Next Article in Journal
The Mechanism of Manipulating Chirality and Chiral Sensing Based on Chiral Plexcitons in a Strong-Coupling Regime
Previous Article in Journal
The Influence Mechanism of Quantum Well Growth and Annealing Temperature on In Migration and Stress Modulation Behavior
Previous Article in Special Issue
MoS2-Based Memristor: Robust Resistive Switching Behavior and Reliable Biological Synapse Emulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Strain-Induced Ferromagnetism in Monolayer T″-Phase VTe2: Unveiling Magnetic States and Anisotropy for Spintronics Advancement

1
Department of Physics, College of Science, Shanghai University, 99 Shangda Road, Shanghai 200444, China
2
Department of Physics, National University of Singapore, Singapore 117542, Singapore
3
Institute of Materials Research and Engineering (IMRE), Agency for Science, Technology and Research (A*STAR), 2 Fusionopolis Way, Innovis #08-03, Singapore 138634, Singapore
4
Materials Genome Institute (MGI), Shanghai University, 333 Nanchen Road, Shanghai 200444, China
5
EACOMP, Shenzhen 518055, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2024, 14(8), 704; https://doi.org/10.3390/nano14080704
Submission received: 18 March 2024 / Revised: 14 April 2024 / Accepted: 15 April 2024 / Published: 18 April 2024

Abstract

:
Two-dimensional (2D) ferromagnets have attracted significant interest for their potential in spintronic device miniaturization, especially since the discovery of ferromagnetic ordering in monolayer materials such as CrI3 and Fe3GeTe2 in 2017. This study presents a detailed investigation into the effects of the Hubbard U parameter, biaxial strain, and structural distortions on the magnetic characteristics of T″-phase VTe2. We demonstrate that setting the Hubbard U to 0 eV provides an accurate representation of the observed structural, magnetic, and electronic features for both bulk and monolayer T″-phase VTe2. The application of strain reveals two distinct ferromagnetic states in the monolayer T″-phase VTe2, each characterized by minor structural differences, but notably different magnetic moments. The T″-1 state, with reduced magnetic moments, emerges under compressive strain, while the T″-2 state, featuring increased magnetic moments, develops under tensile strain. Our analysis also compares the magnetic anisotropy between the T and T″ phases of VTe2, highlighting that the periodic lattice distortion in the T″-phase induces an in-plane anisotropy, which makes it a material with an easy-axis of magnetization. Monte Carlo simulations corroborate our findings, indicating a high Curie temperature of approximately 191 K for the T″-phase VTe2. Our research not only sheds light on the critical aspects of the VTe2 system but also suggests new pathways for enhancing low-dimensional magnetism, contributing to the advancement of spintronics and straintronics.

1. Introduction

The experimental confirmation of two-dimensional (2D) long-range ferromagnetic (FM) ordering in materials like CrI3 and CrGeTe3 has sparked considerable research interest in the realm of low-dimensional magnetism, serving both fundamental physics exploration and practical applications [1,2,3,4,5,6]. Subsequently, numerous other 2D ferromagnets have been reported using different methodologies, including high-throughput screening and machine learning [7,8,9,10,11,12,13]. The unexpected revelation of such long-range 2D magnetic ordering caused a stir in the materials community, as the Mermin–Wagner theorem had previously deemed any low-dimensional long-range ordering impossible at finite temperatures [14]. Further investigations have elucidated that magnetic anisotropy can indeed stabilize long-range magnetism in 2D materials [15]. Alongside the challenging quest for the rare intrinsic magnetic easy axis in 2D magnets, material engineering techniques, such as inducing magnetic anisotropy through uniaxial strain in initially easy-plane magnetic systems, have been successfully employed [16,17,18].
Transition metal dichalcogenides (TMDCs) represent a rich repository of 2D ferromagnets. The magnetic moments observed in TMDCs stem from the partial filling of d orbitals in transition metal ions, facilitating the possibility of long-range magnetic order under specific conditions. Moreover, the presence of heavy elements such as Se and Te impart considerable spin–orbit coupling (SOC) to TMDCs. Notably, tellurides are particularly appealing due to the lower electronegativity and enhanced SOC effects of Te atoms [19]. For instance, group V metal tellurides (VTe2, NbTe2, TaTe2) showcase diverse structural variations and hold promise as a platform for investigating various fundamental physics phenomena, including superconductivity, electro-catalytic activity, excitons, and the quantum spin Hall effect [20]. Experimental studies by Mitsuishi et al. have confirmed the existence of topological surface states in both T and T″ bulk VTe2 [21,22]. Furthermore, the prospect of ferromagnetic TMDCs opens up avenues for further exploration in fundamental physics and the development of novel magneto-electronic/optoelectronic devices [23].
An intriguing physical phenomenon within TMDCs is the charge density wave (CDW), which draws considerable attention due to its wide existence, rich patterns, and capability to be tunable by external influences [24,25,26,27,28,29]. CDWs are periodic modulations of charge density by lattice distortions in real space, alongside the opening of a band gap, below a critical temperature [30]. The CDW patterns in TMDCs have been proposed to depend critically on the d-electron count of the transition metal ions. For example, the most common CDW configurations include the 13 × 13 , 3 × 1 , 2 × 1 , and 2 × 2 rearrangements of metal ions for TMDCs with d1, d1+x (0 < x < 1), d2 and d3 TM ions, respectively [24]. The more related CDW phases for VTe2 are the T′ and T″ phases. As shown in Scheme 1, the T′ phase is characterized by a 2 × 1 reconstruction of the unit cell in the T phase, forming single zigzag chains, and the T″ phase is a 3 × 1 reconstruction with the two V atoms at the two sides moving head-to-head to the middle V atom, forming ribbon or double zigzag chains. The presence of CDWs can significantly alter a material’s properties, such as its mutual suppression with magnetism [31,32,33,34,35,36]. However, recent findings also suggest that CDW transitions can promote the formation of magnetism [37]. Thus, the interplay between CDWs and magnetism remains a subject of ongoing study, with the precise nature of their interaction still being elusive.
Despite this, VTe2 stands out for its diverse phase variants and magnetic properties. In its bulk form, VTe2 assumes the high-symmetric T phase at high temperatures, but undergoes a 3 × 1 charge density wave (CDW) reconstruction, which is named T″ phase in the later sections, at a critical temperature of approximately T C D W   480 K. Within the temperature range of 30–480 K, the T″-like CDW phases of bulk VTe2 are reported to exhibit two types of antiferromagnetic (AFM) behavior [38]. As bulk VTe2 is gradually thinned down to thin film or even monolayer, some fascinating phase evolutions appear. For instance, starting from the T″ in bulk, the phase transforms into the T phase in multilayer thin films, then to the T′ in bilayers, and ultimately to the T phase in monolayers [23,39,40]. Both the T′ and T″ phases are close variants of the T phase. Presently, no long-range magnetic order has been observed in monolayer T-phase VTe2 [41,42], although ferromagnetic behavior has been detected in multilayer T-phase VTe2 [23]. Considering the complex coupling among the charge, lattice, and spin order of VTe2, a close examination of its properties is imperative.
In this work, we conducted a thorough investigation into the multifaceted structural, electronic, and magnetic characteristics of VTe2, considering the impacts of both simulation parameters and structural modifications. Notably, our findings reveal that the periodic distortion present in VTe2 (T″ phase) can induce magnetic anisotropy and effectively stabilize two-dimensional magnetism, even up to a high Curie temperature. These results not only deepen our understanding of VTe2 but also present a promising pathway for enhancing low-dimensional magnetism. By elucidating the intricate interplay between structural distortions and magnetic properties, our study opens up new avenues for advancing the field of low-dimensional magnetism, potentially catalyzing the development of novel magnetic materials for future technological applications.

