Next Article in Journal
Manganese-Based Electrocatalysts for Acidic Oxygen Evolution: Development and Performance Evaluation
Previous Article in Journal
Covalent Organic Framework-Based Nanomembrane with Co-Immobilized Dual Enzymes for Micropollutant Removal
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Understanding the Magnetic Exchange Pathways of Transition-Metal-Doped Monolayer TiS2 Using First-Principles Calculations

Department of Physics, University of North Florida, Jacksonville, FL 32224, USA
*
Author to whom correspondence should be addressed.
Nanomaterials 2025, 15(18), 1435; https://doi.org/10.3390/nano15181435
Submission received: 28 August 2025 / Revised: 7 September 2025 / Accepted: 15 September 2025 / Published: 18 September 2025
(This article belongs to the Section Theory and Simulation of Nanostructures)

Abstract

The ideal crystal symmetry of the 1T-TiS2 lattice results in a non-magnetic structure. However, recent studies have demonstrated that it may become magnetic upon substitution with transition-metal (TM) atoms. In this study, we examine the mechanisms and interactions that allow magnetic exchange through the TiS2 matrix. Using density functional theory, we model the substitutional TM-doped TiS2 (TM = V, Cr, or Mn) system with varying spatial distances to examine the effects on the magnetic exchange. Since pristine 1T-TiS2 is weakly semiconducting, there is a possibility that the introduction of metallic atoms may induce an RKKY-like interaction. We find that the substitution of vanadium produces a standard exchange through the orbital interactions. However, the introduction of chromium and manganese may generate RKKY interactions with the conduction electrons. Overall, a more comprehensive understanding of how different dopants affect magnetic behavior and communicate through the lattice can enable the design of spintronic devices, which offer the potential for more energy-efficient technologies and a deeper understanding of low-dimensional systems.

1. Introduction

Advancements in technologies like machine learning, artificial intelligence, and cryptocurrency continue to increase energy demands [1,2,3], creating a need for further optimizations in energy transfer and memory storage density. The area of spintronics offers a pathway towards more efficient energy technologies by exploring the potential of utilizing low-energy spin interactions over electronic excitations [4,5,6]. For example, magnetic random access memory (MRAM) is a spintronic device that uses electron spin as the key mechanism for data storage [7,8,9]. While various hurdles must be overcome for spintronics and MRAM to replace traditional semiconductor-based memory, there is evidence that supports the idea that MRAM could be used to minimize the downsides of modern memory storage, such as energy loss and heat dissipation [10].
Developing successful spintronic devices requires materials that meet ideal specifications that range from an exact thickness of material to a precise magnetic moment to ground state magnetic behavior (ferromagnetic, antiferromagnetic, etc.) [4,5,6]. Therefore, 2D materials provide the perfect construct for versatility and design [11,12,13]. Further research on 2D materials could provide the precise specifications necessary for spintronic devices to achieve a memory density beyond the current state of 1G RAM, reaching a more competitive [14] or industry-standard number [11,12,13].
Such 2D materials enable the optimization of current systems, such as traditional flash memory and semiconductor memory, in addition to providing integral contributions toward innovative solutions (such as spintronics, valleytronics, etc.) [13,15,16,17]. Moreover, 2D materials offer more than reduced dimensionality. They are powerfully tunable and can give specific properties [18] that depend on various factors, such as exact stoichiometry of elements within a sample or defect engineering [19,20,21,22], external doping either during synthesis or post-synthesis [23,24,25], and even interactions between heterostructure interfaces [26,27]. Whether it is an exact bandgap required for an optoelectronic device or a specific magnetic ground state and/or moment to be incorporated within an SRAM device, 2D materials enable devices to be more optimized than ever.
Transition metal dichalcogenides (TMDs) are a family of van der Waals (vdW) materials characterized by their chemical formula of MX2, with M representing a transition metal (Ti, Mo, and W, etc.) and X representing a chalcogen (S, Se, Te) [28]. TMDs exhibit layers of chalcogen atoms between singular layers of transition metal atoms and often exist in a 1T (space group 164, P 3 ¯ m 1 ) or 2H (space group 194, P 6 3 / m m c ) configuration [29].
TiS2 is a TMD that has been shown computationally to exhibit a controllable magnetic state through doping [30]. While TiS2 in its bulk form has been widely studied [25,31], questions remain regarding how doping will impact a TiS2 thin film or monolayer [32,33,34]. Among the various possible applications of TiS2 are uses within the field of spintronics, in ion batteries [35,36], and in photovoltaic devices [37,38,39]. Recent computational studies demonstrate how magnetism can be introduced into the TiS2 matrix via substitutional transition-metal doping [30,32,33,40]. Therefore, our goal is to expand on the existing literature and gain an understanding of how magnetism communicates through the TiS2 lattice.
In this study, we examine the nature of the magnetic exchange interactions within monolayer 1T-TiS2. To achieve this, we specifically investigate the effects of substituting transition-metal atoms into the Ti sublattice. Under ideal conditions, Ti is typically non-magnetic due to the standard 4+ oxidation state. However, if the Ti site is directly substituted with transition-metal (TM) atoms (V, Cr, Mn), the additional electrons dope the site and provide a net magnetism in the lattice. Using density functional theory, we model the exchange interaction of TiS2 with one and two TM atoms (TM = V, Cr, or Mn). In the two-TM-atom system, we examine the magnetic ground state and exchange interaction with varying spatial separations along the zig-zag direction. The goal is to determine how the electrons in the TiS2 lattice mediate the magnetic interactions. Since pristine TiS2 is weakly semiconducting, there is a possibility that the introduction of metallic atoms may introduce an RKKY-like interaction through the conduction electrons. In general, we find that the substitution of vanadium produces a standard exchange through the orbital interactions. However, the introduction of chromium and manganese may generate the presence of RKKY interactions with the conduction electrons, although only Cr produces a weakly metallic state.

