Giant spin Seebeck effect in two-dimensional V2S2O altermagnet
Abstract
Altermagnet is an uncommon category of antiferromagnets distinguished by their non-overlapping spin-bands, drawing significant attention from researchers. However, while reports on their electronic and magnetic properties are increasing rapidly, the study on the transport properties is still in early stage. Therefore, we explored the orientational spin-dependent transport features of altermagnet V2S2O using Boltzmann transport technique. This altermagnet had 1.15 eV direct band gap energy and a critical temperature of 746 K. We obtained a directional spin-dependent feature in the band structure whose effect spans through all the spin-dependent transport parameters. We found a low isotropic lattice thermal conductivity of magnitude 0.2 Wm-1K-1 at 300 K. Above all, the V2S2O altermagnet displayed a giant spin-dependent Seebeck coefficient of about 1.8 mVK-1 at 300 K and at a small electron or hole doping. This value is multiple times greater than reported values for most transport materials. Besides, we also found a maximum figure of merit of 0.86 in the hole-doped systems. Thus, our findings suggest the possibility of pure spin current generation for possible applications in spintronics and thermoelectricity.
Keywords
INTRODUCTION
Altermagnetic (AM) materials are symmetry-compensated antiferromagnets whose band structures are characterized by spin splitting[1]. This implies that over the entire Brillouin zone, the up and down spin bands of the AM materials do not entirely overlap. Rather, the band structure exhibits
In parallel with the discovery of the AM behavior in RuO2 and other bulk materials, it has been predicted that two-dimensional (2D) materials such as Mn2ClI, V2S2O, MnTeMoO6, V2F7Cl, Cr2SO, V2SeTeO, Cr2O2, CrMoC2S6, Mn2ClF, RuF4, and Cr2SeO can also display an AM behavior[4-6,17,18]. However, most studies have concentrated on basic unconventional properties with minimal reports on thermoelectricity in AM materials[19-24]. Recently, Sukhachov et al. suggested that AM materials could be employed for effective thermoelectricity[25]. Using the AM spin-splitting effect, Bai et al. also reported efficient spin-charge conversion in RuO2 altermagnet[26]. Besides, Lyu and Li pointed out that the transport properties in AM materials were directional dependent[27]. Indeed, the AM behavior has been experimentally confirmed in RuO2 and MnTe using angle-resolved photoemission spectroscopy[28,29]. Fan et al. also reported experimental work that the AM spin splitting effect could produce anisotropic spin currents in AM material with polarization, which depends on the crystal orientation of the material[30]. Moreover. Bai et al. also reported a few microvolts in altermagnet RuO2 spin-charge conversion efficiency using spin Seebeck effect measurement[26]. Motivated by the predicted potentials of unconventional spin-splitting effect in AM materials, we propose to leverage directional spin-dependent thermoelectric properties and predict a massive spin-Seebeck coefficient in V2S2O monolayer AM material.
Indeed, thermoelectricity is a fertile field for testing and predicting high-performance transport materials via theory-based material search techniques. Although precise prediction of these high-performance transport materials computationally entails vigorous optimization, the experimental approach is even more complicated. In our work, we inspect the orientational spin-dependent transport properties at finite temperatures and aim to propose a giant spin Seebeck effect in the V2S2O altermagnet for the spin-polarized current generation in spintronics and thermoelectric device applications.
