3a) was found to be about three electrons per site, predominantly comprising a mixture of d x y , d y z , and d z x orbitals (Figure 3c). On the other hand, for Cr3+ at sublattice 2 (Figure 3b), the degeneracy of e g orbitals was lifted, with the d z 2 orbital largely occupying the CBM; while the electronic states in the energy range of approximately 0.5 to 1 eV below the Fermi level were mainly occupied by d x y , d z x , and d y z orbitals (Figure 3d).
Table 1 also summarizes the magnetic transition temperatures of the nitride MXenes as determined by Monte Carlo simulations. According to the Monte Carlo simulations, the magnetic ground states of V2NF2, V2NO2, Mo2NO2, and Mo2NF2 exhibit an AFM3 pattern, which agrees with the prediction of the DFT calculations. On the other hand, the Monte Carlo simulations predicted Cr2NF2 and Mn2NO2 to have non-collinear magnetic structures. The simulations revealed that Mn2NO2 displays a magnetic configuration similar to conical spin spiral (Figure 5a), with the magnetization at the atomic site k of sublattice i in layer j, with R k ( i , j ) approximated by M k ( i , j ) = M i ( sin ( θ i ) cos ( q · R k ( i , j ) + ϕ i , j ) , sin ( θ i ) sin ( q · R k ( i , j ) + ϕ i , j ) , cos ( θ i ) ) , where θ 1 ( 2 ) ≈ 43 (20) and q ( π a , π b , 0 ) , with a and b representing the lengths of the lattice vector along the a - and b -axes, respectively. The initial phase of magnetization, ϕ i , j is summarized in Figure 5c. It is worth noting that the long-range magnetic ordering in V2NF2, Mo2NO2, and Mo2NF2 is predicted to prevail above room temperature. Such room temperature magnetic behavior in nitride MXenes was also predicted by Kumar et al. [30], where Cr2NF2 and Mn2NT2 (T = F, O, and OH) are reported to show ferromagnetism, with Curie temperatures well above room temperature; pristine and La-doped Ti3C2 MXenes are reported to have ferromagnetism and antiferromagnetism coexisting at temperatures as high as 300 K [61]. Currently, most of the experimentally discovered 2D ferro-/antiferromagnets have their Curie/Néel temperatures limited below room temperature [2,3,5,6,7,8,9,10,11,12,13], with only VSe2 [4] and MnSex [7] showing strong ferromagnetism above 300 K. Our results and previous studies suggest that the MXenes are an important family of 2D materials to search for a new generation of room temperature ferro-/antiferromagnets.
For Cr2NF2, the Monte Carlo simulation at 0 K predicted a magnetic configuration (Figure 5b) that is similar to an out-of-plane Néel-type spin spiral, with the magnetization M k ( i , j ) = M i ( cos ( q · R k ( i , j ) + ϕ i , j ) , 0 , sin ( q · R k ( i , j ) + ϕ i , j ) ) , where q = ( 0.2 × 2 π a , 0 , 0 ) , where a represents the length of the lattice vector along the a -axis. The formation of an intralayer spin spiral magnetization in Cr2NF2 is induced by frustrated isotropic couplings between Cr atoms, which can be described by a simplified model involving four values of the interatomic exchange coupling parameter J (Figure 6a–d). The strong antiferromagnetic couplings (J1 = −55 meV and J2 = −52 meV) between interlayer Cr atoms explain three key features of the Monte Carlo simulation results: (i) nearly antiparallel magnetization between Cr atoms from successive layers; (ii) nearly parallel magnetization between intralayer Cr atoms along the b-axis; and (iii) a small initial phase difference ( Δ ϕ 30 ) (Figure 5d) between intralayer Cr atoms within the same unit cell. The frustrated isotropic couplings are attributed to the competing interactions between intralayer ferromagnetic coupling J3 (18 meV) and interlayer ferromagnetic coupling J4 (28 meV), which favor ferromagnetic and antiferromagnetic couplings of intralayer Cr atoms along the a-axis and b-axis, respectively. Since magnetization of intralayer Cr atoms within the same unit cell are highly in phase, and the wavevector q is parallel to the a-axis, the value of q can be approximated using the classical JN-JNN model for one-dimensional frustrated ferromagnets, i.e., qa = | J N | /(4JNN) [62]. Using JNN = 2 × 18 meV and JNN = 28 meV, we obtain qa = 0.19, which is similar to the prediction of the Monte Carlo simulation (qa = 0.20). Interestingly, although the interacting ion pairs of J5 are separated by comparable distances with similar atomic environments when compared with those of J1 and J2, J5 has a much weaker antiferromagnetic coupling (−4.4 meV) relative to J1 and J2. This unusual weak coupling is caused by orbital ordering in Cr2NF2, and is explained in detail in the next paragraph.
To understand the origin of antiferromagnetic and ferromagnetic interactions among Cr ions, we explain the magnetic interactions using two processes: the antiferromagnetic direct exchange process and the ferromagnetic double exchange process (Figure 6a–d). For J1 and J2, the interatomic distances between interacting ions are 2.772 Å and 3.149 Å, respectively, and the major contribution to the antiferromagnetic coupling originates from the direct exchange process between the Cr ions; the contribution from double exchange to J1 is negligible since the overlap between p-ligands and d-metals is lacking due to the non-occupancy of d x 2 y 2 orbitals in metal ions; on the other hand, the distance between the connecting ligand and Cr ions in J2 coupling is too large, making double exchange processes unfavorable.
For the ferromagnetic J3 interaction, a double exchange process arises as the electron hops between the occupied d y z orbital in Cr3+ and unoccupied d z 2 orbital in Cr2+, mediated by a non-magnetic N ion. While for the double exchange process in J4’s interaction, an electron from the occupied d z 2 in Cr2+ couples with the unoccupied d z 2 in Cr3+, through a non-magnetic N ion. Finally, a strong direct exchange process is expected in J5, as the separation between the interacting ions is only 2.946 Å. Nevertheless, a ferromagnetic double exchange process also arises from the electron coupling between the occupied d y z orbital in Cr3+ and unoccupied d z 2 orbital in Cr3+ at the opposite site. The competing interaction between the double exchange process and direct exchange process results in the weak coupling strength of J5 (−4.4 meV).