2. Materials and Methods

The first-principles calculations were performed based on the projector augmented-wave method as implemented in the Vienna ab initio simulation package (VASP 5.4.4) [43,44,45]. The exchange–correlation interaction was treated with the Perdew–Burke–Ernzerhof form of the generalized gradient approximation (GGA) [46]. The plane-wave cutoff energy was set to 500 eV. The lattice parameters and atomic positions were fully relaxed until the energy and force on each atom converged to less than 1 × 10−5 eV and 1 × 10−3 eV/Å, respectively. For magnetic anisotropy energy, the spin–orbital coupling was included with a higher cutoff criterion of 1 × 10−8 eV for the energy convergence. The K-mesh density was determined following the rule that ai × ki ≈ 50, where ai (unit Å) is the lattice parameter of lattice vector i (i = 1, 2, 3), and ki is the number of K-points corresponding to lattice vector i.
The VAMPIRE 5.0 package was used for metropolis Monte Carlo (MC) simulation [47,48], which was based on the following Heisenberg spin Hamiltonian:
H ~ = i j J i j S i · S j A i S i z   2
where S i and S j are the classical Heisenberg spins with unit magnitude at the site i and j. J is the nearest neighboring spin–spin coupling, and A is the magnetic anisotropic energy. The simulated spin systems for VTe2 consisted of 9600 spins, which were large enough to derive the converged magnetization–temperature curve. At each temperature, the orientations of the spins were initially randomized, and all the spins were first thermalized for 10,000 equilibrium steps and then by 50,000 averaging steps to derive the thermal equilibrium magnetization.

3. Results

3.1. The Determination of the Hubbard U Value for VTe2

The determination of the Hubbard U value holds paramount importance in ensuring the accuracy of Density Functional Theory (DFT) simulations, especially within vanadium-included compounds characterized by moderate correlation, which poses challenges to the determination process. Thus, it is imperative to exercise caution when evaluating its impacts on both crystalline and electronic structures. This section aims to elucidate our selection of the Hubbard U value, considering its implications on both crystalline and electronic structures.
Our approach begins with a comparative analysis of the relative stability and magnetic configurations of bulk VTe2 (T versus T″) across various U values. As shown in Figure 1, our calculations reveal that bulk T″-phase VTe2 transits into the T phase upon full relaxation of the structures when the Hubbard U exceeds a critical value Ucritical of ~2 eV. This transition is marked by a sharp change in atomic configuration, total magnetic moment, and volume. Conversely, the T″ phase is energetically favorable over T only at U = 0 eV (see Figure 1b). Considering the fact that VTe2 maintains its T″ phase at temperatures ≤ 480 K, we expect U = 0 eV to be a more reasonable value. In addition, the determined lattice parameters at U = 0 eV could well reproduce the crystal structure of bulk T″-phase VTe2. Specifically, the determined lattice parameters, a = 18.814 Å, b = 3.608 Å, c = 9.159 Å, and β = 134.5667° of the bulk T″-phase VTe2, are in favorable agreement with the experimentally determined parameters of a = 18.984 Å, b = 3.5947 Å, c = 9.069 Å, and β = 134.62° [49]. It is noted that the conventional cell is employed for direct comparison with the experimentally reported lattice parameters, whereas the primitive cell is utilized in the DFT simulations. Detailed information regarding the structure of the primitive cell of the bulk T″-phase VTe2, relaxed at U = 0 eV, is provided in Table 1.
The above results primarily delve into the impact of U on structural properties. In this section, we shift our focus to the influence of Hubbard U on magnetic properties. Due to the absence of the bulk T″ phase for U > 2 eV, all structures in this section are fixed to that of U = 0 eV. As shown in Figure 1c, the average magnetic moments (per VTe2 unit) of the bulk T″-phase VTe2 exhibits a striking change from nearly zero to >1 μ B as the Hubbard U parameter increases from 0 to 4 eV. It noted that in Figure 1a, a significant change in magnetic moment observed at U > 2 eV is attributed to the phase transition from bulk T″ to the T phase. When U exceeds 2 eV, the T″ phase becomes highly unstable, transitioning into the T phase upon full structural relaxation. This transition leads to a significant increase in magnetic moment, as the T phase inherently possesses a larger magnetic moment compared to the T″ phase. Conversely, the magnetic moment variation in Figure 1c is solely influenced by the enhanced localization of electrons due to the increased Hubbard U value, as the structures remain fixed. In addition, our results reveal that the bulk T″-phase VTe2 manifests ferrimagnetic behavior only at U = 0 eV and it becomes ferromagnetic for all the other U values studied in this work (Figure 1d). As an example, the FM coupling at U = 2 eV is shown in Figure 1f. The respective magnetic moment for V1 and V2 are 0.53 μ B and −0.16 μ B for the ferrimagnetic coupling (Figure 1e). The results at U = 0 eV are quite close to the experimental observed antiferromagnetic (AFM) state [50]. The ferrimagnetic order obtained from our calculation is also qualitatively consistent with recent work by Won et al., which also confirmed the ferrimagnetic order to be the ground states [38]. Therefore, both the structural stability and the magnetic order of bulk VTe2 are accurately captured by GGA (U = 0 eV) rather than GGA + U.
In addition, our analysis extends to the VTe2 monolayer, in which we also ascertain that U = 0 eV remains a prudent selection. As shown in Figure 2, the utilization of U = 0 eV accurately reproduces the band structure and Fermi surface of the T-phase monolayer that was experimentally observed [21,41,42]. Notably, Wang et. al. have successfully reproduced the experimentally determined band structure (i.e., ARPES) of the T-phase monolayer with the GGA (U = 0 eV) method [41]. Consequently, based on the collective evidence, we confidently affirm that the GGA (U = 0 eV) method stands as a rational choice for modeling VTe2 properties, and all of the following results are obtained by calculations based on this choice.

