Next Article in Journal
Memristor-Based Spiking Neuromorphic Systems Toward Brain-Inspired Perception and Computing
Previous Article in Journal
Controlled Synthesis of Tellurium Nanowires and Performance Optimization of Thin-Film Transistors via Percolation Network Engineering
Previous Article in Special Issue
Structural, Magnetic and THz Emission Properties of Ultrathin Fe/L10-FePt/Pt Heterostructures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigation of Thermoelectric Properties in Altermagnet RuO2

1
Hubei Province Key Laboratory of Systems Science in Metallurgical Process, The State Key Laboratory for Refractories and Metallurgy, Collaborative Innovation Center for Advanced Steels, International Research Institute for Steel Technology, Wuhan University of Science and Technology, Wuhan 430081, China
2
School of Mathematics and Physics Science and Engineering, Hebei University of Engineering, Handan 056038, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2025, 15(14), 1129; https://doi.org/10.3390/nano15141129
Submission received: 31 May 2025 / Revised: 7 July 2025 / Accepted: 18 July 2025 / Published: 21 July 2025

Abstract

An altermagnet, characterized by its distinctive magnetic properties, may hold potential applications in diverse fields such as magnetic materials, spintronics, data storage, and quantum computing. As a prototypical altermagnet, RuO2 exhibits spin polarization and demonstrates the advantageous characteristics of high electrical conductivity and low thermal conductivity. These exceptional properties endow it with considerable promise in the emerging field of thermal spintronics. We studied the electronic structure and thermoelectric properties of RuO2; the constructed RuO2/TiO2/RuO2 all-antiferromagnetic tunnel junction (AFMTJ) exhibited thermally induced magnetoresistance (TIMR), reaching a maximum TIMR of 1756% at a temperature gradient of 5 K. Compared with prior studies on RuO2-based antiferromagnetic tunnel junctions, the novelty of this work lies in the thermally induced magnetoresistance based on its superior thermoelectric properties. In parallel structures, the spin-down current dominates the transmission spectrum, whereas in antiparallel structures, the spin-up current governs the transmission spectrum, underscoring the spin-polarized thermal transport. In addition, thermoelectric efficiency emphasizes the potential of RuO2 to link antiferromagnetic robustness with ferromagnetic spin functionality. These findings promote the development of efficient spintronic devices and spin-based storage technology for waste heat recovery and emphasize the role of spin splitting in zero-magnetization systems.

