Next Article in Journal
Synergistic Nitrogen-Doping and Defect Engineering in Hard Carbon: Unlocking Ultrahigh Rate Capability and Long-Cycling Stability for Sodium-Ion Battery Anodes
Previous Article in Journal
Evaluation of Physical Properties of Ti-Doped BiFeO3 Thin Films Deposited on Fluorine Tin Oxide and Indium Tin Oxide Substrates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Monolayer TiAlTe3: A Perfect Room-Temperature Valleytronic Semiconductor

1
School of Physics and Technology, Institute of Spintronics, University of Jinan, Jinan 250022, China
2
School of Physics and Physical Engineering, Qufu Normal University, Qufu 273165, China
3
Key Laboratory of High-Effciency and Clean Mechanical Manufacture of MOE, School of Mechanical Engineering, Shandong University, Jinan 250061, China
4
Suzhou Research Institute, Shandong University, Suzhou 215009, China
*
Authors to whom correspondence should be addressed.
Materials 2025, 18(10), 2396; https://doi.org/10.3390/ma18102396
Submission received: 15 April 2025 / Revised: 11 May 2025 / Accepted: 14 May 2025 / Published: 21 May 2025
(This article belongs to the Section Electronic Materials)

Abstract

:
Investigating valley-related physics in rare intrinsic ferromagnetic materials with high-temperature stability and viable synthesis methods is of vital importance for advancing fundamental physics and information technology. Through first-principles calculations, we forecast that monolayer TiAlTe3 has superb structural stability, a ferromagnetic coupling mechanism deriving from direct-exchange and superexchange interactions, and a high magnetic transition temperature. We observed spontaneous valley polarization of 103 meV in the bottom conduction band when monolayer TiAlTe3 is magnetized toward an out-of-plane orientation. Additionally, because of its powerful valley-contrasting Berry curvature, the anomalous valley Hall effect emerges under an in-plane electric field. The cooperation of ferromagnetic coupling, a high magnetic transition temperature, and spontaneous valley polarization makes monolayer TiAlTe3 a promising room-temperature ferrovalley material for use in nanoscale spintronics and valleytronics.

1. Introduction

The valley degree of freedom serves as a local minimum (maximum) in the conduction (valence) band, and it can be leveraged to realize information encoding, high-efficiency processing, and non-volatile storage [1,2,3,4,5,6,7,8,9]. The absence of space inversion symmetry in two-dimensional (2D) transition metal dichalcogenides (TMDs) facilitates the emergence of two degenerate yet non-equivalent valleys, making these materials ideal for investigating valleytronic phenomena. Valley polarization can be induced by negating valley degeneracy, enabling selective valley manipulation in logic-oriented valleytronic applications. Although optical pumping allows the achievement of valley polarization in 2D TMDs [10], the rapid carrier recombination severely restricts this technique’s scalability. In order to break the time-reversal symmetry, magnetism can be employed to provide an alternative pathway for valley polarization, but many specific constraints remain. Valley polarization induced by a 1.0 T magnetic field reaches merely 0.1 meV [11], revealing inefficient magnetic manipulation. Magnetic proximity engineering can successfully negate valley degeneracy [12]. However, it introduces a reconfiguration of substantial electronic bands, which suppresses valley properties. Magnetic doping can cause atoms to cluster, so carriers are scattered to degrade valleytronic device performance [13].
To circumvent these issues, ferromagnetic (FM) materials with valley properties, known as ferrovalley (FV) materials [14], are more desirable. The first-discovered 2D FV material, 2H-VSe2 [14], features an intrinsic FM order, which can break time-reversal symmetry. Also, it was found that monolayer VSe2 exhibits spin–valley coupling, allowing regulatable valley polarization via external fields. Subsequently, investigators have reported numerous FV systems, such as ScBrI [15], RuClBr [16], VCGeN4 [17], GdI2 [18], LaBrI [19], and CeI2 [20]. Cheng et al. found that monolayer GdI2 is a compelling FV material for valleytronic applications [18]. Sheng et al. reported that FM monolayer CeI2 has excellent stability and exhibits large valley polarization [20]. Nevertheless, it is worth noting that some shortcomings remain, including low magnetic transition temperatures, low experimental feasibility, and a low level of valley polarization. These limitations affect the application of FV materials, and the core targets remain high levels of valley polarization, high magnetic transition temperatures, and broad applicability. Designing and discovering ideal materials constitutes a formidable challenge for experimental studies and valleytronic implementations [21,22,23,24].
Recently, monolayer XSi2N4 (X = Mo, W) was fabricated via the chemical vapor deposition method in an experiment [25]. Moreover, monolayer TiTe2 has been obtained via molecular beam epitaxy [26]. Here, inspired by monolayer TiTe2, a novel 2D valleytronic material TiAlTe3 is proposed. Via first-principles calculations, we report that monolayer TiAlTe3 is an excellent FV material with superb structural stability and strong FM coupling. When spin–orbital coupling (SOC) is included and monolayer TiAlTe3 is magnetized toward an out-of-plane orientation, a high level of spontaneous valley polarization, amounting to 103 meV, can be observed in the bottom conduction band. Moreover, because of its powerful valley-contrasting Berry curvature, the anomalous valley Hall effect (AVHE) emerges under the influence of an in-plane electric field. Our investigations indicate that monolayer TiAlTe3 is a promising material for both experimental research and practical applications.