3.2. The Structural, Electronic, and Magnetic Attributes of Monolayer T″-Phase VTe2

After choosing the optimum Hubbard U values, in the subsequent sections, our investigation will focus on the structural, electronic, and distinctly magnetic attributes of monolayer T″-phase VTe2, using the T-phase monolayer as a comparative benchmark. We prioritize the examination of ML T″-phase VTe2 for two primary reasons. Firstly, since bulk T″-phase VTe2 exhibits antiferromagnetism, it is reasonable to anticipate the persistence of magnetism in the monolayer counterpart. Secondly, the diminished geometrical symmetries in ML T″-VTe2, in contrast to the honeycomb V atom lattice in ML T-VTe2, are poised to introduce additional magnetic anisotropy; thus, promoting stable, long-range ferromagnetic order.
First, we study the strain-dependent energetic and magnetic properties of both bulk and monolayer T″-phase VTe2. Within ML T″-phase VTe2, notably abrupt changes occur in both magnetic moment and energy as the in-plane lattice constant a traverses a critical threshold value around a0 (a0 = 18.687 Å) (see Figure 3a,c). In contrast, both the magnetic moment and energy of the bulk T″-phase VTe2 exhibit a smoother dependence on strain (see Figure 3b). Upon closer examination, we identify two distinct ferromagnetic states within ML T″-phase VTe2. Figure 3c denotes the two kinds of ferromagnetic states of ML T″-phase VTe2: the T″-1 state at the side of compressive and the T″-2 state at the side of tensile strain, respectively. Similarly, the magnetic moment of the middle V atoms (V1) is larger than the neighboring two V atoms (V2 and V3) for both the T″-1 and T″-2 states. However, there are big differences in the magnitude of the magnetic moments for these two states, that is, the magnetic moments of V1, V2, and V3 in ML T″-1 VTe2 are 0.70 μ B , 0.23 μ B , and 0.23 μ B , respectively (see Figure 3e), while the magnetic moment of V1, V2, and V3 in ML T″-2 VTe2 are 1.46 μ B , 1.23 μ B , and 1.23 μ B , respectively (see Figure 3f). As shown in Figure 3d, the fully relaxed structure of ML T″-1 VTe2 exhibits a smaller surface area than ML T″-2 VTe2, which is consistent with the result that the T″-1 (T″-2) state appeared in the compressive (tensile) area (see Figure 3c). The surface area of T″-2 is in close proximity to the ML T-phase VTe2, although the corresponding surface area of bulk T″-VTe2 is smaller than bulk T-phase VTe2.
The above results suggest V atoms in T″-1 (T″-2) are more delocalized (localized), which is caused by the smaller (larger) V-V distance under compressive (tensile) strain. This could be reflected by the hopping energies (t) between the V atoms. A direct estimation of the hopping energies requires the tight binding method, which is out of the scope of this work, but the hopping energies are known to be proportionally linear to the bandwidth. Moreover, as shown in Figure 4, the bandwidths of the d orbital around the Fermi level for ML T″-1 (d1) and T″-2 VTe2 (d2) are 0.73 and 0.62 eV, respectively. Considering we are using the same U values for these two systems, the spin polarization should be inversely proportional to t, which is in line with the larger magnetism in T″-2. Despite the minor change in atomic configuration, the density of states for ML T″-1 and T″-2 VTe2 exhibit an evident difference. For example, the T″-1 has more spin-down density states located below the Fermi level than T″-2. Consequently, ML T″-2 VTe2 exhibits larger spins splitting between the spin-up and spin-down channels than that of the ML T″-1 VTe2 (see Figure 4). This observation concurs with the anticipation that the former would possess a significantly larger average magnetic moment than the latter. As T″-1 VTe2 is more energetically stable than T″-2 phase when no strain is applied, we will further study the magnetic properties of T″-1 VTe2 in the following section.
To determine the magnetic ground state of the T″-1 VTe2, we examined one ferromagnetic and four antiferromagnetic configurations (refer to Figure 5). In Table 2 are the corresponding total energies, which have been decomposed into the nearest-neighboring V-V magnetic exchange energies (J1, J2, J3, and J4) and the energy from all the other contributions. Our findings validate that the ferromagnetic coupling represents the ground magnetic state of the T″-1 VTe2.
As shown in Table 2, the strongest magnetic interaction is the FM coupling (−20.93 meV per link) between the V2 atoms. Due to the symmetry, the same coupling is expected between the V3 atoms. Notably, the coupling between the V1 chains with either the V2 or V3 chains is also significant (−15.615 meV per link). On the contrary, the couplings between V1 atoms and between V2 and V3 chains exhibit an antiferromagnetic nature, albeit with orders of magnitude smaller than the FM couplings. Thus, we could expect that an overall FM coupling is preferred for the systems, a hypothesis that will be further corroborated by the explicit Monte Carlo simulations detailed below.
The strong coupling between magnetic moments does not necessarily lead to long-range magnetic ordering at a finite temperature in low-dimensional systems unless magnetic anisotropy is included in the system. In contrast to the structural isotropy along the lattices a and b for T-phase VTe2, the 3 × 1 reconstruction in T″-1 VTe2 breaks such symmetry, which may induce magnetic anisotropy in the system. Consequently, we performed simulations on the magnetic anisotropy energy (MAE) of both the ML T- and the T″-1 VTe2.
As shown in Figure 6, the ML T-phase VTe2 exhibits an easy-plane anisotropy, with its magnetic moments preferring to stay in the ab plane (in-plane) rather than along the c direction, while the energy remains isotropic within the ab plane (see Figure 6a,b). On the contrary, apart from its inclination to remain in the in-plane over the out-of-plane direction, the ML T″-1 VTe2 also exhibits an anisotropy within the plane. Specifically, it demonstrates an easy-axis anisotropy perpendicular to the chain (along the a axis) and a substantial magnetic anisotropy energy of 165 μ e V (Figure 6c,d). Therefore, ML T″-1 VTe2 promises to be a stable ferromagnetic material while ML T-phase VTe2 does not exhibit stable long-range ferromagnetic order at finite temperatures.
To precisely determine the Curie temperature, we employed the Heisenberg spin Hamiltonian, which can consider the impacts of both spin–spin interactions and magnetic anisotropy on the temperature-dependent spin behavior of the monolayer T″-1 VTe2 system. This analysis was conducted through Monte Carlo Metropolis simulations utilizing the VAMPIRE software package [48]. Figure 7 illustrates the temperature-dependent magnetization obtained from these simulations. The simulated temperature-dependent magnetization was further fitted to the Curie–Bloch equation in the classical limit, m T = ( 1 T / T c ) β , where T c is the Curie temperature and β is the critical exponent. As shown in Figure 7, there is large net magnetism at low temperatures, demonstrating the pinning of the magnetic moments to a specific direction due to the easy-axis magnetic anisotropy. However, the normalized magnetism progressively decreases with temperature, transitioning to a paramagnetic state at the critical Curie temperature T c of around 191 K. Notably, this Curie temperature significantly surpasses that of the CrI3 monolayer (45 K), confirming the stabilization of long-range collective magnetism in the T″-1 VTe2. This stabilization is attributed to the magnetic anisotropy stemming from periodic structural distortions within the material.