1. Introduction

Spin caloritronics [1,2], a significant research direction in spintronics, investigates the interconnection between spin and thermal transport in materials, potentially enabling new thermoelectric conversion, waste heat recovery, and information processing technologies. The spin Seebeck effect (SSE), as one of the core topics in spin caloritronics, explores the thermally induced spin current and associated spin voltage in magnetic materials, offering a feasible mechanism to generate pure spin current [3,4,5,6]. Spin waves and conduction electrons are widely recognized as two primary carriers mediating spin-Seebeck currents [7]. Crucially, the efficiency of spin and thermal transport mediated by these carriers, particularly conduction electrons, is fundamentally governed by their scattering mechanisms. The electron scattering mechanism refers to the process in which electrons collide with other particles (such as other electrons, impurity atoms, lattice defects, etc.) while moving in a material, thereby changing their direction of motion and energy. In conductive materials, electrons are the main charge carriers, and they move under the influence of an electric field. However, due to the inhomogeneity within the material and the presence of other particles, the trajectory of electron motion is altered. Similarly, lattice vibrations play a key role. Phonons are the quantized representation of lattice vibrations. The phonon scattering mechanism, particularly relevant to electron–phonon interaction, refers to the process in which electrons interact with lattice vibrations (phonons) and change their state of motion. In solid materials, atoms vibrate slightly around their equilibrium positions, and these vibrations can propagate through the lattice. When electrons interact with these lattice vibrations, phonon scattering occurs. Nevertheless, the relationship between spin-dependent thermal transport and material symmetry persists as a critical challenge in developing high-efficiency spin-caloritronic devices [8]. Understanding these scattering mechanisms, their dependence on material structure, temperature, and magnetic order, is therefore paramount for unraveling the complexities of the SSE and addressing the symmetry-related challenge to achieve enhanced device performance.
Recently identified as a distinct category of antiferromagnetism, altermagnetism exhibits momentum-locked spin splitting and alternating spin polarizations in both real-space (crystalline structure) and momentum-space (band structure) [9,10]. This emergent phenomenon has garnered significant interest due to its exceptional properties [9,10,11,12,13,14,15,16]. Similarly to conventional antiferromagnetic (AFM) materials, altermagnets possess zero net magnetization; however, they manifest nonrelativistic spin splitting along high-symmetry band structure lines, enabling ferromagnetic (FM)-like behavior under specific conditions. This hybrid character establishes altermagnets as a link between AFM and FM systems, creating new avenues for spin manipulation in zero-magnetization environments [8]. Certain magnetic space groups (MSGs) with the broken TPτ and symmetries support this nonrelativistic spin splitting, where T, P, U, and τ denote the time reversal, spatial inversion, spinor symmetry, and the half lattice translation [12,13,14]. Given magnetism’s expansion through altermagnetism, integrating altermagnetic electrodes into magnetic tunnel junctions (MTJs) [17,18,19] represents a strategic progression toward achieving enhanced tunneling TMR effects and experimentally validating theoretical predictions.
MTJs comprising the FM/insulator barrier/FM sandwich structure [20,21,22] serve as the fundamental component of spintronic devices in applications such as nonvolatile magnetic random-access memories (MRAMs) and magnetic sensors [23,24]. By altering the magnetization orientation of the two FM electrodes between antiparallel and parallel configurations, an MTJ transitions between high- and low-resistance states—termed the TMR effect—which forms the operational basis of MTJ devices. Notwithstanding these advantages, conventional FM-based MTJs exhibit stray magnetic fields and constrained thermal stability, impeding their scalability and energy efficiency [25]. Conversely, AFM materials possess inherent benefits over FM materials: absence of stray magnetic fields, enhanced robustness against external fields, pronounced anisotropy, ultrafast spin dynamics, and appealing characteristics for AFM spintronics. Traditionally, however, the absence of global spin polarization prevented AFM materials from being utilized as MTJ electrodes. This longstanding perspective is now disrupted by altermagnetism, where momentum-dependent spin splitting facilitates localized spin currents without net magnetization [8].
Recent theoretical and experimental advances have confirmed that AFMTJs can produce the TMR effect when employing spin-splitting AFM electrodes, including RuO2 [16,26], Mn3Sn [27,28], and Mn3Pt [29]. Notably, RuO2 exhibits exceptional stability in acidic environments and tunable electronic structures, as evidenced by its recent applications in oxygen evolution reactions (OER) and solid-state catalysis [30,31], yet its potential in spin-dependent thermoelectric transport remains largely unexplored [32].
RuO2 demonstrates exceptional multifunctional properties, establishing it as a leading candidate for advanced spintronic and spin-caloritronic applications. Being a metallic altermagnet, RuO2 displays distinct spin-splitting in its electronic band structure, alongside high electrical conductivity [33] and ultralow thermal conductivity, making it an exceptional thermoelectric material. Research has confirmed the presence of a giant crystal Nernst effect and crystal thermal Hall effect within this material, which exhibit strong anisotropy concerning the Néel vector [34]. The significant crystal thermal transport stems mainly from three sources of Berry’s curvature in momentum space: the Weyl fermions due to crossings between well-separated bands, the strong spin-flip pseudo-nodal surfaces, and the weak spin-flip ladder transitions, defined by transitions among very weakly spin-split states of similar dispersion crossing the Fermi surface. Furthermore, it has been revealed that the anomalous thermal and electrical transport coefficients in RuO2 are linked by an extended Wiedemann–Franz law in a temperature range much wider than expected for conventional magnets. These findings suggest that altermagnets may play a leading role in realizing spin-caloritronic concepts unattainable with ferromagnets or antiferromagnets.
In this work, we systematically investigate the electronic structure, thermoelectric properties, and thermal spin transport characteristics of the altermagnet RuO2. Our results demonstrate that RuO2 exhibits distinct spin-splitting in its band structure, along with high electrical conductivity and low thermal conductivity. Through first-principle calculations, we further explore the spin current and thermal transport properties of RuO2/TiO2/RuO2 MTJs. Compared with prior studies on RuO2-based antiferromagnetic tunnel junctions, the novelty of this work lies in the thermally induced magnetoresistance based on its superior thermoelectric properties [26,35]. Our findings highlight the significant potential of altermagnetic RuO2 in applications related to thermal spintronics.