THEORETICAL METHODS
All our calculations were conducted in this work using the spin-polarized density functional theory as implemented in the Vienna ab initio simulation package (VASP)[31,32]. We used the Perdew, Burke, and Ernzerhof parametrization within the generalized gradient approximation for an exchange functional[33], including the van der Waals interaction[34] as a correlation functional. A plane wave basis set is employed with an energy cutoff of 600 eV. The energy and force convergence criteria are set to 10-6 eV and 10-3 eV/Å, respectively. A well-converged Monkhorst-Pack scheme-generated k-point mesh of 10 × 10 × 1, which also ensures the convergence of the energy and force, was used[35]. Furthermore, we applied a vacuum distance of 25 Å along the z-axis to prevent adjacent units’ interaction. The Heyd-Scuseria-Ernzerhof screened hybrid functional was used to calculate the spin-polarized electronic band structure[36,37]. To estimate the critical temperature, we applied a simulation of the field cooling and Landau-Lifshitz-Gilbert Huen method as implemented in the VAMPIRE simulation package[38] for the temperature-dependent sub-lattice magnetization calculation. To this end, a supercell of 100 × 100 × 1 with total time steps of 100,000 and 0.1 fs is used. Our directional spin-dependent transport properties were explored using the BoltzTraP code[39]. Since most of the transport properties are carrier relaxation time (τ) dependent, we considered the scattering contributions from the impurity effects (τimp) polar optical phonon (τopt), and acoustic phonon
In addition, we also computed the lattice part of the thermal conductivity with phono3py code[40] using a
RESULTS AND DISCUSSION
Figure 1A reveals the side and top views of monolayer V2S2O. The red and black colors denote vanadium atoms with spin-up and spin-down, while the blue and yellow colors represent the oxygen and sulfur atoms. The monolayer V2S2O has a tetragonal crystal structure and a space group of P4mmm (123). We obtained the lattice constants of a = b = 3.84 Å after structure optimization. The monolayer V2S2O is made of a V2O plane embedded by two S planes with interatomic distances of 1.921 Å and 2.479 Å from V to O and S. This shows that the V2S2O monolayer exhibits high-degree symmetry along in-plane and out of plane. Note that the thermal stability, cleavage energy, and dynamic stability were already reported elsewhere[41]. To investigate the magnetic ground state of monolayer V2S2O, we considered ferromagnetic (FM) and three different antiferromagnetic (AFM) spin configurations in a supercell of dimension 2 × 2 × 1. In Figure 1B-D, we show the different AFM spin configurations. The most stable AFM-Neel state was obtained [Figure 1B], and the energy difference (Eex) between this state and the FM state (Eex = EFM - EAFM) was 536 meV/unit cell. Each V atom had an atomic magnetic moment of magnitude 1.7 μB. However, the net magnetic moment of the unit cell vanishes due to the anti-parallel coupling between the V atoms. The AFM-Neel ground state in the V2S2O monolayer is consistent with other theoretical reports[6,41]. Figure 1E reveals the electronic band structure without SOC. This electronic band structure shows a spin splitting around the X and Y points, while in other parts of the Brillouin zone, both spin-down (↓) and spin-up (↑) bands were completely overlapped. The spin splitting at X and Y points was 0.51 eV at the valence band maximum. AM V2S2O monolayer also showed a semiconducting behavior with a direct band gap of 1.15 eV at the Y and X points. We also computed the critical (Neel) temperature via the model of Heisenberg spin Hamiltonian[42] based on:
Figure 1. (A) Schematic illustration of top and side views of V2S2O monolayer. Magnetic configurations of (B) AFM-Neel, (C)
Here, Si and Sj are the spin moment direction of atoms at the neighboring sites i and j, Jij is the exchange interaction between atoms at the neighboring sites, and ku is the anisotropy energy per atom. Meanwhile, Nα and nα represent the number of atoms in the sublattice and the mean of sublattice magnetization of each atom. Using Eex, the difference between the total energies of the FM and AFM spin states, we extracted the exchange interaction via the relation:
where m and N are the average magnetic moment and the number of magnetic atoms in the system. Figure 1F shows the temperature-dependent sublattice magnetization curve. Overall, we obtained a Neel temperature of 746 K for the V2S2O monolayer.