4. Conclusions

In this study, we conducted a systematic analysis of the correlation between structure and magnetism and their dependence on strain in monolayer (ML) T″-phase VTe2. By setting the Hubbard U parameter to 0 eV, we were able to replicate the experimentally observed properties of both bulk and monolayer T″-phase VTe2 accurately. Under applied strain, we identified two distinct ferromagnetic states in the monolayer T″-phase VTe2: the T″-1 state, which emerges under compressive strain, and the T″-2 state, which appears under tensile strain. These states are characterized by minor structural differences yet exhibit significant variations in their magnetic moments, with the T″-1 state displaying lower magnetic moments and the T″-2 state showing higher magnetic moments. Interestingly, in the absence of strain, the T″-1 VTe2 is energetically more favorable compared to the T″-2 phase.
Furthermore, our comparison of magnetic anisotropy between the T and T″-1 phases of VTe2 reveals that periodic lattice distortion in the T″-1 phase leads to in-plane anisotropy, making the T″-1 VTe2 an easy-axis magnetic material. Monte Carlo simulations support our findings, highlighting a remarkably high Curie temperature of approximately 191 K for T″-1 VTe2. These insights enhance our understanding of the VTe2 system. In addition, our study presents a new and promising direction for stabilizing low-dimensional magnetism, which holds promise for spintronics applications due to its enhanced stability of ferromagnetic ordering in thin layers, facilitating efficient manipulation and control of spin-polarized currents for advanced electronic devices.

Author Contributions

Conceptualization, Y.L., S.W. and X.S.; methodology, X.T. and X.S.; software, J.Z. and X.T.; validation, J.Z. and X.S.; investigation, X.T., J.Z. and X.S.; resources, Y.L. and S.W.; data curation, X.T., J.Z. and X.S.; writing—original draft preparation, X.T., J.Z. and X.S.; writing—review and editing, All authors; supervision, Y.L. and S.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 52373227) and Shanghai Technical Service Center of Science and Engineering Computing, Shanghai University.

Data Availability Statement

Data are contained within the article. The data that support the findings of this study are available on request from the corresponding author.

Acknowledgments

We acknowledge the Centre for Advanced 2D Materials and the Graphene Research Center at the National University of Singapore and the National Supercomputing Centre, Singapore, for providing the computing resources.