2. Theoretical Methods

First-principle calculations are performed based on the density functional theory (DFT) [36] as implemented in the Vienna ab initio simulation package (VASP) [37,38]. The pseudopotentials are described using the projector augmented wave (PAW) method [39], and the exchange-correlation functional is treated within the generalized gradient approximation (GGA) developed by Perdew, Burke, and Ernzerhof (PBE) [40]. The transport properties are computed using the non-equilibrium Green’s function formalism (DFT + NEGF approach) [41,42] and implemented in QuantumATK using its relaxed atomic structure. In QuantumATK, we used the nonrelativistic SG15 pseudopotentials [43], and k-point meshes of 17 × 17 × 25 for bulk RuO2 and TiO2 and 6 × 6 × 303 for RuO2/TiO2/RuO2 AFMTJ. The spin-polarized GGA + U [44,45] method with Ueff = 1.2 eV on Ru 4d [35] orbitals and Ueff = 5 eV on Ti 3d [35] orbitals is used in the calculations. These parameters have been well tested to ensure that the electronic structure around EF calculated by QuantumATK is consistent with that calculated by VASP.

3. Results and Discussion

AFMTJ consists of two infinite RuO2 electrodes sandwiching rutile structure TiO2, which exhibits a lattice mismatch of 1.7% relative to RuO2. Such a small lattice mismatch rate makes the structure of AFMTJ stable and conducive to optimization. The electronic and magnetic properties of all these structures are consistent with previous results [35], indicating that our current calculations are reliable. First, we relaxed the lattice constant. Figure 1 presents the optimized structural configurations of RuO2 and TiO2. The spin-dependent band structure and density of states (DOS) of RuO2 and TiO2 are shown in Figure 2. As shown in Figure 2a, RuO2 exhibits metallic characteristics. The spin degeneracy along the Г-X, Г-Z, X-M, Z-R, and R-A symmetry directions is resolved in the RuO2 band structure, while significant spin splitting emerges along the Г-M and Z-A directions.
RuO2 exhibits metallic characteristics with both conduction and valence bands crossing the Fermi level. In contrast, TiO2 possesses a substantial direct band gap of 2.4 eV, where both bands are positioned away from the Fermi level [46].
The expected thermoelectric parameters and ZT value for RuO2 are shown in Figure 3. In contrast to the declining electrical conductivity, which is suppressed by electron–phonon scattering, the thermal contributions from electrons and phonons grow as temperature increases. Notably, the room-temperature ZT reaches 0.01, suggesting RuO2’s potential as a promising thermoelectric material. The ZT value of RuO2 (0.01 at 300 K) is higher than that of the prototypical altermagnet MnTe2 (ZT = 0.008 at 300 K) [47], and its momentum-locked spin splitting enables superior spin-polarized thermal transport critical for device integration. This trade-off between thermoelectric efficiency and spin functionality highlights the material-specific optimization pathways within the altermagnetic family.
Next, we designed an AFMTJ using RuO2 (001) as electrodes and TiO2 (001) as an insulating barrier layer. Due to both RuO2 and TiO2 having a rutile structure and a similar lattice constant, this AFMTJ is experimentally feasible, as evidenced by the successful epitaxial growth of RuO2/TiO2 heterostructures via magnetron sputtering on TiO2 (001)-oriented substrates [48]. Figure 4c shows the atomic structure of the RuO2/TiO2/RuO2 (001) AFMTJ. The scattering region includes 8 TiO2 layers in the center and 10 RuO2 layers on each side, two infinitely extended RuO2 layers as electrodes. The rutile phase was selected for both RuO2 and TiO2 due to the following reasons: For RuO2, the rutile phase is its sole stable polymorph. In the case of TiO2, the rutile phase exhibits superior thermodynamic stability compared to other polymorphs. Crucially, when both oxides adopt the isostructural rutile configuration, their lattice mismatch is minimized, rendering this system particularly suitable for tunnel junction applications [49]. Parallel spin configuration (PC) and antiparallel spin configuration (APC) of the Ru atoms on the left and right electrodes are shown on top of Figure 4a,b.
The operational temperature range in this study is intrinsically limited by the Néel temperature of RuO2 (TN ≈ 400 K) [50]. Above the TN, RuO2 undergoes a magnetic phase transition from antiferromagnetic to paramagnetic states, accompanied by electronic structure modifications that eliminate spin polarization. This invalidates the premise of our model. Concurrently, the temperature dependence of TIMR becomes negligible at room temperature. Therefore, considering TN and this diminished temperature dependence, we restrict our calculations to the range of approximately 0 K to 400 K, without extending to higher temperatures. The current caused by the temperature difference (ΔT = TR − TL) between the temperature of the left electrode (TL) and the temperature of the right electrode (TR) is plotted in Figure 5. As shown in Figure 5a,c, the total current, the spin-up current, and the spin-down current versus TL at ΔT = 10 K, 20 K, 30 K, and 40 K for PC. It can be observed that a marked disparity emerges between the spin-up and the spin-down current, and the spin-down currents consistently exceed their spin-up currents with TL increasing to 260 K for different ΔT. The total current equals zero when TL reaches 260 K, where spin-up and spin-down electrons flow in different directions, contributing to a net spin current. For APC, the spin-up currents consistently exceed their spin-down currents with TL increasing for different ΔT, and there is no net spin current for different TL. The difference in the magnitude of the spin-polarized current between the PC and APC states correlates well with the difference in the magnitude of the transmission coefficient shown in Figure S3. Based on thermal current characteristics, the TIMR was derived as shown in Figure 5e, under different ΔT, TIMR values decrease with increasing TL. It is also very interesting to study the ΔT dependence of total current and spin current with TL fixed, shown in Figure 6. Here, we fix the left electrode at 30, 60, 90, and 120 K and set the ΔT at 0–100 K. Throughout the entire ΔT range, for PC, spin-down currents remain larger than their spin-up counterparts. For APC, spin-up currents play a dominant role in the total current. As shown in Figure 6e, TIMR values under different TL exhibit a platform with increasing ΔT. And when TL = 30 K, the maximum TIMR can reach 1756%. The spin polarization of altermagnetic materials is a key parameter for TIMR. Increasing temperature reduces spin polarization [51] while simultaneously enhancing thermal movement of electrons and increasing their scattering probability, consequently lowering TIMR. Altermagnets uniquely combine the advantages of both ferromagnets and antiferromagnets, offering a promising platform for advanced spintronic devices. Compared with conventional ferromagnetic MTJs, altermagnetic MTJs exhibit significantly reduced stray fields due to their compensated antiparallel spin structure, minimizing crosstalk in high-density integration. Simultaneously, they overcome the limitation of weak spin-dependent signals typically associated with antiferromagnetic MTJs by generating strong momentum-dependent spin splitting, enabling robust spin-polarized transport comparable to ferromagnetic systems. This synergy of low stray fields and robust spin-polarized transport endows altermagnetic MTJs with unique advantages in realizing high-density, high-speed, and high-stability spintronic devices.
The mechanism of this phenomenon described above is the interaction of the Fermi-Dirac distribution of the electrodes and the spin-polarized transmission spectra, as demonstrated in the Fermi-Dirac distribution figure (Figure 7). Remarkably, transmission spectra under both parallel (Figure 8) and antiparallel (Figure 9) Néel vector configurations exhibit spin-polarized transport. For PC, in comparison between spin-up and spin-down electron transmission coefficients, the spin-down electrons play the dominant role. For APC, the spin-up electrons exhibit a higher transmission probability, dominating the transport behavior.