3.2. Magnetoelastic Coupling in Cr2NF2

The material Cr2NF2 possesses a significant ferroelastic strain (10.9%), i.e., it has three equally stable orientation variants in the lattice structure that can be switched by applying an external stress. The ferroelastic strain is defined as ( ( b a 1 ) × 100 % ) , with lattice parameters a = 6.03 Å and b = 6.689 Å, calculated along the three diagonal lines ( a 1 , a 2 , and a 3 ) of a deformed hexagon (Figure 7a). The shorter lattice parameter, a, is laying along a 1 , and the propagation vector of the spin spiral magnetization is perpendicular with respect to a 1 , showing coupling between the ferroelastic strain and magnetic order.
In addition to the previously mentioned orientation state (Figure 7a), there are two equivalent orientation states that can be achieved in Cr2NF2, by switching the values of a 2 (Figure 7b) or a 3 (Figure 7c) to 6.03 Å. The transitions between the two orientation states are analyzed using the nudged elastic band (NEB) method and result in an activation energy barrier of 56 meV per atom. In comparison to other reported 2D ferroelastic materials, Cr2NF2 has a low ferroelastic strain relative to BP5 (41.4%) [63], borophane (42%) [64], and phosphorene (37.9%) [65], but is similar to AgF2 (13.5%) [66] and t-YN (14.4%) [67]. The activation energy barrier of Cr2NF2 is also moderate, comparable to AgF2 (51 meV) [66] and t-YN (33 meV) [67], but lower than that of BP5 (320 meV) [63], borophane (100 meV) [64], and phosphorene (200 meV) [65]. These results suggest that the moderate activation energy barrier of Cr2NF2 allows for experimental manipulation of its ferroelasticity while maintaining stability, which could facilitate the control of its spin spiral magnetization.

3.3. Orbital-Ordering-Induced Magnetoelectric Coupling

It is noteworthy that structural optimization has revealed the existence of stabilized ferroelectric (FE) phases in the materials V2NO2, Cr2NF2, and Mo2NF2. These FE phases are characterized by energies that are 31 meV, 70 meV, and 47 meV higher than the energies of their corresponding ground states, respectively.
As an illustration, we use V2NO2 to highlight the role of orbital ordering in magnetoelectric coupling among the aforementioned MXenes. Figure 8a,b present contour plots of the spin density of V2NO2 in the antiferroelectric (AFE) phase (ground state) and FE phase (excited state), respectively. Both phases exhibit AFM3 orders. The transition from the AFE phase to the FE phase results in a significant increase in the average metal–ligand bond length of the VL ion (circled in black in Figure 8a,b) in the lower layer and a decrease in the average metal–ligand bond length of the VU ion (circled in red in Figure 8a,b) in the upper layer. This in turn leads to the transformation of VL and VU ions from V4+ and V3+ to V3+ and V4+, respectively. As a consequence, the distribution of V4+ and V3+ ions becomes asymmetrical, with the ratio of V4+ to V3+ being 1:3 and 3:1 in the upper and lower layers of V2NO2, respectively.
To further verify the stability of the AFM3 order of V2NO2, we compared the energies of the FE phase at different magnetic orders and performed Monte Carlo simulations. The results confirmed that the AFM3 order remains the most stable magnetic configuration. The energy required to switch from the AFE to the FE phase is estimated to be 51 meV per V atom using the nudged elastic band (NEB) method. The Berry phase method calculates the polarization along the a- and c-axes to be 6.9 µC/cm2 and 2.3 µC/cm2, respectively, expressed in bulk form by considering its thickness. Since the magnetic moment of V3+ (≈2 µ B ) is larger than that of V4+ (≈1 µ B ), the asymmetrical distribution of V4+ and V3+ ions in the FE phase results in a net average magnetic moment of 0.25 µ B per V atom, demonstrating the coupling between the electrical and magnetic degrees of freedom.
Nevertheless, the excitation energy (32 meV) of the ferroelectric phase of V2NO2 is likely too high to be stabilized experimentally. One potential method to induce spontaneous ferroelectric polarization is to fabricate Janus MXenes, which are MXene structures with different compositions of the upper and lower functional groups. The absence of inversion symmetry in Janus MXenes leads to a large dipole moment. However, this type of ferroelectric polarization usually cannot be switched by an external electric field, which limits its application in nanoelectronic devices.
In this study, we demonstrate that ferroelectric polarization and magnetization in the Janus MXene Mo8N4F7O can be altered by an external electric field or external strain. We found out that the Janus MXenes V8N4O7F, Cr8N4F7O, and Mo8N4F7O have excited states characterized by orbital ordering rearrangements, and the excitation energies of these Janus MXenes (18.5 meV, 18.6 meV, and 6.6 meV, respectively) are much lower than those of their non-Janus counterparts, i.e., V2NO2, Cr2NF2, and Mo2NF2. Interesting results are found in the transition of Mo8N4F7O from the ground state phase (P1 phase) to the excited phase (N1 phase). Our results reveal that Mo8N4F7O is a bipolar magnetic semiconductor (BMS) in both the P1 and N1 phases. BMSs are characterized by a VBM and CBM that are fully spin polarized in opposite spin channels (Figure 9a,b), offering an ideal platform to manipulate the polarization of spin current using gate voltage [52,68,69]. In the P1 (N1) phase, the VBM is dominated by electronic contributions from lower (upper)-layer Mo ions, while the CBM is dominated by electronic contributions from upper (lower)-layer Mo ions, with a band gap of 1.4 (0.4) eV. The transition from the P1 to N1 phase also results in an increase in the ferroelastic strain, ( b a 1 ) × 100 % (Figure 9c), from 6.0% to 13.1%, with a P 1 = 6.57 Å and b P 1 = 6.96 Å, while a N 1 = 6.38 Å and b N 1 = 7.22 Å.
By linearly interpolating a and b between ( a P 1 , b P 1 ) and ( a N 1 , b N 1 ), we are able to further reduce the excitation energy until the N1 and P1 phases become nearly energetically degenerate (around a ferroelastic strain of 10.2%). To illustrate the orbital ordering rearrangements upon the transition, we show the (contour plot of) spin density of the P1 phase (Figure 9c) and highlight the changes upon transition (Figure 9d,e). In the P1 phase, the Mo atoms in the lower layer (75% F and 25% O as surface functional groups) have a higher average magnetic moment than the Mo atoms in the upper layer (with only F as a surface functional group), thus give a net magnetic moment of +0.25 µ B per Mo. Upon transition to the N1 phase, the bond length between MoU (circled in red in Figure 9c) and the nearest N ion along the negative z-direction (of local coordinate system for d-orbitals) increases significantly, thereby increasing the occupancy of the d z 2 orbital and magnetic moment at the MoU site; since the N ion now receives less electrons from the MoU site, MoL ions (circled in black in Figure 9c) donate additional electrons to the N ion, resulting in a slight decrease in occupancy of d x y orbitals and the magnetic moments at the MoL site. As a result, the transition from the P1 phase to N1 phase changes the average magnetic moment from +0.25 µ B per Mo to −0.25 µ B per Mo.
Because of the broken inversion symmetry in Mo8N4F7O and the difference in electronegativity of O and F, we expect an intrinsic out-of-plane electric field directed from the lower layer to the upper layer. This prediction is confirmed by our calculations of electronic energies under an applied electric field (Figure 9f): both P1 and N1 phases have their electronic energies decrease (increase), with the applied electric field directing upward (downward). However, the electronic energy of the N1 phase decreases more significantly with an upward applied electric field, and increases less significantly with a downward electric field, when compared with the energy of the P1 phase under the same conditions, implying that the overall strength of the intrinsic electric field is higher in the N1 phase relative to the P1 phase. From Figure 9c, we can see that in the P1 phase, the total ionic charge of Mo ions from the upper layer is higher than from those in the lower layer; such an asymmetrical distribution of low and high ionic charge on the Mo ions results in an intrinsic out-of-plane electric field that is opposite to the intrinsic electric field due to the asymmetrical surface functional groups, thus reducing the overall strength of the intrinsic electric field. On the other hand, the intrinsic electric field, due to both the asymmetrical distribution of low/high-ionic-charge Mo ions and of surface functional groups, is pointing upwards, thus enhancing the overall strength of the intrinsic electric field. The results show that at a ferroelastic strain of 10.2%, the energies of Mo8N4F7O in the P1 and N1 phases become degenerate, where switching of the phase can be achieved by application of an external electric field, making non-volatile electric control of spin polarization possible. It should be noted that the origin of magnetoelectric coupling in our study is due to orbital ordering, which is similar to the previously reported multiferroicity in LuFe2O4 [70]. In LuFe2O4, the non-centrosymmetric distribution of Fe2+ and Fe3+ would give rise to spontaneous ferroelectricity and ferrimagnetism. To quantify the coupling between the magnetic order, electronic polarization, ionic displacement, and lattice distortion, a detailed study using a Landau-like model [71,72] is required in the future.