Conflicts of Interest

Author Xiaohe Song was employed by the company EACOMP. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Gong, C.; Li, L.; Li, Z.; Ji, H.; Stern, A.; Xia, Y.; Cao, T.; Bao, W.; Wang, C.; Wang, Y.; et al. Discovery of intrinsic ferromagnetism in two-dimensional van der Waals crystals. Nature 2017, 546, 265–269. [Google Scholar] [CrossRef] [PubMed]
  2. Huang, B.; Clark, G.; Navarro-Moratalla, E.; Klein, D.R.; Cheng, R.; Seyler, K.L.; Zhong, D.; Schmidgall, E.; McGuire, M.A.; Cobden, D.H.; et al. Layer-dependent ferromagnetism in a van der Waals crystal down to the monolayer limit. Nature 2017, 546, 270–273. [Google Scholar] [CrossRef] [PubMed]
  3. 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] [PubMed]
  4. Matsukura, F.; Tokura, Y.; Ohno, H. Control of magnetism by electric fields. Nat. Nanotechnol. 2015, 10, 209–220. [Google Scholar] [CrossRef] [PubMed]
  5. Feng, Y.P.; Shen, L.; Yang, M.; Wang, A.; Zeng, M.; Wu, Q.; Chintalapati, S.; Chang, C.-R. Prospects of spintronics based on 2D materials. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2017, 7, e1313. [Google Scholar] [CrossRef]
  6. Cheng, H.-X.; Zhou, J.; Ji, W.; Zhang, Y.-N.; Feng, Y.-P. Two-dimensional intrinsic ferrovalley GdI2 with large valley polarization. Phys. Rev. B 2021, 103, 125121. [Google Scholar] [CrossRef]
  7. Deng, Y.; Yu, Y.; Song, Y.; Zhang, J.; Wang, N.Z.; Sun, Z.; Yi, Y.; Wu, Y.Z.; Wu, S.; Zhu, J.; et al. Gate-tunable room-temperature ferromagnetism in two-dimensional Fe3GeTe2. Nature 2018, 563, 94–99. [Google Scholar] [CrossRef] [PubMed]
  8. 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]
  9. Miao, N.; Xu, B.; Zhu, L.; Zhou, J.; Sun, Z. 2D Intrinsic Ferromagnets from van der Waals Antiferromagnets. J. Am. Chem. Soc. 2018, 140, 2417–2420. [Google Scholar] [CrossRef] [PubMed]
  10. Jiang, Z.; Wang, P.; Xing, J.; Jiang, X.; Zhao, J. Screening and Design of Novel 2D Ferromagnetic Materials with High Curie Temperature above Room Temperature. ACS Appl. Mater. Interfaces 2018, 10, 39032–39039. [Google Scholar] [CrossRef] [PubMed]
  11. Zhou, J.; Feng, Y.P.; Shen, L. Atomic-orbital-free intrinsic ferromagnetism in electrenes. Phys. Rev. B 2020, 102, 180407(R). [Google Scholar] [CrossRef]
  12. Gong, C.; Zhang, X. Two-dimensional magnetic crystals and emergent heterostructure devices. Science 2019, 363, eaav4450. [Google Scholar] [CrossRef] [PubMed]
  13. Lu, S.; Zhou, Q.; Guo, Y.; Zhang, Y.; Wu, Y.; Wang, J. Coupling a Crystal Graph Multilayer Descriptor to Active Learning for Rapid Discovery of 2D Ferromagnetic Semiconductors/Half-Metals/Metals. Adv. Mater. 2020, 32, e2002658. [Google Scholar] [CrossRef] [PubMed]
  14. Mermin, N.D.; Wagner, H. Absence of Ferromagnetism or Antiferromagnetism in One- or Two-Dimensional Isotropic Heisenberg Models. Phys. Rev. Lett. 1966, 17, 1133–1136. [Google Scholar] [CrossRef]
  15. Gibertini, M.; Koperski, M.; Morpurgo, A.F.; Novoselov, K.S. Magnetic 2D materials and heterostructures. Nat. Nanotechnol. 2019, 14, 408–419. [Google Scholar] [CrossRef] [PubMed]
  16. Torelli, D.; Thygesen, K.S.; Olsen, T. High throughput computational screening for 2D ferromagnetic materials: The critical role of anisotropy and local correlations. 2D Mater. 2019, 6, 045018. [Google Scholar] [CrossRef]
  17. Chua, R.; Zhou, J.; Yu, X.; Yu, W.; Gou, J.; Zhu, R.; Zhang, L.; Liu, M.; Breese, M.B.H.; Chen, W. Room Temperature Ferromagnetism of Monolayer Chromium Telluride with Perpendicular Magnetic Anisotropy. Adv. Mater. 2021, 33, e2103360. [Google Scholar] [CrossRef] [PubMed]
  18. Zhou, J.; Song, X.; Yang, M.; Chai, J.; Wong, N.L.M.; Shen, L.; Wang, S.; Feng, Y.P. A first principles study of uniaxial strain-stabilized long–range ferromagnetic ordering in electrenes. J. Mater. Chem. C 2021, 9, 16576–16580. [Google Scholar] [CrossRef]
  19. Niu, C.; Qiu, G.; Wang, Y.; Zhang, Z.; Si, M.; Wu, W.; Ye, P.D. Gate-tunable strong spin-orbit interaction in two-dimensional tellurium probed by weak antilocalization. Phys. Rev. B 2020, 101, 205414. [Google Scholar] [CrossRef]
  20. Su, J.; Liu, K.; Wang, F.; Jin, B.; Guo, Y.; Liu, G.; Li, H.; Zhai, T. van der Waals 2D transition metal tellurides. Adv. Mater. Interfaces 2019, 6, 1900741. [Google Scholar] [CrossRef]
  21. Mitsuishi, N.; Sugita, Y.; Bahramy, M.S.; Kamitani, M.; Sonobe, T.; Sakano, M.; Shimojima, T.; Takahashi, H.; Sakai, H.; Horiba, K.; et al. Switching of band inversion and topological surface states by charge density wave. Nat. Commun. 2020, 11, 2466. [Google Scholar] [CrossRef] [PubMed]
  22. Naik, M.H.; Regan, E.C.; Zhang, Z.; Chan, Y.-H.; Li, Z.; Wang, D.; Yoon, Y.; Ong, C.S.; Zhao, W.; Zhao, S. Intralayer charge-transfer moiré excitons in van der Waals superlattices. Nature 2022, 609, 52–57. [Google Scholar] [CrossRef] [PubMed]
  23. 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]
  24. Whangbo, M.H.; Canadell, E. Analogies between the concepts of molecular chemistry and solid-state physics concerning structural instabilities. Electronic origin of the structural modulations in layered transition metal dichalcogenides. J. Am. Chem. Soc. 1992, 114, 9587–9600. [Google Scholar] [CrossRef]
  25. Manzeli, S.; Ovchinnikov, D.; Pasquier, D.; Yazyev, O.V.; Kis, A. 2D transition metal dichalcogenides. Nat. Rev. Mater. 2017, 2, 17033. [Google Scholar] [CrossRef]
  26. Strocov, V.N.; Shi, M.; Kobayashi, M.; Monney, C.; Wang, X.; Krempasky, J.; Schmitt, T.; Patthey, L.; Berger, H.; Blaha, P. Three-dimensional electron realm in VSe2 by soft-X-ray photoelectron spectroscopy: Origin of charge-density waves. Phys. Rev. Lett. 2012, 109, 086401. [Google Scholar] [CrossRef] [PubMed]
  27. Si, J.G.; Lu, W.J.; Wu, H.Y.; Lv, H.Y.; Liang, X.; Li, Q.J.; Sun, Y.P. Origin of the multiple charge density wave order in 1T−VSe2. Phys. Rev. B 2020, 101, 235405. [Google Scholar] [CrossRef]