4. Conclusions

In conclusion, we have studied the spin-dependent thermoelectric transport properties of variable magnetic RuO2 and its application in AFMTJ through first-principle calculations and quantum transport simulations. As a momentum-locked altermagnetic material, RuO2 exhibits zero net magnetization with nonrelativistic spin splitting in its band structure, enabling localized spin polarization critical for spin caloritronic applications. The results show that the ZT of RuO2 reaches 0.01 due to the influence of electron transport and the phonon scattering mechanism. Furthermore, the designed RuO2/TiO2/RuO2 AFMTJ demonstrates significant TIMR, reaching a maximum TIMR of 1756% at a temperature gradient of 5 K. In parallel structures, the spin-down current dominates the transmission spectrum, whereas in antiparallel structures, the spin-up current governs the transmission spectrum, underscoring the spin-polarized thermal transport. The small lattice mismatch (1.7%) between RuO2 and TiO2 ensures structural stability, while spin-polarized thermal transport highlights the interplay between altermagnetism and thermoelectric efficiency. Looking forward, the high TIMR efficiency of RuO2-based antiferromagnetic tunnel junctions (AFMTJs) offers a direct pathway for next-generation energy harvesting systems. This capability enables the conversion of ubiquitous low-grade waste heat gradients (<150 °C) into spin-polarized currents via the spin-dependent Seebeck effect. Such currents can directly drive ultra-low-power spintronic devices—including wireless sensor network nodes, biomedical implants, or non-volatile memory switching units—without intermediate power conversion stages. Crucially, RuO2’s room-temperature functionality and compatibility with CMOS processing could accelerate the development of self-powered, energy-autonomous electronics, leveraging waste thermal energy from industrial processes or microelectronic systems. These findings underscore the potential of altermagnetic materials in bridging the gap between antiferromagnetic robustness and ferromagnetic spin functionality, advancing high-efficiency, low-energy spintronic devices for waste heat recovery and spin-based memory technologies.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano15141129/s1, Figure S1. Spin-resolved HOMO and LUMO of RuO2/TiO2/RuO2 device under PC state. Figure S2. Spin-resolved HOMO and LUMO of RuO2/TiO2/RuO2 device under APC state. Figure S3. Transmission coefficient versus energy of RuO2/TiO2/RuO2 devices at (a) total, (b) along the z-direction, (c) spin-up and (d) spin-down.