2. Computational Details

Based on density functional theory, first-principles calculations were carried out by means of the projector-augmented wave method using the Vienna Ab initio Simulation Package [27,28,29]. The exchange-correlation potential was set according to the Perdew–Burke–Ernzerhof (PBE) functional of generalized gradient approximation [30]. Spin polarization was included in all calculations. The cutoff energy was set to 500 eV. A vacuum space of 25 Å in the c direction was adopted to avoid interaction between adjacent layers. The total energy and force convergence criteria were set to 10−6 eV and 0.01 eV/Å, respectively. The PBE+U method [31] was applied to describe the Ti-3d orbitals, and U and J were set to 3 and 0 eV, respectively [32,33]. For calibrating the PBE+U method, the Heyd–Scuseria–Ernzerhof (HSE06) hybrid functional [34] was applied. The Brillouin zone was sampled utilizing a Г-centered 17 × 17 × 1 k mesh for the unit cell and a mesh with dimensions of 17 × 9 × 1 for the rectangle supercell. A 4 × 4 × 1 supercell was used to analyze phonon dispersion using PHONOPY software (the version number: 2.15.1) [35,36]. An Ab initio molecular dynamics (AIMD) [37] simulation was performed over a 4 × 4 × 1 supercell at a temperature of 300 K for 8 ps. Berry curvature was calculated using the WANNIER90 package (the version number: 3.1.0) [38,39].