Now, we explore the spin-dependent transport properties such as spin-dependent Seebeck coefficients, electrical conductivity, electronic thermal conductivity (Ke), and lattice thermal conductivity (KL). First, the spin-dependent Seebeck coefficient [S(↑↓)] is obtained using:
Here, q, T, and L are the elementary charge, the temperature, and the generalized linear spin-dependent transport coefficient expressed as:
The conductivity tensor over the bands (i) and the k-points (k) can be expressed as σ(↑↓)(i,k) = e2τi,kV2, where V represents group velocity. Figure 2 reveals the spin-dependent Seebeck coefficients for the spin-up channel (S↑) and spin-down channel (S↓). In the x-direction, the Seebeck coefficient of the spin-down channel comes first before the spin-up channel Seebeck coefficient in both doped systems. Meanwhile, in the y-direction, the Seebeck coefficient of the spin-up channel appears first before the spin-down component in both the electron and hole-doped systems. This feature is the direct consequence of the directional spin-dependent band structure in Figure 1E.
Figure 2. Spin-dependent Seebeck coefficient (S↑ and S↓) along x-direction at (A) 100 K vs. chemical potentials; (B) 200 K vs. chemical potentials; (C) 300 K vs. chemical potentials; and y-direction at (D) 100 K vs. chemical potentials; (E) 200 K vs. chemical potentials; and (F) 300 K vs. chemical potentials.
Note that the spin Seebeck coefficient does not depend on τ. However, other transport properties, such as electronic thermal and electrical conductivities, strongly depend on it. Hence, it is crucial to estimate τ accurately. Nonetheless, in most studies, the energy-independent constant relaxation time approach has been adopted with limitations in accounting for some scattering effects. This usually results in an overestimation of the thermoelectric performance. To remedy this, in our study, we applied the temperature-energy dependent relaxation time considering the contributions from the acoustic and optical phonons as implemented in the works of Casu et al. and Marfoua et al.[43,44]. Besides, we also considered the impurity scattering effect using the relation of Brooks-Herring. Note that the impurity scattering effect can be written as:
where the ZI, nI, ϵs, ϵ0 and m represent the impurity charge, the ionized impurity concentration, the relative dielectric constant, the vacuum’s permittivity, and the effective mass, respectively. Also,
where the sum is over all longitudinal optical polar phonons with energy
where D, ρ, v, E, and m are the band energy deformation potential, mass density, the sound average velocity, the energy of the electron, and the effective mass, respectively. To estimate the energy and temperature dependent τ, the intrinsic material quantities such as the dominant longitudinal optical phonon frequency
The calculated intrinsic parameters for hole and electron carriers in the spin-up channel for the energy and temperature-dependent relaxation time
Orientation | Carrier type | DP | m*/m | V (ms-1) | ε∞ | εL | Ρ (kgm-3) | ħωLO (meV) |
x | Hole | 2.59 | 0.14 | 7752 | 1.89 | 3.67 | 3410 | 70 |
Electron | 1.09 | 0.13 | 7752 | 1.89 | 3.67 | 3410 | 70 | |
y | Hole | 3.05 | 0.09 | 7752 | 1.89 | 3.67 | 3410 | 70 |
Electron | 1.69 | 0.12 | 7752 | 1.89 | 3.67 | 3410 | 70 |
The calculated intrinsic parameters for the energy-dependent relaxation time in the spin-down channel for hole and electron carriers
Orientation | Carrier type | DP | m*/m | V (ms-1) | ε ∞ | εL | Ρ (kgm-3) | ħωLO (meV) |
x | Hole | 3 | 0.09 | 7752 | 1.89 | 3.67 | 3410 | 70 |
Electron | 1.64 | 0.11 | 7752 | 1.89 | 3.67 | 3410 | 70 | |
y | Hole | 2.64 | 0.15 | 7752 | 1.89 | 3.67 | 3410 | 70 |
Electron | 1.14 | 0.16 | 7752 | 1.89 | 3.67 | 3410 | 70 |
We present the temperature and spin-dependent carrier relaxation time along the y-direction (τy) and
Figure 3A and B reveals the temperature and spin-dependent electrical conductivity as a function of chemical potential. In the x-direction, the electrical conductivity was entirely influenced by the carriers of spin-down (blue color) because the contribution from the spin-up channel carriers (red color) was negligible. Meanwhile, in the y-direction, the opposite behavior was found. This behavior originates from the electronic band structure where the spin-down (spin-up) channel controls the bandgap in the
Figure 3. Spin-dependent electrical conductivities along (A) x-direction and (B) y-direction at different temperatures.