Author Contributions

Conceptualization, J.L. and S.Z.; methodology, X.L.; software, C.N. and X.L.; validation, J.L.; formal analysis, J.L., S.Z. and S.W.; data curation, J.L.; writing—original draft preparation, J.L.; writing—review and editing, S.Z. 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 under Grants (No. 11704291, No. 12174296, and No. 52475210), the High-Performance Computing Center of Wuhan University of Science and Technology, and the Science Research Project of Hebei Education Department (No. QN2022034).

Data Availability Statement

The data presented in this study are available upon request from the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bauer, G.E.W.; Saitoh, E.; Van Wees, B.J. Spin caloritronics. Nat. Mater. 2012, 11, 391–399. [Google Scholar] [CrossRef] [PubMed]
  2. Goennenwein, S.T.B.; Bauer, G.E.W. Electron spins blow hot and cold. Nat. Nanotechnol. 2012, 7, 145–147. [Google Scholar] [CrossRef] [PubMed]
  3. Uchida, K.; Takahashi, S.; Harii, K.; Ieda, J.; Koshibae, W.; Ando, K.; Maekawa, S.; Saitoh, E. Observation of the spin Seebeck effect. Nature 2008, 455, 778–781. [Google Scholar] [CrossRef] [PubMed]
  4. 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. [Google Scholar] [CrossRef] [PubMed]
  5. Uchida, K.-I.; Adachi, H.; Ota, T.; Nakayama, H.; Maekawa, S.; Saitoh, E. Observation of longitudinal spin-Seebeck effect in magnetic insulators. Appl. Phys. Lett. 2010, 97, 172505. [Google Scholar] [CrossRef]
  6. Wu, S.M.; Pearson, J.E.; Bhattacharya, A. Paramagnetic spin Seebeck effect. Phys. Rev. Lett. 2015, 114, 186602. [Google Scholar] [CrossRef] [PubMed]
  7. Hirobe, D.; Sato, M.; Kawamata, T.; Shiomi, Y.; Uchida, K.-I.; Iguchi, R.; Koike, Y.; Maekawa, S.; Saitoh, E. One-dimensional spinon spin currents. Nat. Phys. 2017, 13, 30–34. [Google Scholar] [CrossRef]
  8. Eacret, J.S.; Gonzales, D.M.; Franks, R.G.; Burns, J.M. Immunization with merozoite surface protein 2 fused to a Plasmodium-specific carrier protein elicits strain-specific and strain-transcending, opsonizing antibody. Sci. Rep. 2019, 9, 9022. [Google Scholar] [CrossRef] [PubMed]
  9. 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. [Google Scholar] [CrossRef]
  10. Šmejkal, L.; Sinova, J.; Jungwirth, T. Emerging research landscape of altermagnetism. Phys. Rev. X 2022, 12, 040501. [Google Scholar] [CrossRef]
  11. Noda, Y.; Ohno, K.; Nakamura, S. Momentum-dependent band spin splitting in semiconducting MnO2: A density functional calculation. Phys. Chem. Chem. Phys. 2016, 18, 13294–13303. [Google Scholar] [CrossRef] [PubMed]
  12. Hayami, S.; Yanagi, Y.; Kusunose, H. Momentum-dependent spin splitting by collinear antiferromagnetic ordering. J. Phys. Soc. Jpn. 2019, 88, 123702. [Google Scholar] [CrossRef]
  13. Yuan, L.D.; Wang, Z.; Luo, J.W.; Rashba, E.I.; Zunger, A. Giant momentum-dependent spin splitting in centrosymmetric low-Z antiferromagnets. Phys. Rev. B 2020, 102, 014422. [Google Scholar] [CrossRef]
  14. Yuan, L.D.; Wang, Z.; Luo, J.W.; Zunger, A. Prediction of low-Z collinear and noncollinear antiferromagnetic compounds having momentum-dependent spin splitting even without spin-orbit coupling. Phys. Rev. Mater. 2021, 5, 014409. [Google Scholar] [CrossRef]
  15. Šmejkal, L.; González-Hernández, R.; Jungwirth, T.; Sinova, J. Crystal time-reversal symmetry breaking and spontaneous Hall effect in collinear antiferromagnets. Sci. Adv. 2020, 6, eaaz8809. [Google Scholar] [CrossRef] [PubMed]
  16. 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. [Google Scholar] [CrossRef]