2. Methodology

Density functional theory calculations performed within this paper are done with the QuantumATK software [41,42,43]. Each calculation employs a spin-polarized meta-generalized gradient approximation (SMGGA) and SCAN functional, utilizing a linear combination of atomic orbitals (LCAO) basis set. Compared to LDA and GGA, SCAN improves predictions of ground-state structures, magnetic moments, and electronic properties, especially in strongly correlated systems [44,45], and often matches or exceeds the accuracy of hybrid functionals but at a computational cost closer to GGA functionals [45]. A PseudoDojo pseudopotential is applied [43] as well as a k-point sampling of 3 × 3 × 1. The iteration control parameter is set to 10−7 Hartrees, and the maximum force for geometry optimization is 0.05 eV/Å. The benchmarking process ensures agreement with previously established experimental data for the pristine TiS2 bandgap range. Notably, each system is geometry optimized, converting experimentally determined atomic configurations into absolute DFT-determined ground states. Geometry optimization is especially necessary when changing orbital structures within the lattice, as atomic repositioning will occur as a result of evolving ground-state electron densities.
Magnetism induced by transition metal doping in TiS2 is analyzed by directly substituting titanium atoms within an 8 × 8 × 1 supercell with two transition metal atoms (V, Cr, Mn), as demonstrated in Figure 1. As the dopants provide additional electrons to the lattice, it is expected that they will gain a significant magnetic moment. Comparing the expected magnetic moment ( μ B ) with the calculated magnetic moment provides a benchmarking method to ensure that the calculations are properly evaluated. It should be noted that the SCAN function typically provides an overestimation for the magnetic moment in an attempt to correct for underestimations of the GGA calculation [44,46,47].
To examine the exchange interactions between TM dopants, we examine the energy differences between ferromagnetic and antiferromagnetic configurations. For each system (shown in Figure 1), two sets of initial spin parameters are tested. One set of initial spins corresponds with a ferromagnetic configuration, while the other corresponds with an antiferromagnetic alignment. Through a comparison of the total energies, we can determine the global ground state between the configurations. With the ground state known, we can evaluate the energy of the other configuration without any optimization, allowing us to compare only the electron shifts in the system. A comparison of the energies provides the exchange interaction.
Using a Heisenberg exchange analysis, the given energy requirement of a spin-flip for a transition-metal dopant corresponds with the interaction energy. The Heisenberg Hamiltonian is given by
H ^ = i j J i j S i · S j ,
where H ^ is the Hamiltonian operator, J i j is the exchange interaction term, and S i and S j are the spin operators for sites i and j, respectively. A positive exchange term J i j indicates a ferromagnetic system where spins energetically favor an aligned configuration, whereas a negative J i j corresponds with an antiferromagnetic configuration. To calculate the exchange interaction, we determine the total energy from both the geometry-optimized AFM and FM configurations. Once the ground state is determined, the non-ground state configuration is recalculated using the ground state structure without a geometry optimization.
In addition, determining the exchange interaction J i j as a function of the dopant separation helps provide the type of magnetic exchange present. A standard Heisenberg exchange J e x is controlled through orbital interactions and has either ferromagnetic or antiferromagnetic character. Additionally, the exchange typically drops off as J e x e r as the magnetic centers are separated [48].
Another interaction is the Ruderman–Kittel–Kasuya–Yosida (RKKY) interaction J R K K Y [49,50,51]. This interaction is typically governed by conduction electrons and provides modulation between ferromagnetic and antiferromagnetic character, dropping off in strength as the dopants are separated. Therefore, it is typically given by
J R K K Y 1 r 3 c o s 2 k F r
where r is the separation distance, and k F is the wavevector at the Fermi level. As such, examining the relationship between magnetic exchange energy and separation between magnetic atoms provides insight regarding the interaction mechanisms.