4. Conclusions

In conclusion, using first-principles calculations, we have shown that multiple nitride MXenes (V2NO2, V2NF2, Cr2NF2, Mo2NO2, Mo2NF2, Mn2NO2) are magnetic semiconductors, with band gaps ranging from 0.20 eV to 2.1 eV, and exhibit orbital ordering. Among them, Cr2NF2 exhibits spin spiral magnetization induced by orbital ordering, where the propagation vector is coupled with the ferroelastic strain of the structure, suggesting the magnetic order can be switched by applying external stress. We also find that the Janus MXene Mo8N4F7O exhibits two distinct ferroelectric phases marked by different orbital orderings, where the energy difference between the two phases can be reduced by applying an external strain or external electric field. Remarkably, the rearrangement of the orbital ordering during transition would also reverse the direction of net magnetism, illustrating the coupling between the strain tensor, electrical polarization, and magnetization. Our results show that orbital ordering can play an important role in coupling the electronic, magnetic, and mechanical properties of 2D nitride MXenes, making nitride MXenes important candidates for potential application in spintronic devices.

Author Contributions

Investigation, S.T.; Writing—original draft, S.T.; Writing—review & editing, S.T. and H.-T.J.; Supervision, H.-T.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data available on request from the authors.

Acknowledgments

This work was supported by the National Science and Technology Council (NSTC), Taiwan. H.-T.J. also acknowledges support from NCHC, CINC-NTU, AS-iMATE-111-12, and CQT-NTHU-MOE, Taiwan.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gao, Y.; Gao, M.; Lu, Y. Two-dimensional multiferroics. Nanoscale 2021, 13, 19324–19340. [Google Scholar] [CrossRef]
  2. Tian, S.; Zhang, J.F.; Li, C.; Ying, T.; Li, S.; Zhang, X.; Liu, K.; Lei, H. Ferromagnetic van der Waals Crystal VI3. J. Am. Chem. Soc. 2019, 141, 5326–5333. [Google Scholar] [CrossRef]
  3. Zhang, Z.; Shang, J.; Jiang, C.; Rasmita, A.; Gao, W.; Yu, T. Direct Photoluminescence Probing of Ferromagnetism in Monolayer Two-Dimensional CrBr3. Nano Lett. 2019, 19, 3138–3142. [Google Scholar] [CrossRef]
  4. Bonilla, M.; Kolekar, S.; Ma, Y.; Diaz, H.C.; Kalappattil, V.; Das, R.; Eggers, T.; Gutierrez, H.R.; Phan, M.H.; Batzill, M. Strong room-temperature ferromagnetism in VSe2 monolayers on van der Waals substrates. Nat. Nanotechnol. 2018, 13, 289–293. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, T.; Xu, X.; Dai, Y.; Huang, B.; Ma, Y. Intrinsic ferromagnetic triferroicity in bilayer T’-VTe2. Appl. Phys. Lett. 2022, 120, 192903. [Google Scholar] [CrossRef]
  6. Li, J.; Zhao, B.; Chen, P.; Wu, R.; Li, B.; Xia, Q.; Guo, G.; Luo, J.; Zang, K.; Zhang, Z.; et al. Synthesis of Ultrathin Metallic MTe2 (M = V, Nb, Ta) Single-Crystalline Nanoplates. Adv. Mater. 2018, 30, 1801043. [Google Scholar] [CrossRef] [PubMed]
  7. OHara, D.J.; Zhu, T.; Trout, A.H.; Ahmed, A.S.; Luo, Y.K.; Lee, C.H.; Brenner, M.R.; Rajan, S.; Gupta, J.A.; McComb, D.W.; et al. Room Temperature Intrinsic Ferromagnetism in Epitaxial Manganese Selenide Films in the Monolayer Limit. Nano Lett. 2018, 18, 3125–3131. [Google Scholar] [CrossRef]
  8. O’Neill, A.; Rahman, S.; Zhang, Z.; Schoenherr, P.; Yildirim, T.; Gu, B.; Su, G.; Lu, Y.; Seidel, J. Enhanced Room Temperature Ferromagnetism in Highly Strained 2D Semiconductor Cr2Ge2Te6. ACS Nano 2023, 17, 735–742. [Google Scholar] [CrossRef]
  9. Ostwal, V.; Shen, T.; Appenzeller, J. Efficient Spin-Orbit Torque Switching of the Semiconducting Van Der Waals Ferromagnet Cr2Ge2Te6. Adv. Mater. 2020, 32, 1906021. [Google Scholar] [CrossRef]
  10. Cai, X.; Song, T.; Wilson, N.P.; Clark, G.; He, M.; Zhang, X.; Taniguchi, T.; Watanabe, K.; Yao, W.; Xiao, D.; et al. Atomically Thin CrCl3: An In-Plane Layered Antiferromagnetic Insulator. Nano Lett. 2019, 19, 3993–3998. [Google Scholar] [CrossRef]
  11. Lee, Y.; Son, S.; Kim, C.; Kang, S.; Shen, J.; Kenzelmann, M.; Delley, B.; Savchenko, T.; Parchenko, S.; Na, W.; et al. Giant Magnetic Anisotropy in the Atomically Thin van der Waals Antiferromagnet FePS3. Adv. Electron. Mater. 2023, 9, 2200650. [Google Scholar] [CrossRef]
  12. Wang, X.; Du, K.; Liu, Y.Y.F.; Hu, P.; Zhang, J.; Zhang, Q.; Owen, M.H.S.; Lu, X.; Gan, C.K.; Sengupta, P.; et al. Raman spectroscopy of atomically thin two-dimensional magnetic iron phosphorus trisulfide (FePS3) crystals. 2D Mater. 2016, 3, 031009. [Google Scholar] [CrossRef]