  17. Wolf, S.A.; Awschalom, D.D.; Buhrman, R.A.; Daughton, J.M.; von Molnár, 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] [PubMed]
  18. Žutić, I.; Fabian, J.; Sarma, S.D. Spintronics: Fundamentals and applications. Rev. Mod. Phys. 2004, 76, 323. [Google Scholar] [CrossRef]
  19. Lenz, J.; Edelstein, S. Magnetic sensors and their applications. IEEE Sens. J. 2006, 6, 631–649. [Google Scholar] [CrossRef]
  20. Julliere, M. Tunneling between ferromagnetic films. Phys. Lett. A 1975, 54, 225–226. [Google Scholar] [CrossRef]
  21. Moodera, J.S.; Kinder, L.R.; Wong, T.M.; Meservey, R. Large magnetoresistance at room temperature in ferromagnetic thin film tunnel junctions. Phys. Rev. Lett. 1995, 74, 3273. [Google Scholar] [CrossRef] [PubMed]
  22. Miyazaki, T.; Tezuka, N. Giant magnetic tunneling effect in Fe/Al2O3/Fe junction. J. Magn. Magn. Mater. 1995, 139, L231–L234. [Google Scholar] [CrossRef]
  23. Tsymbal, E.Y.; Mryasov, O.N.; LeClair, P.R. Spin-dependent tunnelling in magnetic tunnel junctions. J. Phys. Condens. Matter 2003, 15, R109. [Google Scholar] [CrossRef]
  24. Zhang, X.G.; Butler, W.H. Band structure, evanescent states, and transport in spin tunnel junctions. J. Phys. Condens. Matter 2003, 15, R1603. [Google Scholar] [CrossRef]
  25. Chen, T.; Jiang, Y.; Keen, L. Slices of parameter space for meromorphic maps with two asymptotic values. Ergod. Theory Dyn. Syst. 2023, 43, 99–139. [Google Scholar] [CrossRef]
  26. Shao, D.F.; Zhang, S.H.; Li, M.; Eom, C.B.; Tsymbal, E.Y. Spin-neutral currents for spintronics. Nat. Commun. 2021, 12, 7061. [Google Scholar] [CrossRef] [PubMed]
  27. Dong, J.; Li, X.; Gurung, G.; Zhu, M.; Zhang, P.; Zheng, F.; Tsymbal, E.Y.; Zhang, J. Tunneling magnetoresistance in noncollinear antiferromagnetic tunnel junctions. Phys. Rev. Lett. 2022, 128, 197201. [Google Scholar] [CrossRef] [PubMed]
  28. Chen, X.; Higo, T.; Tanaka, K.; Nomoto, T.; Tsai, H.; Idzuchi, H.; Shiga, M.; Sakamoto, S.; Ando, R.; Kosaki, H.; et al. Octupole-driven magnetoresistance in an antiferromagnetic tunnel junction. Nature 2023, 613, 490–495. [Google Scholar] [CrossRef] [PubMed]
  29. Qin, P.; Yan, H.; Wang, X.; Chen, H.; Meng, Z.; Dong, J.; Zhu, M.; Cai, J.; Feng, Z.; Zhou, X.; et al. Room-temperature magnetoresistance in an all-antiferromagnetic tunnel junction. Nature 2023, 613, 485–489. [Google Scholar] [CrossRef] [PubMed]
  30. Han, Y.; Luo, Q.; Hao, X.; Li, X.; Wang, F.; Hu, W.; Wu, K.; Lü, S.; Sadler, P.J. Reactions of an organoruthenium anticancer complex with 2-mercaptobenzanilide—A model for the active-site cysteine of protein tyrosine phosphatase 1B. Dalton Trans. 2011, 40, 11519–11529. [Google Scholar] [CrossRef] [PubMed]
  31. Tanaka, K.; Pfennig, N. Fermentation of 2-methoxyethanol by Acetobacterium malicum sp. nov. and Pelobacter venetianus. Arch. Microbiol. 1988, 149, 181–187. [Google Scholar] [CrossRef]
  32. Shelud’kO, A.V.; Filip’eCheva, Y.A.; Telesheva, E.M.; Burov, A.M.; Evstigneeva, S.S.; Burygin, G.L.; Petrova, L.P. Characterization of carbohydrate-containing components of Azospirillum brasilense Sp245 biofilms. Microbiology 2018, 87, 610–620. [Google Scholar] [CrossRef]
  33. Tang, J.; Guan, D.; Xu, H.; Zhao, L.; Arshad, U.; Fang, Z.; Zhu, T.; Kim, M.; Pao, C.-W.; Hu, Z.; et al. Undoped ruthenium oxide as a stable catalyst for the acidic oxygen evolution reaction. Nat. Commun. 2025, 16, 801. [Google Scholar] [CrossRef] [PubMed]
  34. Zhou, X.; Feng, W.; Zhang, R.W.; Šmejkal, L.; Sinova, J.; Mokrousov, Y.; Yao, Y. Crystal thermal transport in altermagnetic RuO2. Phys. Rev. Lett. 2024, 132, 056701. [Google Scholar] [CrossRef] [PubMed]