Next, we calculate the total effective spin-dependent Seebeck coefficient (Seff-spin) and effective
Figure 4A-D reveals the computed results. Along the x-direction (y-direction), the Seff-spin and Seff-charge are determined by the spin-down carriers (spin-up carrier) since σ↑(σ↓) is negligible. Due to this feature, we expect the generation of a pure spin-polarized current in a particular direction when a heat gradient is applied along that direction. At 300 K, we obtained a maximum Seff-spin of 1.8 mV/K in both electron and hole-doped systems along the x and y directions. Note that both Seff-spin and Seff-charge have approximately the same magnitude at all temperatures along the x and y directions. This feature results from the negligible value of either σ↑ or σ↓ along each direction as discussed earlier. Indeed, we achieved a giant effective Seebeck coefficient in the V2S2O monolayer altermagnet compared with previously reported structures. For instance, the maximum spin Seebeck coefficients reported for GaMnAs, Pt/YIG, Cr2O3, HfMnGe, EuTiO3 and V2SeTeO monolayers are ~5 μV/K, ~6 μV/K, 20 μV/K, ~80 μV/K, 0.1 mV/K, and 0.64 mV/K respectively at room temperature[45-49].
Figure 4. Effective spin Seebeck coefficients at different temperatures in (A) x-direction vs. chemical potentials and in (B) y-direction vs. chemical potentials, and effective charge Seebeck coefficients in (C) x-direction vs. chemical potentials, and in (D) y-direction vs. chemical potentials.
We also investigated the electronic and lattice thermal conductivities. Figure 5A and B shows the Ke for the spin-up (red color) and spin-down (blue color) carriers along the x-direction and y-direction as a function of temperature and chemical potential. Similar to the σ, the electronic part of the thermal conductivity is also entirely determined by either of the spin carriers depending on the direction under consideration. To be precise, the spin-up (spin-down) carriers were wholly responsible for the Ke in the y-direction
Figure 5. Electronic thermal conductivity along (A) x-direction and (B) y-direction for spin-up and spin-down at different temperatures (red color for spin-up and blue color for spin-down); (C) The lattice thermal conductivity vs. temperature along x-direction (color black) and y-direction (color red).
The KL is given by:
where τi, vi, and Ci are the phonon lifetime, group velocity, and mode heat capacity. Besides, the phonon mode, number of unit cells, and volume of the system are represented with i, N, and V. Figure 5C shows the lattice part of the thermal conductivity of the V2S2O monolayer along the x and y directions. The KL was decreased with increasing temperature. This behavior agrees with the evolution law of KL. For instance, at 100 K we obtained an isotropic KL of 0.5 Wm-1K-1 in both directions. This value decreased to 0.2 Wm-1K-1 at 300 K. The low KL at 300 K may be an indication of relatively stronger phonon-phonon interaction leading to substantial anharmonic scattering. Due to this stronger scattering, the phonon mean free path will be decreased, and this may result in low KL. Overall, we found that the total thermal conductivity of the system was mostly governed by the Ke. For comparison, we present the total thermal conductivity of the V2S2O monolayer with other well-studied MoS2 in Supplementary Table 1.Finally, we evaluated the spin and charge dimensionless figure of merit (ZT). We obtained the dimensionless ZT using the transport parameters in the relations:
where σ = σ↑ + σ↓, and Ke is the sum of the electronic part of the thermal conductivities of the spin-down and spin-up carriers. Figure 6A-D reveals the computed results. Due to the directional spin-dependence of the σ, we also obtained ZTCharge = ZTspin along the y and x directions. At 300 K, we obtained a maximum ZT of 0.86 in the hole-doped systems, while it became 0.63 in the electron-doped systems.