  13. Suzuki, H.; Liu, H.; Bertinshaw, J.; Ueda, K.; Kim, H.; Laha, S.; Weber, D.; Yang, Z.; Wang, L.; Takahashi, H.; et al. Proximate ferromagnetic state in the Kitaev model material α-RuCl3. Nat. Commun. 2021, 12, 4512. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, X.X.; Li, L.; Weber, D.; Goldberger, J.; Mak, K.F.; Shan, J. Gate-tunable spin waves in antiferromagnetic atomic bilayers. Nat. Mater. 2020, 19, 838–842. [Google Scholar] [CrossRef] [PubMed]
  15. Huang, B.; Clark, G.; Klein, D.R.; MacNeill, D.; Navarro-Moratalla, E.; Seyler, K.L.; Wilson, N.; McGuire, M.A.; Cobden, D.H.; Xiao, D.; et al. Electrical control of 2D magnetism in bilayer CrI3. Nat. Nanotechnol. 2018, 13, 544–548. [Google Scholar] [CrossRef]
  16. Jiang, S.; Li, L.; Wang, Z.; Mak, K.F.; Shan, J. Controlling magnetism in 2D CrI3 by electrostatic doping. Nat. Nanotechnol. 2018, 13, 549–553. [Google Scholar] [CrossRef]
  17. Verzhbitskiy, I.A.; Kurebayashi, H.; Cheng, H.; Zhou, J.; Khan, S.; Feng, Y.P.; Eda, G. Controlling the magnetic anisotropy in Cr2Ge2Te6 by electrostatic gating. Nat. Electron. 2020, 3, 460–465. [Google Scholar] [CrossRef]
  18. Telford, E.J.; Dismukes, A.H.; Dudley, R.L.; Wiscons, R.A.; Lee, K.; Chica, D.G.; Ziebel, M.E.; Han, M.G.; Yu, J.; Shabani, S.; et al. Coupling between magnetic order and charge transport in a two-dimensional magnetic semiconductor. Nat. Mater. 2022, 21, 754–760. [Google Scholar] [CrossRef]
  19. Tahir, R.; Fatima, S.; Zahra, S.A.; Akinwande, D.; Li, H.; Jafri, S.H.M.; Rizwan, S. Multiferroic and ferroelectric phases revealed in 2D Ti3C2Tx MXene film for high performance resistive data storage devices. Npj 2D Mater. Appl. 2023, 7, 7. [Google Scholar] [CrossRef]
  20. Zhao, M.; Chen, J.; Wang, S.S.; An, M.; Dong, S. Multiferroic properties of oxygen-functionalized magnetic i-MXene. Phys. Rev. Mater. 2021, 5, 094408. [Google Scholar] [CrossRef]
  21. Zhang, J.J.; Lin, L.; Zhang, Y.; Wu, M.; Yakobson, B.I.; Dong, S. Type-II Multiferroic Hf2VC2F2 MXene Monolayer with High Transition Temperature. J. Am. Chem. Soc. 2018, 140, 9768–9773. [Google Scholar] [CrossRef] [PubMed]
  22. Frey, N.C.; Bandyopadhyay, A.; Kumar, H.; Anasori, B.; Gogotsi, Y.; Shenoy, V.B. Surface-Engineered MXenes: Electric Field Control of Magnetism and Enhanced Magnetic Anisotropy. ACS Nano 2019, 13, 2831–2839. [Google Scholar] [CrossRef] [PubMed]
  23. Li, M.; Lu, J.; Luo, K.; Li, Y.; Chang, K.; Chen, K.; Zhou, J.; Rosen, J.; Hultman, L.; Eklund, P.; et al. Element Replacement Approach by Reaction with Lewis Acidic Molten Salts to Synthesize Nanolaminated MAX Phases and MXenes. J. Am. Chem. Soc. 2019, 141, 4730–4737. [Google Scholar] [CrossRef] [PubMed]
  24. Jawaid, A.; Hassan, A.; Neher, G.; Nepal, D.; Pachter, R.; Kennedy, W.J.; Ramakrishnan, S.; Vaia, R.A. Halogen Etch of Ti3AlC2 MAX Phase for MXene Fabrication. ACS Nano 2021, 15, 2771–2777. [Google Scholar] [CrossRef]
  25. Sun, Z.; Yuan, M.; Lin, L.; Yang, H.; Nan, C.; Li, H.; Sun, G.; Yang, X. Selective Lithiation–Expansion–Microexplosion Synthesis of Two-Dimensional Fluoride-Free Mxene. ACS Mater. Lett. 2019, 1, 628–632. [Google Scholar] [CrossRef]
  26. Huang, X.; Wu, P. A Facile, High-Yield, and Freeze-and-Thaw-Assisted Approach to Fabricate MXene with Plentiful Wrinkles and Its Application in On-Chip Micro-Supercapacitors. Adv. Funct. Mater. 2020, 30, 1910048. [Google Scholar] [CrossRef]
  27. Urbankowski, P.; Anasori, B.; Hantanasirisakul, K.; Yang, L.; Zhang, L.; Haines, B.; May, S.J.; Billinge, S.J.L.; Gogotsi, Y. 2D molybdenum and vanadium nitrides synthesized by ammoniation of 2D transition metal carbides (MXenes). Nanoscale 2017, 9, 17722–17730. [Google Scholar] [CrossRef]
  28. Soundiraraju, B.; George, B.K. Two-Dimensional Titanium Nitride (Ti2N) MXene: Synthesis, Characterization, and Potential Application as Surface-Enhanced Raman Scattering Substrate. ACS Nano 2017, 11, 8892–8900. [Google Scholar] [CrossRef]
  29. Urbankowski, P.; Anasori, B.; Makaryan, T.; Er, D.; Kota, S.; Walsh, P.L.; Zhao, M.; Shenoy, V.B.; Barsoum, M.W.; Gogotsi, Y. Synthesis of two-dimensional titanium nitride Ti4N3 (MXene). Nanoscale 2016, 8, 11385–11391. [Google Scholar] [CrossRef]
  30. Kumar, H.; Frey, N.C.; Dong, L.; Anasori, B.; Gogotsi, Y.; Shenoy, V.B. Tunable Magnetism and Transport Properties in Nitride MXenes. ACS Nano 2017, 11, 7648–7655. [Google Scholar] [CrossRef]