3. Results

Figure 1a,b show the optimized crystal structures of monolayer TiAlTe3 with a hexagonal lattice structure and a P3m1 space group (No. 156). Each unit cell includes one Ti atom, one Al atom, and three Te atoms, and the stacking sequence is Te-Al-Te-Ti-Te. Clearly, broken space inversion symmetry can be observed in monolayer TiAlTe3. The optimized lattice constant a of monolayer TiAlTe3 is 4.07 Å. The thermal and dynamic stabilities were evaluated through AIMD simulation and phonon dispersion, respectively. According to Figure 1c, the absence of an imaginary frequency shows the dynamical stability of monolayer TiAlTe3. The total energy fluctuated near the equilibrium value after 8 ps, and no evident disruption was observed for monolayer TiAlTe3 (see Figure 1d), proving that it is thermally stable at room temperature.
The formation energy is calculated as follows:
E form = E TiAlTe 3 E TiTe 2 1 2 E Al 2 Te 2
Here, E TiAlTe 3 , E TiTe 2 , and E Al 2 Te 2 are the energy of the monolayers TiAlTe3, TiTe2, and α-AlTe, respectively. The E form of monolayer TiAlTe3 is −1.75 eV, indicating that individual TiTe2 and α-AlTe systems cannot be obtained through the spontaneous decomposition of monolayer TiAlTe3. For the hexagonal structure, two independent elastic constants of C11 and C12 were obtained: 52.33 Nm−1 and 15.31 Nm−1, respectively. These Cij meet the Born criterion [40,41], namely, C11 > 0 and C11 > |C12|, verifying the structure’s mechanical stability. Naturally, the fabrication scheme for monolayer TiAlTe3 is discussed due to its excellent stability. We hypothesize that monolayer TiAlTe3 can be synthesized by depositing an α-AlTe structure on monolayer TiTe2 [26].
To investigate the magnetic ground state of monolayer TiAlTe3, the total energy difference between antiferromagnetic (AFM) and FM states was calculated by employing one rectangle supercell. Under a U value of 3 eV, the results show that the FM state is energetically lower than the AFM state by 44.6 meV for monolayer TiAlTe3. The energy differences between FM and AFM states as a function of U value are plotted in Figure S1. The total magnetic moment per unit cell was determined to be 1 μB. The magnetic moment is primarily from the Ti atom (see Figure 1a,b), and thus the contributions of Al and Te atoms are very small, a phenomenon that can be attributed to the special configuration. Each Ti atom is coordinated by six Te atoms, forming one trigonal prismatic. Therefore, according to Figure 1e, the 3d orbitals of each Ti atom split into three portions: e1 (dxy, dx2y2), e2 (dxz, dyz), and a (dz2). The electronic configuration of one Ti atom is 3d24s2. In order to form monolayer TiAlTe3, each Ti atom contributes three electrons to Te atoms. One unpaired electron occupies the lower-energy dz2 orbital, which generates a magnetic moment of 1 μB per Ti atom. We can comprehend the FM coupling mechanism of monolayer TiAlTe3 by assessing superexchange (SE) and direct-exchange (DE) interactions. The length between the nearest-neighbor Ti atoms (dTi-Ti) for monolayer TiAlTe3 is 4.07 Å, indicating that AFM DE interactions cannot occur. In addition, based on the Goodenough–Kanamori–Anderson theory [42,43,44], the SE interactions in next-nearest-neighbor Ti atoms via the Te atom (namely, Ti-Te-Ti) also tend to appear FM coupling. As shown in Figure 1b, the bond angles of Ti-Te-Ti for monolayer TiAlTe3 were discovered to be 88.5° and 90.6° in the upper and bottom SE routes, respectively, both being close to 90°. Hence, the intrinsic FM coupling of monolayer TiAlTe3 is governed by the SE interaction between the Te-5p and Ti-3d orbitals. We conclude that Ti3+-related systems are built by multiple coordinate atoms, which can appear in various phases and exhibit diversity.
On the basis of the Mermin–Wagner theory [45], powerful thermal fluctuations can easily destroy long-range FM ordering. Nevertheless, the 2D ferromagnets Cr2Ge2Te6 [46] and CrI3 [47] exhibit magnetic anisotropy governed by magnetic anisotropy energy (MAE), which plays a significant role in realizing long-range FM ordering. The MAE is estimated using the following equation:
E MAE = E x E z
where E x and E z signify the total energy for the magnetization in IP and out-of-plane (OP) orientations, respectively. The MAE calculated was −418 μeV. This confirms that the easy magnetization axis of monolayer TiAlTe3 is in the IP orientation, indicating there is no energy obstacle when the magnetization rotates in the xy plane. In this situation, monolayer TiAlTe3 is a 2D XY magnet. The Berezinsky–Kosterlitz–Thouless transition can happen at a critical temperature:
T C = 1.335 J / k B
where k B and J are the Boltzmann constant and nearest-neighbor exchange parameter ( J = E AFM E FM / 8 S 2 ), respectively. The J as a function of U value is shown in Figure S2. With S = 1 / 2 and J = 22.3 meV in one rectangle supercell, T C was found to be 344 K. The perfect combination of a sizable MAE and a high magnetic transition temperature renders monolayer TiAlTe3 a prospective platform for developing spintronic and valleytronic devices.
The electronic properties under different U values are investigated (see Figure S3), and the calculation by the HSE06 functional is employed for comparative analysis (see Figure S4). The band structure of monolayer TiAlTe3 without SOC is presented in Figure 2a, exhibiting noticeable spin splitting due to magnetic exchange coupling. Evidently, monolayer TiAlTe3 is an indirect semiconductor with a band gap of 0.59 eV. The conduction band minimum (CBM) is located at +K and −K high-symmetry points, while the valence band maximum (VBM) is located between +K (−K) and Г points. The VBM and CBM come from the same spin-up state. We speculate that a rational carrier doping can induce the movement of the Fermi level ( E f ), so the electronic properties of monolayer TiAlTe3 allow for the creation of a half-metal, which facilitates the development of advanced nanodevices. Additionally, in the bottom conduction band, a pair of energetically degenerate valleys emerges at the +K and −K points, and monolayer TiAlTe3 is a potential FV material. Because the space inversion symmetry is naturally broken, the two degenerate valleys have non-equivalent properties [14].