Figure 6. The dimensionless effective spin figure of merit in (A) x and (B) y-directions and the dimensionless charge figure of merit in (C) x and (D) y-directions vs. chemical potentials.
This carrier type dependent ZT feature comes from the electrical conductivity in which σ↑(σ↓) is greater in the hole-doped system along the y-direction (x-direction). Overall, the ZT is controlled by only one spin component in each direction. Our findings may suggest that the V2S2O altermagnet system may be used for potential pure spin current generation in spintronic and thermoelectricity.
CONCLUSIONS
We probed the spin-dependent transport properties of the altermagnet V2S2O monolayer using the
DECLARATIONS
Authors’ contributions
Conception of the idea and design of the study, data curation, formal analysis, writing an original draft, and revising the manuscript: Hong, J.
Performed DFT calculations and data curation, analyzed and interpreted the results, and wrote and revised the manuscript: Ashani, T.M.; Abdullah.
Availability of data and materials
The data that support the findings of this study are available from the corresponding author upon reasonable request.
Financial support and sponsorship
This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korean government (MSIT) (2022R1A2C1004440) and by the National Research Foundation of Korea (2022M3H4A1A040853)
Conflicts of interest
All authors declared that there are no conflicts of interest
Ethical approval and consent to participate
Not applicable.
Consent for publication
Not applicable.
Copyright
© The Author(s) 2025.
Supplementary Materials
REFERENCES
1. I; The PRX Editors. Editorial: altermagnetism-a new punch line of fundamental magnetism. Phys. Rev. X. 2022, 12, 04002.
2. Manchon, A.; Koo, H. C.; Nitta, J.; Frolov, S. M.; Duine, R. A. New perspectives for Rashba spin-orbit coupling. Nat. Mater. 2015, 14, 871-82.
3. Zhang, F.; Lv, L.; Xu, Z.; et al. Prediction of the TiS2 bilayer with self-intercalation: robust ferromagnetic semiconductor with a high curie temperature. J. Phys. Chem. C. 2025, 129, 5577-88.
4. Guo, S. D.; Tao, Y. L.; Wang, G.; Ang, Y. S. How to produce spin-splitting in antiferromagnetic materials. J. Phys. Condens. Matter. 2024, 36, 215804.
5. Qi, Y.; Zhao, J.; Zeng, H. Spin-layer coupling in two-dimensional altermagnetic bilayers with tunable spin and valley splitting properties. Phys. Rev. B. 2024, 110, 014442.
6. Ma, H. Y.; Hu, M.; Li, N.; et al. Multifunctional antiferromagnetic materials with giant piezomagnetism and noncollinear spin current. Nat. Commun. 2021, 12, 2846.
7. Guo, Y.; Liu, H.; Janson, O.; Fulga, I. C.; van, den., Brink., J.; Facio, J. I. Spin-split collinear antiferromagnets: a large-scale ab-initio study. Materials. Today. Physics. 2023, 32, 100991.
8. Šmejkal, L.; Sinova, J.; Jungwirth, T. Beyond conventional ferromagnetism and antiferromagnetism: a phase with nonrelativistic spin and crystal rotation symmetry. Phys. Rev. X. 2022, 12, 031042.
9. Šmejkal, L.; Sinova, J.; Jungwirth, T. Emerging research landscape of altermagnetism. Phys. Rev. X. 2022, 12, 040501.
10. Naka, M.; Hayami, S.; Kusunose, H.; Yanagi, Y.; Motome, Y.; Seo, H. Spin current generation in organic antiferromagnets. Nat. Commun. 2019, 10, 4305.
11. Zhu, Y.; Chen, T.; Li, Y.; et al. Multipiezo effect in altermagnetic V2SeTeO monolayer. Nano. Lett. 2024, 24, 472-8.
12. Bhowal, S.; Spaldin, N. A. Ferroically ordered magnetic octupoles in d-wave altermagnets. Phys. Rev. X. 2024, 14, 011019.