  31. Liechtenstein, A.I.; Anisimov, V.I.; Zaanen, J. Density-functional theory and strong interactions: Orbital ordering in Mott-Hubbard insulators. Phys. Rev. B 1995, 52, R5467–R5470. [Google Scholar] [CrossRef] [PubMed]
  32. da Silva Neto, E.H.; Aynajian, P.; Frano, A.; Comin, R.; Schierle, E.; Weschke, E.; Gyenis, A.; Wen, J.; Schneeloch, J.; Xu, Z.; et al. Ubiquitous Interplay between Charge Ordering and High-Temperature Superconductivity in Cuprates. Science 2014, 343, 393–396. [Google Scholar] [CrossRef]
  33. Bednorz, J.G.; Müller, K.A.; Takashige, M. Superconductivity in Alkaline Earth-Substituted La2CuO4-y. Science 1987, 236, 73–75. [Google Scholar] [CrossRef] [PubMed]
  34. Bednorz, J.G.; Müller, K.A. Perovskite-type oxides—The new approach to high-Tc superconductivity. Rev. Mod. Phys. 1988, 60, 585–600. [Google Scholar] [CrossRef]
  35. Labbé, J.; Bok, J. Superconductivity in Alcaline-Earth-Substituted La2CuO4: A Theoretical Model. Europhys. Lett. 1987, 3, 1225. [Google Scholar] [CrossRef]
  36. Chahara, K.; Ohno, T.; Kasai, M.; Kozono, Y. Magnetoresistance in magnetic manganese oxide with intrinsic antiferromagnetic spin structure. Appl. Phys. Lett. 1993, 63, 1990–1992. [Google Scholar] [CrossRef]
  37. von Helmolt, R.; Wecker, J.; Holzapfel, B.; Schultz, L.; Samwer, K. Giant negative magnetoresistance in perovskitelike La2/3Ba1/3MnOx ferromagnetic films. Phys. Rev. Lett. 1993, 71, 2331–2333. [Google Scholar] [CrossRef]
  38. Raveau, B.; Hervieu, M.; Maignan, A.; Martin, C. The route to CMR manganites: What about charge ordering and phase separation? J. Mater. Chem. 2001, 11, 29–36. [Google Scholar] [CrossRef]
  39. Varignon, J.; Bristowe, N.C.; Ghosez, P. Electric Field Control of Jahn-Teller Distortions in Bulk Perovskites. Phys. Rev. Lett. 2016, 116, 057602. [Google Scholar] [CrossRef]
  40. Goodenough, J.B. Jahn-teller phenomena in solids. Annu. Rev. Mater. Sci. 1998, 28, 1–27. [Google Scholar] [CrossRef]
  41. Chen, L.; Xu, C.; Tian, H.; Xiang, H.; Íñiguez, J.; Yang, Y.; Bellaiche, L. Electric-Field Control of Magnetization, Jahn-Teller Distortion, and Orbital Ordering in Ferroelectric Ferromagnets. Phys. Rev. Lett. 2019, 122, 247701. [Google Scholar] [CrossRef] [PubMed]
  42. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef] [PubMed]
  43. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  44. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  45. 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, 1505–1509. [Google Scholar] [CrossRef]
  46. Sun, W.; Xie, Y.; Kent, P.R.C. Double transition metal MXenes with wide band gaps and novel magnetic properties. Nanoscale 2018, 10, 11962–11968. [Google Scholar] [CrossRef]
  47. Li, Y.; Deng, J.; Zhang, Y.F.; Jin, X.; Dong, W.H.; Sun, J.T.; Pan, J.; Du, S. Nonvolatile electrical control of spin polarization in the 2D bipolar magnetic semiconductor VSeF. Npj Comput. Mater. 2023, 9, 50. [Google Scholar] [CrossRef]
  48. Zhang, Y.; Cui, Z.; Sa, B.; Miao, N.; Zhou, J.; Sun, Z. Computational design of double transition metal MXenes with intrinsic magnetic properties. Nanoscale Horiz. 2022, 7, 276–287. [Google Scholar] [CrossRef]
  49. Reshak, A.H. Microcrystalline β-RbNd(MoO4)2: Spin polarizing DFT+U. RSC Adv. 2015, 5, 44960–44968. [Google Scholar] [CrossRef]
  50. Aykol, M.; Wolverton, C. Local environment dependent GGA+U method for accurate thermochemistry of transition metal compounds. Phys. Rev. B 2014, 90, 115105. [Google Scholar] [CrossRef]
  51. Franchini, C.; Podloucky, R.; Paier, J.; Marsman, M.; Kresse, G. Ground-state properties of multivalent manganese oxides: Density functional and hybrid density functional calculations. Phys. Rev. B 2007, 75, 195128. [Google Scholar] [CrossRef]
  52. Sheppard, D.; Xiao, P.; Chemelewski, W.; Johnson, D.D.; Henkelman, G. A generalized solid-state nudged elastic band method. J. Chem. Phys. 2012, 136, 074103. [Google Scholar] [CrossRef]
  53. King-Smith, R.D.; Vanderbilt, D. Theory of polarization of crystalline solids. Phys. Rev. B 1993, 47, 1651–1654. [Google Scholar] [CrossRef] [PubMed]
  54. Stokes, H.T.; Hatch, D.M. Introduction to Isotropy Subgroups and Displacive Phase Transitions; Brigham Young University: Provo, Utah, 2006. [Google Scholar]
  55. Pizzi, G.; Vitale, V.; Arita, R.; Blügel, S.; Freimuth, F.; Géranton, G.; Gibertini, M.; Gresch, D.; Johnson, C.; Koretsune, T.; et al. Wannier90 as a community code: New features and applications. J. Phys. Condens. Matter 2020, 32, 165902. [Google Scholar] [CrossRef]