Significantly, the magnetization orientation can be varied from IP to OP under an external magnetic field. According to Figure 2b, monolayer TiAlTe3 still shows indirect semiconductor properties, and the band gap is 0.47 eV. The SOC negates the degeneracy between +K and −K valleys in the bottom conduction band, and spontaneous valley polarization equal to 103 meV is induced in monolayer TiAlTe3. This corresponds to employing an enormous external magnetic field of about 515–1030 T. The valley polarization of 103 meV is superior to that of numerous other FV systems discovered, such as ScBrI (90 meV) [15], VCGeN4 (46 meV) [17], and LaBrI (59 meV) [19]. According to Figure 2c, when monolayer TiAlTe3 is magnetized toward a −z orientation, the valley polarization and spin are concurrently reversed. The valley states in monolayer TiAlTe3 can be regulated by transforming its magnetization orientation.
The related physical mechanism of spontaneous valley polarization stems from the synergistic interplay between SOC and spin-polarized effects. Firstly, when the spin-polarized effect is not involved, the spin degeneracy can be split by the SOC effect at the +K and −K valleys. Due to the protection of time-reversal symmetry, the opposite spin states at the +K and −K valleys can be observed:
E SO + K E SO K
E SO + K = E SO K
Here, E and ( ) are the energy of the valley in the bottom conduction band and the spin symbol, respectively. Secondly, as shown in Figure 2a, when the SOC effect is excluded, the identical spin symbol in the +K and −K valleys is forced by spin-polarized effect; that is,
E SP + K = E SP K
Therefore, when both SOC and spin-polarized effects are simultaneously included, the +K and −K valleys in monolayer TiAlTe3 do not exhibit energetic degeneracy, resulting in spontaneous valley polarization.
By analyzing the orbital contributions to the +K and −K valleys, a quantitative model linking valley polarization with magnetization orientation was derived. Due to the broken time-reversal symmetry, the interplay between opposite spin states was negated. Therefore, the SOC Hamiltonian is written as follows:
H ^ SOC = λ SOC L ^ S ^ H ^ SOC 0 = λ SOC S ^ z L ^ z cos θ + 1 2 L ^ + e i φ sin θ + 1 2 L ^ e + i φ sin θ
Here, L ^ is the orbital angular momentum, with the coordinates x , y , z , and S ^ is the spin angular momentum, with the coordinates x , y , z . The θ and φ are polar angles. As shown in Figure 3, the +K and −K valleys in the bottom conduction band mainly arise from the Ti-3dxy and Ti-3dx2y2 orbitals, and the group symmetry of monolayer TiAlTe3 is C3v. So, the basis functions are expressed as follows:
ϕ c τ = 3 d x 2 y 2 + i τ 3 d x y / 2
Here, τ = ± 1 and c are the valley index and conduction band, respectively. Based on L ^ and L ^ z , the atomic orbitals can be expressed as follows:
3 d x 2 y 2 = 2 , 2 + 2 , + 2 / 2
3 d x y = i 2 , 2 2 , + 2 / 2
Hence, the energy levels for the +K and −K valleys in the bottom conduction band can be written as follows:
E c + K = ψ c + H ^ SOC 0 ψ c + = 2 , 2 H ^ SOC 0 2 , 2 = 2 λ SOC S ^ z cos θ
E c K = ψ c H ^ SOC 0 ψ c = 2 , + 2 H ^ SOC 0 2 , + 2 = + 2 λ SOC S ^ z cos θ
Valley polarization is expressed as follows:
Δ c = E c K E c + K = 4 λ SOC S ^ z cos θ = 4 α SOC cos θ
Here, α SOC is the constant associated with SOC. The trend of valley polarization derived from the model is in agreement with that from first-principles calculations, as illustrated in Figure 2b.
After the spontaneous valley polarization of monolayer TiAlTe3 was identified, the Berry curvature was explored to determine electronic transport properties. We calculated the Berry curvature via the Kudo formula [48]:
Ω z k = n n n f n 2 Im ψ n k υ ^ x ψ n k ψ n k υ ^ y ψ n k E n E n 2
Here, f n is the Fermi–Dirac distribution function, and ψ n k is the Bloch wave function with an eigenvalue of E n . υ ^ x and υ ^ y are velocity operators in the x and y directions, respectively. The calculated Berry curvature of monolayer TiAlTe3 is plotted in Figure 4a. Apparently, the Berry curvature reveals opposite signs and identical absolute values in the neighborhood of the +K and −K valleys, which indicates the valley-contrasting property of monolayer TiAlTe3.
The Bloch electrons can obtain an anomalous velocity, υ E × Ω , when an IP electric field is applied [49]. The carriers can shift to opposite boundaries owing to powerful valley-contrasting Berry curvature. The immense spontaneous valley polarization is realized in monolayer TiAlTe3, so the E f can be moved between the +K and −K valleys in the bottom conduction band. The observed instances of hole shift in p-type monolayer TiAlTe3 under an IP electric field are displayed in Figure 4b. For magnetization in the +z direction, as illustrated in Figure 4b, the spin-up holes from the +K valley can transversely shift to the left boundary. For magnetization in the −z direction, the spin-down holes from the −K valley move to the right boundary, according to Figure 4b. As a consequence, AVHE is demonstrated in monolayer TiAlTe3. Furthermore, accompanied by the AVHE, extra-spin Hall current and charge can occur, because carriers with the same spin and charge are accumulated on the boundary of the material. The remarkable synergistic combination of valley Hall current, charge, and spin provides a promising pathway for integrating electronics, spintronics, and valleytronics.