  28. Zhang, K.; Si, C.; Lian, C.-S.; Zhou, J.; Sun, Z. Mottness collapse in monolayer 1T-TaSe2 with persisting charge density wave order. J. Mater. Chem. C 2020, 8, 9742–9747. [Google Scholar] [CrossRef]
  29. Li, L.J.; O’Farrell, E.C.; Loh, K.P.; Eda, G.; Ozyilmaz, B.; Castro Neto, A.H. Controlling many-body states by the electric-field effect in a two-dimensional material. Nature 2016, 529, 185–189. [Google Scholar] [CrossRef] [PubMed]
  30. Grüner, G. The dynamics of charge-density waves. Rev. Mod. Phys. 1988, 60, 1129. [Google Scholar] [CrossRef]
  31. Qian, X.; Liu, J.; Fu, L.; Li, J. Quantum spin Hall effect in two-dimensional transition metal dichalcogenides. Science 2014, 346, 1344–1347. [Google Scholar] [CrossRef] [PubMed]
  32. Kim, H.; Shim, J.H.; Kim, S.; Park, J.H.; Kim, K.; Min, B.I. Unusual Pressure-Induced Quantum Phase Transition from Superconducting to Charge-Density Wave State in Rare-Earth-Based Heusler LuPd2 In Compound. Phys. Rev. Lett. 2020, 125, 157001. [Google Scholar] [CrossRef] [PubMed]
  33. Kusmartseva, A.F.; Sipos, B.; Berger, H.; Forro, L.; Tutis, E. Pressure induced superconductivity in pristine 1T-TiSe2. Phys. Rev. Lett. 2009, 103, 236401. [Google Scholar] [CrossRef] [PubMed]
  34. Chen, C.; Singh, B.; Lin, H.; Pereira, V.M. Reproduction of the Charge Density Wave Phase Diagram in 1T-TiSe2 Exposes its Excitonic Character. Phys. Rev. Lett. 2018, 121, 226602. [Google Scholar] [CrossRef] [PubMed]
  35. Kolincio, K.K.; Roman, M.; Klimczuk, T. Enhanced Mobility and Large Linear Nonsaturating Magnetoresistance in the Magnetically Ordered States of TmNiC2. Phys. Rev. Lett. 2020, 125, 176601. [Google Scholar] [CrossRef] [PubMed]
  36. Stoner, E.C. Collective electron ferronmagnetism. Proc. R. Soc. Lond. A 1938, 165, 372–414. [Google Scholar]
  37. Zhou, J.; Wang, Z.; Wang, S.; Feng, Y.P.; Yang, M.; Shen, L. Coexistence of ferromagnetism and charge density waves in monolayer LaBr2. Nanoscale Horiz. 2023, 8, 1054–1061. [Google Scholar] [CrossRef] [PubMed]
  38. Won, D.; Kiem, D.H.; Cho, H.; Kim, D.; Kim, Y.; Jeong, M.Y.; Seo, C.; Kim, J.; Park, J.G.; Han, M.J.; et al. Polymorphic spin, charge, and lattice waves in vanadium ditelluride. Adv. Mater. 2020, 32, e1906578. [Google Scholar] [CrossRef]
  39. Coelho, P.M.; Lasek, K.; Nguyen Cong, K.; Li, J.; Niu, W.; Liu, W.; Oleynik, I.I.; Batzill, M. Monolayer modification of VTe2 and its charge density wave. J. Phys. Chem. Lett. 2019, 10, 4987–4993. [Google Scholar] [CrossRef] [PubMed]
  40. Dai, T.; Kang, S.; Ma, X.; Dang, S.; Li, H.; Ruan, Z.; Zhou, W.; Hu, P.; Li, S.; Wu, S. Multiple Transitions of Charge Density Wave Order in Epitaxial Few-Layered 1T′-VTe2 Films. J. Phys. Chem. C 2019, 123, 18711–18716. [Google Scholar] [CrossRef]
  41. Wang, Y.; Ren, J.; Li, J.; Wang, Y.; Peng, H.; Yu, P.; Duan, W.; Zhou, S. Evidence of charge density wave with anisotropic gap in a monolayer VTe2 film. Phys. Rev. B 2019, 100, 241404. [Google Scholar] [CrossRef]
  42. Sugawara, K.; Nakata, Y.; Fujii, K.; Nakayama, K.; Souma, S.; Takahashi, T.; Sato, T. Monolayer VTe2: Incommensurate Fermi surface nesting and suppression of charge density waves. Phys. Rev. B 2019, 99, 241404. [Google Scholar] [CrossRef]
  43. Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558–561. [Google Scholar] [CrossRef] [PubMed]
  44. Kresse, G.; Hafner, J. Ab initio molecular-dynamics simulation of the liquid-metal--amorphous-semiconductor transition in germanium. Phys. Rev. B 1994, 49, 14251–14269. [Google Scholar] [CrossRef] [PubMed]
  45. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  46. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed]
  47. Evans, R.F.; Atxitia, U.; Chantrell, R.W. Quantitative simulation of temperature-dependent magnetization dynamics and equilibrium properties of elemental ferromagnets. Phys. Rev. B 2015, 91, 144425. [Google Scholar] [CrossRef]
  48. Evans, R.F.; Fan, W.J.; Chureemart, P.; Ostler, T.A.; Ellis, M.O.; Chantrell, R.W. Atomistic spin model simulations of magnetic nanomaterials. J. Phys. Condens. Matter 2014, 26, 103202. [Google Scholar] [CrossRef] [PubMed]
  49. Bronsema, K.D.; Bus, G.W.; Wiegers, G.A. The crystal structure of vanadium ditelluride, V1+xTe2. J. Solid State Chem. 1984, 53, 415–421. [Google Scholar] [CrossRef]
  50. Han, G.H.; Duong, D.L.; Keum, D.H.; Yun, S.J.; Lee, Y.H. van der Waals metallic transition metal dichalcogenides. Chem. Rev. 2018, 118, 6297–6336. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. The structures of (a) T-phase, (b) T′-phase, and (c) T″-phase VTe2. The red and brown balls represent the V and Te atoms, respectively. The black rhombus and rectangles are the unit cells for the three phases. Each subplot includes a top view on the left and a side view on the right. In the T′-phase VTe2, single zigzag chains (indicated by blue lines in (b)) are formed by the head-to-head movement of the V pairs, while the T″ phase is a 3 × 1 reconstruction with the two V atoms on the two sides moving head-to-head towards the middle V atom, forming double zigzag chains (indicated by blues in (c)).
Scheme 1. The structures of (a) T-phase, (b) T′-phase, and (c) T″-phase VTe2. The red and brown balls represent the V and Te atoms, respectively. The black rhombus and rectangles are the unit cells for the three phases. Each subplot includes a top view on the left and a side view on the right. In the T′-phase VTe2, single zigzag chains (indicated by blue lines in (b)) are formed by the head-to-head movement of the V pairs, while the T″ phase is a 3 × 1 reconstruction with the two V atoms on the two sides moving head-to-head towards the middle V atom, forming double zigzag chains (indicated by blues in (c)).
Nanomaterials 14 00704 sch001