  56. Mostofi, A.A.; Yates, J.R.; Lee, Y.S.; Souza, I.; Vanderbilt, D.; Marzari, N. wannier90: A tool for obtaining maximally-localised Wannier functions. Comput. Phys. Commun. 2008, 178, 685–699. [Google Scholar] [CrossRef]
  57. Liechtenstein, A.; Katsnelson, M.; Antropov, V.; Gubanov, V. Local spin density functional approach to the theory of exchange interactions in ferromagnetic metals and alloys. J. Magn. Magn. Mater. 1987, 67, 65–74. [Google Scholar] [CrossRef]
  58. He, X.; Helbig, N.; Verstraete, M.J.; Bousquet, E. TB2J: A python package for computing magnetic interaction parameters. Comput. Phys. Commun. 2021, 264, 107938. [Google Scholar] [CrossRef]
  59. Müller, G.P.; Hoffmann, M.; Dißelkamp, C.; Schürhoff, D.; Mavros, S.; Sallermann, M.; Kiselev, N.S.; Jónsson, H.; Blügel, S. Spirit: Multifunctional framework for atomistic spin simulations. Phys. Rev. B 2019, 99, 224414. [Google Scholar] [CrossRef]
  60. Hatch, D.M.; Stokes, H.T. Complete listing of order parameters for a crystalline phase transition: A solution to the generalized inverse Landau problem. Phys. Rev. B 2001, 65, 014113. [Google Scholar] [CrossRef]
  61. Iqbal, M.; Fatheema, J.; Noor, Q.; Rani, M.; Mumtaz, M.; Zheng, R.K.; Khan, S.A.; Rizwan, S. Co-existence of magnetic phases in two-dimensional MXene. Mater. Today Chem. 2020, 16, 100271. [Google Scholar] [CrossRef]
  62. Zinke, R.; Richter, J.; Drechsler, S.L. Spiral correlations in frustrated one-dimensional spin-1/2 Heisenberg J1–J2–J3 ferromagnets. J. Phys. Condens. Matter 2010, 22, 446002. [Google Scholar] [CrossRef]
  63. Wang, H.; Li, X.; Sun, J.; Liu, Z.; Yang, J. BP5 monolayer with multiferroicity and negative Poisson’s ratio: A prediction by global optimization method. 2D Mater. 2017, 4, 045020. [Google Scholar] [CrossRef]
  64. Kou, L.; Ma, Y.; Tang, C.; Sun, Z.; Du, A.; Chen, C. Auxetic and Ferroelastic Borophane: A Novel 2D Material with Negative Possion’s Ratio and Switchable Dirac Transport Channels. Nano Lett. 2016, 16, 7910–7914. [Google Scholar] [CrossRef] [PubMed]
  65. Wu, M.; Zeng, X.C. Intrinsic Ferroelasticity and/or Multiferroicity in Two-Dimensional Phosphorene and Phosphorene Analogues. Nano Lett. 2016, 16, 3236–3241. [Google Scholar] [CrossRef]
  66. Xu, X.; Ma, Y.; Zhang, T.; Lei, C.; Huang, B.; Dai, Y. Prediction of two-dimensional antiferromagnetic ferroelasticity in an AgF2 monolayer. Nanoscale Horiz. 2020, 5, 1386–1393. [Google Scholar] [CrossRef] [PubMed]
  67. Xu, B.; Xiang, H.; Yin, J.; Xia, Y.; Liu, Z. A two-dimensional tetragonal yttrium nitride monolayer: A ferroelastic semiconductor with switchable anisotropic properties. Nanoscale 2018, 10, 215–221. [Google Scholar] [CrossRef]
  68. Li, X.; Wu, X.; Li, Z.; Yang, J.; Hou, J.G. Bipolar magnetic semiconductors: A new class of spintronics materials. Nanoscale 2012, 4, 5680–5685. [Google Scholar] [CrossRef]
  69. Li, J.; Li, X.; Yang, J. A review of bipolar magnetic semiconductors from theoretical aspects. Fundam. Res. 2022, 2, 511–521. [Google Scholar] [CrossRef]
  70. Liu, F.; Hao, Y.; Ni, J.; Zhao, Y.; Zhang, D.; Fabbris, G.; Haskel, D.; Cheng, S.; Xu, X.; Yin, L.; et al. Pressure-induced charge orders and their postulated coupling to magnetism in hexagonal multiferroic LuFe2O4. npj Quantum Mater. 2023, 8, 1. [Google Scholar] [CrossRef]
  71. Lejman, M.; Paillard, C.; Juvé, V.; Vaudel, G.; Guiblin, N.; Bellaiche, L.; Viret, M.; Gusev, V.E.; Dkhil, B.; Ruello, P. Magnetoelastic and magnetoelectric couplings across the antiferromagnetic transition in multiferroic BiFeO3. Phys. Rev. B 2019, 99, 104103. [Google Scholar] [CrossRef]
  72. Lu, X.Z.; Wu, X.; Xiang, H.J. General microscopic model of magnetoelastic coupling from first principles. Phys. Rev. B 2015, 91, 100405. [Google Scholar] [CrossRef]
Figure 1. The spin densities of AFM patterns. Four AFM phases: (a) AFM1, (b) AFM2, (c) AFM3, and (d) AFM4 are considered in our calculations. Red and blue colors indicate up- and down-spin densities; the cyan, gray, and red spheres represent transition metals (TM), nitrogen, and terminal groups (oxygen or fluorine), respectively.
Figure 1. The spin densities of AFM patterns. Four AFM phases: (a) AFM1, (b) AFM2, (c) AFM3, and (d) AFM4 are considered in our calculations. Red and blue colors indicate up- and down-spin densities; the cyan, gray, and red spheres represent transition metals (TM), nitrogen, and terminal groups (oxygen or fluorine), respectively.
Nanomaterials 13 02644 g001