4. Conclusions

In conclusion, monolayer TiAlTe3 has theoretically been identified as an FM material with excellent structural stability and a high magnetic transition temperature. Spontaneous valley polarization of 103 meV could be observed when the magnetization orientation is OP. The valley states in monolayer TiAlTe3 could be modulated by transforming its magnetization orientation. Via the doping of the carriers, the AVHE was realized in monolayer TiAlTe3. Our research demonstrates that monolayer TiAlTe3 exhibits spin-valley coupling, offering a potential platform for spintronic and valleytronic explorations.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/ma18102396/s1. Figure S1. The energy differences ( Δ E = E AFM E FM ) between FM and AFM states as a function of U value. Figure S2. The nearest-neighbor exchange parameter J of monolayer TiAlTe3 as a function of U value. Figure S3. Under the FM ground state, the band structures of monolayer TiAlTe3 calculated from PBE+U method without SOC. The U values are chosen as (a) 0 eV, (b) 1 eV, (c) 2 eV, and (d) 4 eV. Figure S4. Under the FM ground state, the band structure of monolayer TiAlTe3 by the HSE06 functional without SOC.

Author Contributions

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

Funding

This work was supported by the National Natural Science Foundation of China (Grant No. 62071200), the Taishan Scholars Project of Shandong Province (Grant No. tsqn202306028), the Shandong Provincial Natural Science Foundation (Grant No. ZR2023QE187), and the Basic Research Program of Jiangsu (Grant No. BK20230254).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rycerz, A.; Tworzydło, J.; Beenakker, C. Valley filter and valley valve in graphene. Nat. Phys. 2007, 3, 172–175. [Google Scholar] [CrossRef]
  2. Mak, K.F.; Xiao, D.; Shan, J. Light-valley interactions in 2D semiconductors. Nat. Photonics 2018, 12, 451–460. [Google Scholar] [CrossRef]
  3. Wolf, S.A.; Awschalom, D.D.; Buhrman, R.A.; Daughton, J.M.; von Molnar, S.; Roukes, M.L.; Chtchelkanova, A.Y.; Treger, D.M. Spintronics: A spin-based electronics vision for the future. Science 2001, 294, 1488–1495. [Google Scholar] [CrossRef]
  4. Schaibley, J.R.; Yu, H.; Clark, G.; Rivera, P.; Ross, J.S.; Seyler, K.L.; Yao, W.; Xu, X. Valleytronics in 2D materials. Nat. Rev. Mater. 2016, 1, 16055. [Google Scholar] [CrossRef]
  5. Vitale, S.A.; Nezich, D.; Varghese, J.O.; Kim, P.; Gedik, N.; Jarillo-Herrero, P.; Xiao, D.; Rothschild, M. Valleytronics: Opportunities, challenges, and paths forward. Small 2018, 14, 1801483. [Google Scholar] [CrossRef] [PubMed]
  6. Xiao, D.; Liu, G.B.; Feng, W.; Xu, X.; Yao, W. Coupled Spin and Valley Physics in Monolayers of MoS2 and Other Group-VI Dichalcogenides. Phys. Rev. Lett. 2012, 108, 196802. [Google Scholar] [CrossRef]
  7. Xiao, D.; Yao, W.; Niu, Q. Valley-contrasting physics in graphene: Magnetic moment and topological transport. Phys. Rev. Lett. 2007, 99, 236809. [Google Scholar] [CrossRef]
  8. Pacchioni, G. Valleytronics with a twist. Nat. Rev. Mater. 2020, 5, 480. [Google Scholar] [CrossRef]
  9. Yao, W.; Xiao, D.; Niu, Q. Valley-dependent optoelectronics from inversion symmetry breaking. Phys. Rev. B 2008, 77, 235406. [Google Scholar] [CrossRef]
  10. Zeng, H.; Dai, J.; Yao, W.; Xiao, D.; Cui, X. Valley polarization in MoS2 monolayers by optical pumping. Nat. Nanotechnol. 2012, 7, 490–493. [Google Scholar] [CrossRef]