Figure 1. The evolution of preferred phase, relative phase stability, and magnetic order in bulk VTe2 as a function of U. (a) Evolution of the crystalline phase, magnetic moment per V atom, and volume per three VTe2 units as a function of U values, with bulk T″-phase VTe2 as the starting point for the fully structural relaxation in DFT calculations. (b) The relative stability of bulk T″-phase VTe2 versus bulk T-phase VTe2 at different U values. (c) Comparison of the calculated average magnetic moments for bulk T−- and bulk T″-phase VTe2 with fixed atomic positions during DFT calculations. (d) The magnetic moment of V atoms (V1 and V2 denote two types of V atoms shown in (e)) in the T″-phase VTe2 bulk at different U. The AFM and FM within the image refer to the magnetic configurations illustrated in (e,f). (e,f) illustrate the magnetic ordering in each layer of bulk T″-phase VTe2 at U = 0 eV and U = 2 eV, respectively. The red and brown balls represent the V and Te atoms, respectively. The cyan color indicates down-spin density, and the yellow color denotes up-spin density.
Figure 1. The evolution of preferred phase, relative phase stability, and magnetic order in bulk VTe2 as a function of U. (a) Evolution of the crystalline phase, magnetic moment per V atom, and volume per three VTe2 units as a function of U values, with bulk T″-phase VTe2 as the starting point for the fully structural relaxation in DFT calculations. (b) The relative stability of bulk T″-phase VTe2 versus bulk T-phase VTe2 at different U values. (c) Comparison of the calculated average magnetic moments for bulk T−- and bulk T″-phase VTe2 with fixed atomic positions during DFT calculations. (d) The magnetic moment of V atoms (V1 and V2 denote two types of V atoms shown in (e)) in the T″-phase VTe2 bulk at different U. The AFM and FM within the image refer to the magnetic configurations illustrated in (e,f). (e,f) illustrate the magnetic ordering in each layer of bulk T″-phase VTe2 at U = 0 eV and U = 2 eV, respectively. The red and brown balls represent the V and Te atoms, respectively. The cyan color indicates down-spin density, and the yellow color denotes up-spin density.
Nanomaterials 14 00704 g001
Figure 2. (a) Calculated band structure with projection on V and Te, and (b) Fermi surface of the T-phase VTe2 monolayer by using U = 0 eV. The Fermi is aligned to 0 eV. The color bar indicates the density of states at the Fermi level. The red (blue) color denotes high (low) density of states. The band structure could well match the energy band detected by the ARPES detection. Moreover, the experimentally detected Fermi surface could also be accurately reproduced [21,41,42].
Figure 2. (a) Calculated band structure with projection on V and Te, and (b) Fermi surface of the T-phase VTe2 monolayer by using U = 0 eV. The Fermi is aligned to 0 eV. The color bar indicates the density of states at the Fermi level. The red (blue) color denotes high (low) density of states. The band structure could well match the energy band detected by the ARPES detection. Moreover, the experimentally detected Fermi surface could also be accurately reproduced [21,41,42].
Nanomaterials 14 00704 g002
Figure 3. (a) Evolution of the relative energy and magnetic moment per formula unit in (a) ML T″-phase VTe2 and (b) bulk T″-phase VTe2 as a function of strain. (c) A closer examination of figure (a) by narrowing the ratio a/a0 into the range [0.98, 1.04]. It is noted that a0 = 18.687 Å and 18.814 Å for monolayer and bulk T″-phase VTe2. T″-1 and T″-2 denotes the two kinds of magnetic states. (d) The surface area of ML T″-1 and T″-2 VTe2 compared to bulk T″-phase VTe2. The surface area of ML and bulk T-phase VTe2 are given for comparison. The surface area is defined by the in-plane area of the unit cell with the c axis perpendicular to the VTe2 monolayers. (e) Magnetic order of ML T″-1 VTe2. Magnetic moments of V1, V2, and V3 are 0.70 μ B , 0.23 μ B , and 0.23 μ B , respectively. (f) Magnetic order of ML T″-2 VTe2. Magnetic moments of V1, V2, and V3 are 1.46 μ B , 1.23 μ B , 1.23 μ B , respectively. The red and brown balls represent the V and Te atoms, respectively. It is noted that the ground-state structures of ML T″ VTe2 have the lattice constant a of 18.687 (18.786) Å for T″-1 (T″-2).
Figure 3. (a) Evolution of the relative energy and magnetic moment per formula unit in (a) ML T″-phase VTe2 and (b) bulk T″-phase VTe2 as a function of strain. (c) A closer examination of figure (a) by narrowing the ratio a/a0 into the range [0.98, 1.04]. It is noted that a0 = 18.687 Å and 18.814 Å for monolayer and bulk T″-phase VTe2. T″-1 and T″-2 denotes the two kinds of magnetic states. (d) The surface area of ML T″-1 and T″-2 VTe2 compared to bulk T″-phase VTe2. The surface area of ML and bulk T-phase VTe2 are given for comparison. The surface area is defined by the in-plane area of the unit cell with the c axis perpendicular to the VTe2 monolayers. (e) Magnetic order of ML T″-1 VTe2. Magnetic moments of V1, V2, and V3 are 0.70 μ B , 0.23 μ B , and 0.23 μ B , respectively. (f) Magnetic order of ML T″-2 VTe2. Magnetic moments of V1, V2, and V3 are 1.46 μ B , 1.23 μ B , 1.23 μ B , respectively. The red and brown balls represent the V and Te atoms, respectively. It is noted that the ground-state structures of ML T″ VTe2 have the lattice constant a of 18.687 (18.786) Å for T″-1 (T″-2).
Nanomaterials 14 00704 g003
Figure 4. (a) A comparison of the spin-polarized total density of states (DOS) and projected density of states (PDOS) for (a) ML T″-1 VTe2 and (b) ML T″-2 VTe2. The Fermi level is aligned to 0 eV. d1 and d2 are the bandwidths of the d orbital around the Fermi level for ML T″-1 and T″-2 VTe2, respectively.
Figure 4. (a) A comparison of the spin-polarized total density of states (DOS) and projected density of states (PDOS) for (a) ML T″-1 VTe2 and (b) ML T″-2 VTe2. The Fermi level is aligned to 0 eV. d1 and d2 are the bandwidths of the d orbital around the Fermi level for ML T″-1 and T″-2 VTe2, respectively.
Nanomaterials 14 00704 g004