Figure 2. Lattice structure of space groups: (a) P 3 ¯ m 1 , (b) P 2 1 / m , and (c) P 12 1 / m 1 . The cyan and dark blue spheres in (b,c) represent transition metals (TM) in sublattice 1 and sublattice 2, respectively. The distortion of the parent structure ( P 3 ¯ m 1 ) can be decomposed into displacements with irreducible representations (d) Γ 1 + ( a , 0 , 0 ) , (e) Γ 3 + ( a , 3 a , 0 ) , (f) Γ 3 + ( 0 , a , 0 ) , (g) M 2 ( 0 , a , 0 ) , and (h) M 1 ( 0 , a , 0 ) . The algorithm used to obtain these irreducible representations is summarized in Dorian and Harolds’ work [60].
Figure 2. Lattice structure of space groups: (a) P 3 ¯ m 1 , (b) P 2 1 / m , and (c) P 12 1 / m 1 . The cyan and dark blue spheres in (b,c) represent transition metals (TM) in sublattice 1 and sublattice 2, respectively. The distortion of the parent structure ( P 3 ¯ m 1 ) can be decomposed into displacements with irreducible representations (d) Γ 1 + ( a , 0 , 0 ) , (e) Γ 3 + ( a , 3 a , 0 ) , (f) Γ 3 + ( 0 , a , 0 ) , (g) M 2 ( 0 , a , 0 ) , and (h) M 1 ( 0 , a , 0 ) . The algorithm used to obtain these irreducible representations is summarized in Dorian and Harolds’ work [60].
Nanomaterials 13 02644 g002
Figure 3. Contour plot of spin density in sublattices of Cr2NF2 with (a) lower ionization, sublattice 1, and (b) higher ionization, sublattice 2. The cyan, gray, and red spheres represent Cr, N, and F ions, respectively; the solid black arrows specify the axes of the local coordinate system for d orbitals. Projected band structure of Cr2NF2 on d x y , d y z , d z 2 , d z x , and d x 2 y 2 orbitals in (c) sublattice 1 and (d) sublattice 2.
Figure 3. Contour plot of spin density in sublattices of Cr2NF2 with (a) lower ionization, sublattice 1, and (b) higher ionization, sublattice 2. The cyan, gray, and red spheres represent Cr, N, and F ions, respectively; the solid black arrows specify the axes of the local coordinate system for d orbitals. Projected band structure of Cr2NF2 on d x y , d y z , d z 2 , d z x , and d x 2 y 2 orbitals in (c) sublattice 1 and (d) sublattice 2.
Nanomaterials 13 02644 g003
Figure 4. Projected band structures of (a) V2NO2, (b) V2NF2, (c) Cr2NF2, (d) Mo2NO2, (e) Mo2NF2, and (f) Mn2NO2 on the t 2 g and e g orbitals of metals M, and also on the p orbitals of the ligands (N, O, or F).
Figure 4. Projected band structures of (a) V2NO2, (b) V2NF2, (c) Cr2NF2, (d) Mo2NO2, (e) Mo2NF2, and (f) Mn2NO2 on the t 2 g and e g orbitals of metals M, and also on the p orbitals of the ligands (N, O, or F).
Nanomaterials 13 02644 g004
Figure 5. Spin textures of (a) Mn2NO2 at 0 K and (b) Cr2NF2 at 0 K generated using a Monte Carlo simulation in a 50 × 100 lattice with periodic boundaries. The color map is the magnitude of spin S i polarized along the out-of-plane axis; the green (black) arrows in sublattice 1 (2) represent the in-plane components of spin S i ; the cyan and blue spheres correspond to metal ions in sublattices 1 and 2, respectively. Spin components of (c) Mn2NO2 projected to a–b plane and (d) Cr2NF2 projected to a–c plane; ϕ i , j represents the initial phase of the spin components in sublattice i in layer j.
Figure 5. Spin textures of (a) Mn2NO2 at 0 K and (b) Cr2NF2 at 0 K generated using a Monte Carlo simulation in a 50 × 100 lattice with periodic boundaries. The color map is the magnitude of spin S i polarized along the out-of-plane axis; the green (black) arrows in sublattice 1 (2) represent the in-plane components of spin S i ; the cyan and blue spheres correspond to metal ions in sublattices 1 and 2, respectively. Spin components of (c) Mn2NO2 projected to a–b plane and (d) Cr2NF2 projected to a–c plane; ϕ i , j represents the initial phase of the spin components in sublattice i in layer j.
Nanomaterials 13 02644 g005
Figure 6. (ad): Top view of Cr2NF2, where light blue, dark blue, gray, and red spheres correspond to chromium in the lower layer, chromium in the upper layer, nitrogen, and oxygen ions, respectively. The atomic bonds are represented in blue colored dashed lines, while green colored cylinders mark the Cr ion pairs with isotropic coupling: (a) J1 (dark green) and J2 (light green), (b) J3, (c) J4, and (d) J5.
Figure 6. (ad): Top view of Cr2NF2, where light blue, dark blue, gray, and red spheres correspond to chromium in the lower layer, chromium in the upper layer, nitrogen, and oxygen ions, respectively. The atomic bonds are represented in blue colored dashed lines, while green colored cylinders mark the Cr ion pairs with isotropic coupling: (a) J1 (dark green) and J2 (light green), (b) J3, (c) J4, and (d) J5.