13. Guo, S.; Guo, X.; Cheng, K.; Wang, K.; Ang, Y. S. Piezoelectric altermagnetism and spin-valley polarization in Janus monolayer Cr2SO. Appl. Phys. Lett. 2023, 123, 082401.
14. Gao, Z. F.; Qu, S.; Zeng, B.; et al. AI-accelerated discovery of altermagnetic materials. arXiv 2023, arXiv:2311.04418. Available online: http://arxiv.org/abs/2311.04418 (accessed 2025-05-14).
15. Mazin, I. I. Altermagnetism in MnTe: origin, predicted manifestations, and routes to detwinning. Phys. Rev. B. 2023, 107, L100418.
16. Mazin, I. I.; Koepernik, K.; Johannes, M. D.; González-Hernández, R.; Šmejkal, L. Prediction of unconventional magnetism in doped FeSb2. Proc. Natl. Acad. Sci. U. S. A. 2021, 118, e2108924118.
17. Sødequist, J.; Olsen, T. Two-dimensional altermagnets from high throughput computational screening: symmetry requirements, chiral magnons, and spin-orbit effects. Appl. Phys. Lett. 2024, 124, 182409.
18. Zeng, S.; Zhao, Y. Description of two-dimensional altermagnetism: categorization using spin group theory. Phys. Rev. B. 2024, 110, .054406.
19. Occhialini, C. A.; Martins, L. G. P.; Fan, S.; et al. Strain-modulated anisotropic electronic structure in superconducting RuO2 films. Phys. Rev. Materials. 2022, 6, 084802.
20. Gonzalez, Betancourt., R. D.; Zubáč, J.; Gonzalez-Hernandez, R.; et al. Spontaneous anomalous hall effect arising from an unconventional compensated magnetic phase in a semiconductor. Phys. Rev. Lett. 2023, 130, 036702.
21. Šmejkal, L.; Hellenes, A. B.; González-hernández, R.; Sinova, J.; Jungwirth, T. Giant and tunneling magnetoresistance in unconventional collinear antiferromagnets with nonrelativistic spin-momentum coupling. Phys. Rev. X. 2022, 12, 011028.
22. Ahn, K.; Hariki, A.; Lee, K.; Kuneš, J. Antiferromagnetism in RuO2 as d-wave pomeranchuk instability. Phys. Rev. B. 2019, 99, 184432.
23. Hayami, S.; Yanagi, Y.; Kusunose, H. Momentum-dependent spin splitting by collinear antiferromagnetic ordering. J. Phys. Soc. Jpn. 2019, 88, 123702.
24. Yuan, L.; Wang, Z.; Luo, J.; Rashba, E. I.; Zunger, A. Giant momentum-dependent spin splitting in centrosymmetric low-Z antiferromagnets. Phys. Rev. B. 2020, 102, 014422.
25. Sukhachov, P. O.; Hodt, E. W.; Linder, J. Thermoelectric effect in altermagnet-superconductor junctions. Phys. Rev. B. 2024, 110, 094508.
26. Bai, H.; Zhang, Y. C.; Zhou, Y. J.; et al. Efficient spin-to-charge conversion via altermagnetic spin splitting effect in antiferromagnet RuO2. Phys. Rev. Lett. 2023, 130, 216701.
27. Lyu, K.; Li, Y. Orientation-dependent spin-polarization and transport properties in altermagnet based resonant tunneling junctions. Results. Phys. 2024, 59, 107564.
28. Fedchenko, O.; Minár, J.; Akashdeep, A.; et al. Observation of time-reversal symmetry breaking in the band structure of altermagnetic RuO2. Sci. Adv. 2024, 10, eadj4883.
29. Krempaský, J.; Šmejkal, L.; D'Souza, S. W.; et al. Altermagnetic lifting of Kramers spin degeneracy. Nature 2024, 626, 517-22.