3. Results

Using density functional theory, we investigate the electronic and magnetic properties of the TM dopants in TiS2 by examining the electronic band structure, density of states, magnetic moment (as determined by the Mulliken population), and total energy of the systems. Additionally, a novel method is employed to determine the magnetic exchange energy between dopant atoms, which is established by approximating the magnetic exchange energy between two magnetic atoms as the energy required to flip the spin state of one of the atoms. Assuming titanium and sulfur atoms within a pristine TiS2 sample are non-magnetic due to the symmetry present in covalent bonding, we are capable of directly substituting transition metals (V, Cr, Mn) into titanium sites, resulting in an isolation of magnetism within the sample.
As V, Cr, and Mn provide additional electrons relative to the Ti atom they are replacing, it follows that each doped system should have a magnetic moment with respect to the transition metal dopants. For example, an 8 × 8 × 1 supercell of TiS2 with two titanium atom sites that have been directly substituted with vanadium atoms will have a magnetic moment of approximately one Bohr magneton at each dopant atom. DFT-based magnetic moment analysis provides a method for benchmarking our results against what we expect through fundamental chemistry (Table 1).
Within this paper, transition metal-doped systems are defined based on their geometries as shown in Figure 1. Prefixes 1 and 2 denote the number of transition metal dopants present within the lattice (1 or 2), and subscripts denote the titanium separations between dopants (1, 2, 3, or 4). In terms of vanadium dopant analysis, we consider that directly substituting a V atom for a Ti results in one additional electron in an otherwise perfectly covalently bonded system. The additional electron then provides a magnetic moment of one Bohr magneton, corresponding with a spin- 1 2 system. The DFT calculation for the single-V-doped 1T-TiS2 system demonstrates a magnetic moment of 1.073 μ B with a moment of 1.091 μ B localized on the V atom, in agreement with expectations.
The systems with two V dopants all have total magnetic moments beyond two Bohr magnetons, which also agrees with expectations. As shown in Figure 2, each magnetic moment is positive, indicating ferromagnetic alignment for each dopant. The magnetic exchange interactions are analyzed through the total energy difference between the antiferromagnetic and ferromagnetic configurations. System 2V1 has the strongest magnetic exchange with 0.23 eV, which quickly decays as shown in Figure 3a. Systems 2V2 and 2V3 have magnetic exchanges of 0.0538 eV at 6.619 Å and 0.0103 eV at 10.218 Å respectively, with system 2V4 having 0 exchange at 13.667 Å. The V-doped system exhibits a standard Heisenberg exchange interaction, as indicated by its exponential decay behavior.
While the initial exchange interaction is a bit high, this is due to the initial separation of the TM atoms being fairly close together, which leads to considerable orbital overlap and a high exchange interaction. In most chalcogenide materials, the exchange is measured to be about an order of magnitude lower [52,53]. However, those materials have exchange distances closer to our second or third positions.