Nanomaterials 13 02644 g006
Figure 7. Ferroelastic transition of Cr2NF2. Top view of lower layer at (a) ground state with a 2 = a 3 > a 1 , (b) ground state with a 1 = a 3 > a 2 , and (c) ground state with a 1 = a 2 > a 3 .
Figure 7. Ferroelastic transition of Cr2NF2. Top view of lower layer at (a) ground state with a 2 = a 3 > a 1 , (b) ground state with a 1 = a 3 > a 2 , and (c) ground state with a 1 = a 2 > a 3 .
Nanomaterials 13 02644 g007
Figure 8. Contour plot of spin density of (a) antiferroelctric phase (ground state), (b) ferroelectric phase of V2NO2; where the yellow and blue colored isosurfaces represent spin-up and spin-down density, respectively. The V ion in the lower (upper) layer circled in black (red) is referred to as VL (VU).
Figure 8. Contour plot of spin density of (a) antiferroelctric phase (ground state), (b) ferroelectric phase of V2NO2; where the yellow and blue colored isosurfaces represent spin-up and spin-down density, respectively. The V ion in the lower (upper) layer circled in black (red) is referred to as VL (VU).
Nanomaterials 13 02644 g008
Figure 9. The density of states of Mo8N4F7O in (a) P1 phase and (b) N1 phase. (c) Contour plot of spin density of Mo8N4F7O in P1 phase with ferroelastic strain of 10.2%, where the blue and red colored isosurfaces represent spin-up and spin-down density, respectively. The Mo ion circled in black (red) is referred as MoL (MoU), where (d,e) illustrates the change in orbital ordering of MoU and MoL during the transition from the P1 to the N1 phase, respectively. (f) Energy of Mo8N4F7O in the P1 and N1 phases as a function of applied electric field in the out-of-plane direction at the ferroelastic strain of 10.2%.
Figure 9. The density of states of Mo8N4F7O in (a) P1 phase and (b) N1 phase. (c) Contour plot of spin density of Mo8N4F7O in P1 phase with ferroelastic strain of 10.2%, where the blue and red colored isosurfaces represent spin-up and spin-down density, respectively. The Mo ion circled in black (red) is referred as MoL (MoU), where (d,e) illustrates the change in orbital ordering of MoU and MoL during the transition from the P1 to the N1 phase, respectively. (f) Energy of Mo8N4F7O in the P1 and N1 phases as a function of applied electric field in the out-of-plane direction at the ferroelastic strain of 10.2%.
Nanomaterials 13 02644 g009
Table 1. Magnetic moments in sublattice 1 (2) are shown without (with) parentheses. Using a Monte Carlo (MC) simulation, we predicted the magnetic order, Néel temperatures of the AFM lattice structures and transition temperatures of non-collinear magnetic structures. The energy of 2 × 2 supercells with different magnetic configurations (Figure 1a–d) are calculated using DFT and listed with reference to the energy of their respective magnetic ground state.
Table 1. Magnetic moments in sublattice 1 (2) are shown without (with) parentheses. Using a Monte Carlo (MC) simulation, we predicted the magnetic order, Néel temperatures of the AFM lattice structures and transition temperatures of non-collinear magnetic structures. The energy of 2 × 2 supercells with different magnetic configurations (Figure 1a–d) are calculated using DFT and listed with reference to the energy of their respective magnetic ground state.
MXeneBand Gap/eVMagnetic Moment/µBMagnetic Order (MC)Transition Temperature/KE (AFM1)/meVE (AFM2)/meVE (AFM3)/meVE (AFM4)/meVE (FM)/meV
V2NO20.781.9 (1.2)AFM31401061310-44
V2NF20.852.7 (2.1)AFM3380871808141304
Cr2NF22.13.9 (3.1)Néel spin spiral1203943990340549
Mo2NO20.952.4 (1.9)AFM335013201586012982796
Mo2NF21.753.9 (3.4)AFM3630119518630-1266
Mn2NO20.203.9 (3.4)Conical spin spiral110440334336590
“-” indicates DFT calculation failed to converge.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Teh, S.; Jeng, H.-T. Magnetoelastic and Magnetoelectric Coupling in Two-Dimensional Nitride MXenes: A Density Functional Theory Study. Nanomaterials 2023, 13, 2644. https://doi.org/10.3390/nano13192644

AMA Style

Teh S, Jeng H-T. Magnetoelastic and Magnetoelectric Coupling in Two-Dimensional Nitride MXenes: A Density Functional Theory Study. Nanomaterials. 2023; 13(19):2644. https://doi.org/10.3390/nano13192644

Chicago/Turabian Style

Teh, Sukhito, and Horng-Tay Jeng. 2023. "Magnetoelastic and Magnetoelectric Coupling in Two-Dimensional Nitride MXenes: A Density Functional Theory Study" Nanomaterials 13, no. 19: 2644. https://doi.org/10.3390/nano13192644

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