  11. MacNeill, D.; Heikes, C.; Mak, K.F.; Anderson, Z.; Kormanyos, A.; Zolyomi, V.; Park, J.; Ralph, D.C. Breaking of Valley Degeneracy by Magnetic Field in Monolayer MoSe2. Phys. Rev. Lett. 2015, 114, 037401. [Google Scholar] [CrossRef] [PubMed]
  12. Norden, T.; Zhao, C.; Zhang, P.; Sabirianov, R.; Petrou, A.; Zeng, H. Giant valley splitting in monolayer WS2 by magnetic proximity effect. Nat. Commun. 2019, 10, 4163. [Google Scholar] [CrossRef] [PubMed]
  13. Peng, R.; Ma, Y.; Zhang, S.; Huang, B.; Dai, Y. Valley Polarization in Janus Single-Layer MoSSe via Magnetic Doping. J. Phys. Chem. Lett. 2018, 9, 3612–3617. [Google Scholar] [CrossRef] [PubMed]
  14. Tong, W.Y.; Gong, S.J.; Wan, X.; Duan, C.G. Concepts of ferrovalley material and anomalous valley Hall effect. Nat. Commun. 2016, 7, 13612. [Google Scholar] [CrossRef]
  15. Jia, K.; Dong, X.J.; Li, S.S.; Ji, W.X.; Zhang, C.W. Spontaneous valley polarization and valley-nonequilibrium quantum anomalous Hall effect in Janus monolayer ScBrI. Nanoscale 2023, 15, 8395–8405. [Google Scholar] [CrossRef]
  16. Sun, H.; Li, S.S.; Ji, W.X.; Zhang, C.W. Valley-dependent topological phase transition and quantum anomalous valley Hall effect in single-layer RuClBr. Phys. Rev. B 2022, 105, 195112. [Google Scholar] [CrossRef]
  17. Jia, K.; Dong, X.J.; Li, S.S.; Ji, W.X.; Zhang, C.W. Tunable abundant valley Hall effect and chiral spin-valley locking in Janus monolayer VCGeN4. Nanoscale 2024, 16, 8639–8649. [Google Scholar] [CrossRef]
  18. 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]
  19. Jiang, P.; Kang, L.; Li, Y.L.; Zheng, X.; Zeng, Z.; Sanvito, S. Prediction of the two-dimensional Janus ferrovalley material LaBrI. Phys. Rev. B 2021, 104, 035430. [Google Scholar] [CrossRef]
  20. Sheng, K.; Chen, Q.; Yuan, H.K.; Wang, Z.Y. Monolayer CeI2: An intrinsic room-temperature ferrovalley semiconductor. Phys. Rev. B 2022, 105, 075304. [Google Scholar] [CrossRef]
  21. Zhang, S.J.; Zhang, C.W.; Zhang, S.F.; Ji, W.X.; Li, P.; Wang, P.J.; Li, S.S.; Yan, S.S. Intrinsic Dirac half-metal and quantum anomalous Hall phase in a hexagonal metal-oxide lattice. Phys. Rev. B 2017, 96, 205433. [Google Scholar] [CrossRef]
  22. Wu, B.; Song, Y.L.; Ji, W.X.; Wang, P.J.; Zhang, S.F.; Zhang, C.W. Quantum anomalous Hall effect in an antiferromagnetic monolayer of MoO. Phys. Rev. B 2023, 107, 214419. [Google Scholar] [CrossRef]
  23. Lou, W.H.; Liu, G.; Ma, X.G.; Yang, C.L.; Feng, L.X.; Liu, Y.; Gao, X.C. Enhancing photocatalytic water splitting via GeC/SGaSnP Z-scheme heterojunctions with built-in electric fields. J. Mater. Chem. A 2025, 13, 4356–4366. [Google Scholar] [CrossRef]
  24. Lou, W.H.; Gao, J.M.; Liu, G.; Ali, A.; Jia, B.N.; Guan, X.N.; Ma, X.G. Enhancing hydrogen evolution reaction performance through defect engineering in WTeX (X = S, Se) monolayers: A first-principles study. Int. J. Hydrog. Energy 2024, 87, 620–629. [Google Scholar] [CrossRef]
  25. Hong, Y.L.; Liu, Z.; Wang, L.; Zhou, T.; Ma, W.; Xu, C.; Feng, S.; Chen, L.; Chen, M.L.; Sun, D.M.; et al. Chemical vapor deposition of layered two-dimensional MoSi2N4 materials. Science 2020, 369, 670–674. [Google Scholar] [CrossRef] [PubMed]
  26. Song, Z.; Lei, B.; Cao, Y.; Qi, J.; Peng, H.; Wang, Q.; Huang, L.; Lu, H.; Lin, X.; Wang, Y.L.; et al. Epitaxial fabrication of two-dimensional TiTe2 monolayer on Au(111) substrate with Te as buffer layer. Chin. Phys. B 2019, 28, 056801. [Google Scholar] [CrossRef]
  27. Kresse, G.; Hafner, J. Ab initio molecular-dynamics simulation of the liquid-metal-amorphous-semiconductor transition in germanium. Phys. Rev. B 1994, 49, 14251. [Google Scholar] [CrossRef]