30. Fan, Y.; Wang, Q.; Wang, W.; et al. Robust magnetic-field-free perpendicular magnetization switching by manipulating spin polarization direction in RuO2/[Pt/Co/Pt] heterojunctions. ACS. Nano. 2024, 18, 26350-8.
31. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15-50.
32. 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.
33. Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865-8.
34. Dion, M.; Rydberg, H.; Schröder, E.; Langreth, D. C.; Lundqvist, B. I. Van der Waals density functional for general geometries. Phys. Rev. Lett. 2004, 92, 246401.
35. Pack, J. D.; Monkhorst, H. J. “Special points for Brillouin-zone integrations”-a reply. Phys. Rev. B. 1977, 16, 1748.
36. Heyd, J.; Scuseria, G. E.; Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2003, 118, 8207-15.
37. Heyd, J.; Scuseria, G. E.; Ernzerhof, M. Erratum: “Hybrid functionals based on a screened Coulomb potential” [J. J. Chem. Phys. 2006, 124, 219906.
38. Vampire Curie temperature simulation tutorial. https://vampire.york.ac.uk/tutorials/simulation/curie-temperature/ (accessed 2025-05-14).
39. Madsen, G. K.; Singh, D. J. BoltzTraP. A code for calculating band-structure dependent quantities. Comput. Phys. Commun. 2006, 175, 67-71.
40. Togo, A.; Chaput, L.; Tanaka, I. Distributions of phonon lifetimes in Brillouin zones. Phys. Rev. B. 2015, 91, 094306.
41. Yu, Y. High storage capacity and small volume change of potassium-intercalation into novel vanadium oxychalcogenide monolayers V2S2O, V2Se2O and V2Te2O: an ab initio DFT investigation. Appl. Surf. Sci. 2021, 546, 149062.
42. 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.
43. Casu, G.; Bosin, A.; Fiorentini, V. Efficient thermoelectricity in Sr2Nb2O7 with energy-dependent relaxation times. Phys. Rev. Materials. 2020, 4, 075404.
44. Marfoua, B.; Hong, J. Graphene induced high thermoelectric performance in ZnO/graphene heterostructure. Adv. Mater. Interfaces. 2023, 10, 2202387.
45. Ullah, A.; Bezzerga, D.; Hong, J. Giant spin seebeck effect with highly polarized spin current generation and piezoelectricity in flexible V2SeTeO altermagnet at room temperature. Mater. Today. Phys. 2024, 47, 101539.
46. Jaworski, C. M.; Yang, J.; Mack, S.; Awschalom, D. D.; Heremans, J. P.; Myers, R. C. Observation of the spin-Seebeck effect in a ferromagnetic semiconductor. Nat. Mater. 2010, 9, 898-903.
47. Reitz, D.; Li, J.; Yuan, W.; Shi, J.; Tserkovnyak, Y. Spin Seebeck effect near the antiferromagnetic spin-flop transition. Phys. Rev. B. 2020, 102, 020408.
48. Wadhwa, P.; Bosin, A.; Filippetti, A. Giant spin-dependent Seebeck effect from fully spin-polarized carriers in n-doped EuTiO3: a prototype material for spin-caloritronic applications. J. Mater. Chem. A. 2023, 11, 6842-53.
Cite This Article

How to Cite
Download Citation
Export Citation File:
Type of Import
Tips on Downloading Citation
Citation Manager File Format
Type of Import
Direct Import: When the Direct Import option is selected (the default state), a dialogue box will give you the option to Save or Open the downloaded citation data. Choosing Open will either launch your citation manager or give you a choice of applications with which to use the metadata. The Save option saves the file locally for later use.
Indirect Import: When the Indirect Import option is selected, the metadata is displayed and may be copied and pasted as needed.
About This Article
Special Issue
Copyright
Data & Comments
Data

Comments
Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at [email protected].