  35. Jiang, Y.Y.; Wang, Z.A.; Samanta, K.; Zhang, S.H.; Xiao, R.C.; Lu, W.J.; Sun, Y.P.; Tsymbal, E.Y.; Shao, D.F. Prediction of giant tunneling magnetoresistance in Ru O2/TiO2/RuO2 (110) antiferromagnetic tunnel junctions. Phys. Rev. B 2023, 108, 174439. [Google Scholar] [CrossRef]
  36. Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Phys. Rev. 1964, 136, B864. [Google Scholar] [CrossRef]
  37. Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558. [Google Scholar] [CrossRef] [PubMed]
  38. 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. [Google Scholar] [CrossRef] [PubMed]
  39. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758. [Google Scholar] [CrossRef]
  40. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed]
  41. Taylor, J.; Guo, H.; Wang, J. Ab initio modeling of quantum transport properties of molecular electronic devices. Phys. Rev. B 2001, 63, 245407. [Google Scholar] [CrossRef]
  42. Brandbyge, M.; Mozos, J.-L.; Ordejón, P.; Taylor, J.; Stokbro, K. Density-functional method for nonequilibrium electron transport. Phys. Rev. B 2002, 65, 165401. [Google Scholar] [CrossRef]
  43. Hamann, D.R. Optimized norm-conserving Vanderbilt pseudopotentials. Phys. Rev. B—Condens. Matter Mater. Phys. 2013, 88, 085117. [Google Scholar] [CrossRef]
  44. 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] [PubMed]
  45. Dudarev, S.L.; Botton, G.A.; Savrasov, S.Y.; Humphreys, C.J.; Sutton, A.P. Electron-energy-loss spectra and the structural stability of nickel oxide: An LSDA+ U study. Phys. Rev. B 1998, 57, 1505. [Google Scholar] [CrossRef]
  46. Guan, D.; Xu, H.; Huang, Y.; Jing, C.; Tsujimoto, Y.; Xu, X.; Lin, Z.; Tang, J.; Wang, Z.; Sun, X.; et al. Operando studies redirect spatiotemporal restructuration of model coordinated oxides in electrochemical oxidation. Adv. Mater. 2025, 37, 2413073. [Google Scholar] [CrossRef] [PubMed]
  47. Xu, Y.; Li, W.; Wang, C.; Chen, Z.; Wu, Y.; Zhang, X.; Li, J.; Lin, S.; Chen, Y.; Pei, Y. MnTe2 as a novel promising thermoelectric material. J. Mater. 2018, 4, 215–220. [Google Scholar] [CrossRef]
  48. Fields, S.S.; Callahan, P.G.; Combs, N.G.; Cress, C.D.; Bennett, S.P. Orientation control and mosaicity in heteroepitaxial RuO2 thin films grown through reactive direct current sputtering. Cryst. Growth Des. 2024, 24, 4604–4612. [Google Scholar] [CrossRef]
  49. Katal, R.; Masudy-Panah, S.; Tanhaei, M.; Farahani, M.H.D.A.; Jiangyong, H. A review on the synthesis of the various types of anatase TiO2 facets and their applications for photocatalysis. Chem. Eng. J. 2020, 384, 123384. [Google Scholar] [CrossRef]
  50. Pawula, F.; Fakih, A.; Daou, R.; Hébert, S.; Mordvinova, N.; Lebedev, O.; Pelloquin, D.; Maignan, A. Multiband transport in RuO2. Phys. Rev. B 2024, 110, 064432. [Google Scholar] [CrossRef]
  51. Suchaneck, G. Tunnel Spin-Polarization of ferromagnetic metals and ferrimagnetic oxides and its effect on tunnel magnetoresistance. Electron. Mater. 2022, 3, 227–234. [Google Scholar] [CrossRef]
Figure 1. (a) The atomic structure of RuO2; (b) the atomic structure of TiO2.
Figure 1. (a) The atomic structure of RuO2; (b) the atomic structure of TiO2.
Nanomaterials 15 01129 g001
Figure 2. (a) Band structure and (b) DOS for RuO2; (c) band structure and (d) DOS for TiO2. (e) The 3D Brillouin zone of RuO2; (f) coordinates of RuO2’s high-symmetry points.
Figure 2. (a) Band structure and (b) DOS for RuO2; (c) band structure and (d) DOS for TiO2. (e) The 3D Brillouin zone of RuO2; (f) coordinates of RuO2’s high-symmetry points.
Nanomaterials 15 01129 g002