Figure 5. The (a) ferromagnetic and (b) four antiferromagnetic configurations adopted for T″-1 VTe2. The red and brown balls represent V and Te atoms, respectively. J1, J2, J3, and J4 are the four unique types of exchange energies between the nearest neighboring V-V exchange paths in ML T″-1 VTe2. The yellow and cyan colors denote the up-polarized spin density and down-polarized spin density, respectively. Note that the FM order is the ground state magnetic configuration.
Figure 5. The (a) ferromagnetic and (b) four antiferromagnetic configurations adopted for T″-1 VTe2. The red and brown balls represent V and Te atoms, respectively. J1, J2, J3, and J4 are the four unique types of exchange energies between the nearest neighboring V-V exchange paths in ML T″-1 VTe2. The yellow and cyan colors denote the up-polarized spin density and down-polarized spin density, respectively. Note that the FM order is the ground state magnetic configuration.
Nanomaterials 14 00704 g005
Figure 6. Magnetic anisotropy of ML T- and T″-1 VTe2. (a,b) illustrate the relative energy as a function of the spin rotated angles (red squares) for ML T-phase VTe2 in the ac plane (out of plane) and the ab plane (in plane), respectively. (c,d) illustrate the relative energy as a function of the spin rotated angles for ML T″-1 VTe2 in the ac plane and the ab plane, respectively. The black curves are a guide for the eyes.
Figure 6. Magnetic anisotropy of ML T- and T″-1 VTe2. (a,b) illustrate the relative energy as a function of the spin rotated angles (red squares) for ML T-phase VTe2 in the ac plane (out of plane) and the ab plane (in plane), respectively. (c,d) illustrate the relative energy as a function of the spin rotated angles for ML T″-1 VTe2 in the ac plane and the ab plane, respectively. The black curves are a guide for the eyes.
Nanomaterials 14 00704 g006
Figure 7. The temperature dependence of normalized magnetization for monolayer T″-1 VTe2 (blue solid squares). The red curve is the fitted Curie–Bloch equation in the classical limit from the simulated temperature-dependent magnetization.
Figure 7. The temperature dependence of normalized magnetization for monolayer T″-1 VTe2 (blue solid squares). The red curve is the fitted Curie–Bloch equation in the classical limit from the simulated temperature-dependent magnetization.
Nanomaterials 14 00704 g007
Table 1. The structure parameters of bulk T″-phase VTe2 relaxed at U = 0 eV. The first three rows present the lattice vectors a, b, and c, with the three columns indicating the Cartesian components of the lattice vectors. The subsequent nine rows detail the ion positions, with the three columns specifying the fractional coordinates relative to the three lattice vectors.
Table 1. The structure parameters of bulk T″-phase VTe2 relaxed at U = 0 eV. The first three rows present the lattice vectors a, b, and c, with the three columns indicating the Cartesian components of the lattice vectors. The subsequent nine rows detail the ion positions, with the three columns specifying the fractional coordinates relative to the three lattice vectors.
Latticexyz
a7.440300
b6.58373.46600
c−10.8930−2.69218.5749
Atomabc
V10.34680.34680.70819
V20.65320.65320.29181
V3000
Te10.627940.627940.98386
Te20.372060.372060.01614
Te30.088520.088530.38103
Te40.911470.911480.61897
Te50.704040.704040.69656
Te60.295960.295960.30344
Table 2. The summary of the different magnetic configurations and the deduced magnetic exchange energies. FM, AMF_1, AMF_2, AMF_3, and AMF_4 are the magnetic configurations and J1, J2, J3, and J4 are the exchange energies between the nearest neighboring V-V, as shown in Figure 5. Etot is the total energy of T″-1 VTe2 in different magnetic configurations. The integrals in the first five rows are the number of the respective exchange couplings in specific the magnetic configurations. Positive values indicate the ferromagnetic coupling while negative values are for AFM coupling. The values in the last row are the deduced exchange energies from the different magnetic configurations.
Table 2. The summary of the different magnetic configurations and the deduced magnetic exchange energies. FM, AMF_1, AMF_2, AMF_3, and AMF_4 are the magnetic configurations and J1, J2, J3, and J4 are the exchange energies between the nearest neighboring V-V, as shown in Figure 5. Etot is the total energy of T″-1 VTe2 in different magnetic configurations. The integrals in the first five rows are the number of the respective exchange couplings in specific the magnetic configurations. Positive values indicate the ferromagnetic coupling while negative values are for AFM coupling. The values in the last row are the deduced exchange energies from the different magnetic configurations.
J1 (meV/Link)J2 (meV/Link)J3 (meV/Link)J4 (meV/Link)Etot (eV)
FM81648−194.33
AFM_1−80−40−193.99
AFM_28048−194.08
AFM_38−164−8−193.90
AFM_480−4−8−193.99
2.77−15.61−20.934.85
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

Tang, X.; Zhou, J.; Wong, N.L.M.; Chai, J.; Liu, Y.; Wang, S.; Song, X. Strain-Induced Ferromagnetism in Monolayer T″-Phase VTe2: Unveiling Magnetic States and Anisotropy for Spintronics Advancement. Nanomaterials 2024, 14, 704. https://doi.org/10.3390/nano14080704

AMA Style

Tang X, Zhou J, Wong NLM, Chai J, Liu Y, Wang S, Song X. Strain-Induced Ferromagnetism in Monolayer T″-Phase VTe2: Unveiling Magnetic States and Anisotropy for Spintronics Advancement. Nanomaterials. 2024; 14(8):704. https://doi.org/10.3390/nano14080704

Chicago/Turabian Style

Tang, Xiaoting, Jun Zhou, Nancy Lai Mun Wong, Jianwei Chai, Yi Liu, Shijie Wang, and Xiaohe Song. 2024. "Strain-Induced Ferromagnetism in Monolayer T″-Phase VTe2: Unveiling Magnetic States and Anisotropy for Spintronics Advancement" Nanomaterials 14, no. 8: 704. https://doi.org/10.3390/nano14080704

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