  28. Blochl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953. [Google Scholar] [CrossRef]
  29. Kresse, G.; Furthmuller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169. [Google Scholar] [CrossRef]
  30. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef]
  31. Dudarev, S.L.; Botton, G.A.; Savrasov, S.Y.; Humphreys, C.J.; Sutton, A.P. Electron-energy-loss spectra and the structural stability of nickel oxide: An LSDA+U study. Phys. Rev. B 1998, 57, 1505. [Google Scholar] [CrossRef]
  32. Yekta, Y.; Hadipour, H.; Şaşıoğlu, E.; Friedrich, C.; Jafari, S.A.; Blügel, S.; Mertig, I. Strength of effective Coulomb interaction in two-dimensional transition-metal halides MX2 and MX3 (M = Ti, V, Cr, Mn, Fe, Co, Ni; X = Cl, Br, I). Phys. Rev. Mater. 2021, 5, 034001. [Google Scholar] [CrossRef]
  33. Sheng, K.; Zhang, B.; Wang, Z.Y. Valley Polarization in a TwoDimensional High-Temperature Semiconducting TiInTe3 Honeycomb Ferromagnet. Acta Mater. 2024, 262, 119461. [Google Scholar] [CrossRef]
  34. Krukau, A.V.; Vydrov, O.A.; Izmaylov, A.F.; Scuseria, G.E. Influence of the exchange screening parameter on the performance of screened hybrid functionals. J. Chem. Phys. 2006, 125, 224106. [Google Scholar] [CrossRef] [PubMed]
  35. Anisimov, V.I.; Zaanen, J.; Andersen, O.K. Band theory and Mott insulators: Hubbard U instead of Stoner I. Phys. Rev. B 1991, 44, 943. [Google Scholar] [CrossRef]
  36. Zhang, Y.; Lin, L.F.; Moreo, A.; Dagotto, E. Theoretical study of the crystal and electronic properties of α-RuI3. Phys. Rev. B 2022, 105, 085107. [Google Scholar] [CrossRef]
  37. Bucher, D.; Pierce, L.C.T.; McCammon, J.A.; Markwick, P.R.L. On the use of accelerated molecular dynamics to enhance configurational sampling in ab initio simulations. J. Chem. Theory Comput. 2011, 7, 890–897. [Google Scholar] [CrossRef]
  38. Wu, X.; Vanderbilt, D.; Hamann, D.R. Systematic treatment of displacements, strains, and electric fields in density-functional perturbation theory. Phys. Rev. B 2005, 72, 035105. [Google Scholar] [CrossRef]
  39. Togo, A.; Oba, F.; Tanaka, I. First-principles calculations of the ferroelastic transition between rutile-type and CaCl2-type SiO2 at high pressures. Phys. Rev. B 2008, 78, 134106. [Google Scholar] [CrossRef]
  40. Zhang, J.; Jia, S.; Kholmanov, I.; Dong, L.; Er, D.; Chen, W.; Guo, H.; Jin, Z.; Shenoy, V.B.; Shi, L.; et al. Janus Monolayer Transition-Metal Dichalcogenides. ACS Nano 2017, 11, 8192–8198. [Google Scholar] [CrossRef]
  41. Andrew, R.C.; Mapasha, R.E.; Ukpong, A.M.; Chetty, N. Mechanical properties of graphene and boronitrene. Phys. Rev. B 2012, 85, 125428. [Google Scholar] [CrossRef]
  42. Goodenough, B. Theory of the role of covalence in the Perovskite-type manganites [La, M(II)] MnO3. Phys. Rev. 1955, 100, 564. [Google Scholar] [CrossRef]
  43. Kanamori, J. Superexchange interaction and symmetry properties of electron orbitals. J. Phys. Chem. Solids 1959, 10, 87–98. [Google Scholar] [CrossRef]
  44. Anderson, P.W. New approach to the theory of superexchange interactions. Phys. Rev. 1959, 115, 2. [Google Scholar] [CrossRef]
  45. Mermin, N.D.; Wagner, H. Absence of ferromagnetism or antiferromagnetism in one-or two-dimensionalisotropic Heisenberg models. Phys. Rev. Lett. 1966, 17, 1133. [Google Scholar] [CrossRef]
  46. Gong, C.; Li, L.; Li, Z.; Ji, H.; Stern, A.; Xia, Y.; Cao, T.; Bao, W.; Wang, C.; Wang, Y. Discovery of intrinsic ferromagnetism in two-dimensional van der Waals crystals. Nature 2017, 546, 265–269. [Google Scholar] [CrossRef]