Figure 3. The calculated plots of (a) electrical conductivity, (b) Seebeck coefficient, (c) electronic and phonon thermal conductivity, and (d) ZT value versus temperature for RuO2.
Figure 3. The calculated plots of (a) electrical conductivity, (b) Seebeck coefficient, (c) electronic and phonon thermal conductivity, and (d) ZT value versus temperature for RuO2.
Nanomaterials 15 01129 g003
Figure 4. RuO2/TiO2/RuO2 device: (a) magnetic moment arrangement of PC state; (b) magnetic moment arrangement of APC state. (c) Schematic diagram of the RuO2/TiO2/RuO2 device. The red arrow indicates the direction of the magnetic moment on the Ru atom.
Figure 4. RuO2/TiO2/RuO2 device: (a) magnetic moment arrangement of PC state; (b) magnetic moment arrangement of APC state. (c) Schematic diagram of the RuO2/TiO2/RuO2 device. The red arrow indicates the direction of the magnetic moment on the Ru atom.
Nanomaterials 15 01129 g004
Figure 5. Plot of thermal spin polarization current versus TL for RuO2/TiO2/RuO2 device in the (a) PC and (b) APC; (e) TIMR versus TL. The TL dependence of thermal spin total current for the RuO2/TiO2/RuO2 device is shown in the (c) PC and (d) APC.
Figure 5. Plot of thermal spin polarization current versus TL for RuO2/TiO2/RuO2 device in the (a) PC and (b) APC; (e) TIMR versus TL. The TL dependence of thermal spin total current for the RuO2/TiO2/RuO2 device is shown in the (c) PC and (d) APC.
Nanomaterials 15 01129 g005
Figure 6. Plot of thermal spin polarization current versus ΔT for RuO2/TiO2/RuO2 device in the (a) PC and (b) APC; (e) TIMR versus ΔT. Plot of thermal spin total current versus ΔT for RuO2/TiO2/RuO2 device in the (c) PC and (d) APC.
Figure 6. Plot of thermal spin polarization current versus ΔT for RuO2/TiO2/RuO2 device in the (a) PC and (b) APC; (e) TIMR versus ΔT. Plot of thermal spin total current versus ΔT for RuO2/TiO2/RuO2 device in the (c) PC and (d) APC.
Nanomaterials 15 01129 g006
Figure 7. The Fermi distribution of the electrode at different temperatures.
Figure 7. The Fermi distribution of the electrode at different temperatures.
Nanomaterials 15 01129 g007
Figure 8. Transmission spectra of RuO2/TiO2/RuO2 devices in the PC at (a) total, (b) along the z-direction, (c) spin-up, and (d) spin-down.
Figure 8. Transmission spectra of RuO2/TiO2/RuO2 devices in the PC at (a) total, (b) along the z-direction, (c) spin-up, and (d) spin-down.
Nanomaterials 15 01129 g008
Figure 9. Transmission spectra of RuO2/TiO2/RuO2 devices in the APC at (a) total, (b) along the z-direction, (c) spin-up, and (d) spin-down.
Figure 9. Transmission spectra of RuO2/TiO2/RuO2 devices in the APC at (a) total, (b) along the z-direction, (c) spin-up, and (d) spin-down.
Nanomaterials 15 01129 g009
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

Liu, J.; Ning, C.; Liu, X.; Zhu, S.; Wang, S. Investigation of Thermoelectric Properties in Altermagnet RuO2. Nanomaterials 2025, 15, 1129. https://doi.org/10.3390/nano15141129

AMA Style

Liu J, Ning C, Liu X, Zhu S, Wang S. Investigation of Thermoelectric Properties in Altermagnet RuO2. Nanomaterials. 2025; 15(14):1129. https://doi.org/10.3390/nano15141129

Chicago/Turabian Style

Liu, Jun, Chunmin Ning, Xiao Liu, Sicong Zhu, and Shuling Wang. 2025. "Investigation of Thermoelectric Properties in Altermagnet RuO2" Nanomaterials 15, no. 14: 1129. https://doi.org/10.3390/nano15141129

APA Style

Liu, J., Ning, C., Liu, X., Zhu, S., & Wang, S. (2025). Investigation of Thermoelectric Properties in Altermagnet RuO2. Nanomaterials, 15(14), 1129. https://doi.org/10.3390/nano15141129

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