  47. 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]
  48. Thouless, D.J.; Kohmoto, M.; Nightingale, M.P.; den Nijs, M. Quantized Hall Conductance in a Two-Dimensional Periodic Potential. Phys. Rev. Lett. 1982, 49, 405. [Google Scholar] [CrossRef]
  49. Xiao, D.; Chang, M.C.; Niu, Q. Berry phase effects on electronic properties. Rev. Mod. Phys. 2010, 82, 1959. [Google Scholar] [CrossRef]
Figure 1. The (a) top and (b) side views of monolayer TiAlTe3 along with the spin charge density. The blue, green, and red balls represent Te, Al, and Ti atoms, respectively. (c) The phonon dispersion of monolayer TiAlTe3. (d) The AIMD of monolayer TiAlTe3 at 300 K. (e) The splitting of Ti-3d orbitals under trigonal prismatic crystal field.
Figure 1. The (a) top and (b) side views of monolayer TiAlTe3 along with the spin charge density. The blue, green, and red balls represent Te, Al, and Ti atoms, respectively. (c) The phonon dispersion of monolayer TiAlTe3. (d) The AIMD of monolayer TiAlTe3 at 300 K. (e) The splitting of Ti-3d orbitals under trigonal prismatic crystal field.
Materials 18 02396 g001
Figure 2. In the FM ground state, the band structures of monolayer TiAlTe3 (a) without SOC and (b,c) with SOC for magnetization in (b) positive z and (c) negative z directions, respectively. In (a), the red and blue lines represent spin-up and spin-down states, respectively.
Figure 2. In the FM ground state, the band structures of monolayer TiAlTe3 (a) without SOC and (b,c) with SOC for magnetization in (b) positive z and (c) negative z directions, respectively. In (a), the red and blue lines represent spin-up and spin-down states, respectively.
Materials 18 02396 g002
Figure 3. In the FM ground state, the orbital-projected band structure of monolayer TiAlTe3 with SOC for magnetization in the positive z direction. The red, green, and blue symbols denote the dz2, dyz/dxz, and dx2y2/dxy orbital components of a Ti atom, respectively.
Figure 3. In the FM ground state, the orbital-projected band structure of monolayer TiAlTe3 with SOC for magnetization in the positive z direction. The red, green, and blue symbols denote the dz2, dyz/dxz, and dx2y2/dxy orbital components of a Ti atom, respectively.
Materials 18 02396 g003
Figure 4. The Berry curvature of monolayer TiAlTe3 in the FM ground state when the magnetization orientation is OP (a). (b) Schematic diagram of AVHE for hole-doped monolayer TiAlTe3.
Figure 4. The Berry curvature of monolayer TiAlTe3 in the FM ground state when the magnetization orientation is OP (a). (b) Schematic diagram of AVHE for hole-doped monolayer TiAlTe3.
Materials 18 02396 g004
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

Jia, K.; Zhang, C.-W.; Wang, Z.-R.; Wang, P.-J. Monolayer TiAlTe3: A Perfect Room-Temperature Valleytronic Semiconductor. Materials 2025, 18, 2396. https://doi.org/10.3390/ma18102396

AMA Style

Jia K, Zhang C-W, Wang Z-R, Wang P-J. Monolayer TiAlTe3: A Perfect Room-Temperature Valleytronic Semiconductor. Materials. 2025; 18(10):2396. https://doi.org/10.3390/ma18102396

Chicago/Turabian Style

Jia, Kang, Chang-Wen Zhang, Zi-Ran Wang, and Pei-Ji Wang. 2025. "Monolayer TiAlTe3: A Perfect Room-Temperature Valleytronic Semiconductor" Materials 18, no. 10: 2396. https://doi.org/10.3390/ma18102396

APA Style

Jia, K., Zhang, C.-W., Wang, Z.-R., & Wang, P.-J. (2025). Monolayer TiAlTe3: A Perfect Room-Temperature Valleytronic Semiconductor. Materials, 18(10), 2396. https://doi.org/10.3390/ma18102396

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