Additionally, the density of states for the V-doped system (shown in Figure 4) indicates that the system remains semiconducting upon the addition of the V atoms. Figure 5 breaks down the total density of states into the projected density of states per orbital, which indicates a Fermi level shift from S-p to Ti-d excitations (Figure 5a,b) as observed in the pristine TiS2 system to V-d to Ti-d (Figure 5c–e). It should be noted that in the legends of Figure 4 and Figure 5, l is the orbital angular momentum, where 1 corresponds to p-orbitals and 2 corresponds to d-orbitals. The m is the z component of orbital angular momentum corresponding to the individual d-orbitals. Overall, this suggests that the system interacts through orbital interactions, with the primary exchange being V to Ti.
In the Cr-doped system, Cr has two more electrons than Ti, and as such, we expect the Cr-doped systems to provide a spin-1 for each direct substitution within the TiS2 lattice. We first note that the single Cr system has a net magnetic moment of 2.914 μ B , which is high for a spin-1 system. However, referring to the density of states present in Figure 4, we note the presence of electronic states at the Fermi level, indicating conduction electrons within the system. In addition, the calculation of the magnetic moment is given by
μ S = n ( n + 2 ) μ B ,
where n is the number of unpaired electrons and provides further insight regarding the large magnetic moment. This spin-only formula is especially useful when analyzing transition metal complexes within the first series, as orbital momentum contributions are small [54,55,56,57,58]. Using the spin-only formula, the expected μ S 2.83 μ B is much closer to our calculated value of 2.914 μ B .
Furthermore, the 2-Cr systems have a magnetic ground state configuration that depends on the separation of magnetic atoms present. At a separation of 3.453 Å, the system exists in a ferromagnetic ground state. For further separations, the spins prefer to be misaligned in an antiferromagnetic configuration (shown in Figure 3b). The oscillatory exchange interaction is a characteristic of the RKKY interaction, suggesting that conduction electrons play a role in the exchange of magnetism within the system.
Figure 4 shows the density of states for the Cr-doped system and clearly indicates that the system is weakly metallic by having an apparent density of states at the Fermi energy. Figure 5f–h shows that the states that appear at the Fermi level are a mixing of Cr-d and S-p orbitals, indicative of a conduction pathway. Overall, this further strengthens the argument that Cr is producing a distinct RKKY-like interaction.
The examination of the DOS suggests that the metallic characteristics for the Cr system are produced by a shift of the oxidation state from 4+ to 3+, which explains the higher calculated magnetic moment. The gain of an electron to the chromium atom would leave a hole in the lattice and allow for conductive states at the Fermi level.
Moving to the Mn-doped system, Mn has three more electrons than Ti, and we expect a spin- 3 2 system. The single Mn-doped system has an overall net magnetic moment of 2.995 μ B with the magnetic moment local to the Mn-dopant being 3.102 μ B , which agrees with our initial assumption of a spin- 3 2 system. Similar to the Cr-doped systems, Mn also demonstrates a slight modulation of the magnetic ground state that depends on the separation between the exchanging atoms (shown in Figure 3c). The relative exchange interaction between the first and second separations is drastically reduced compared to the Cr systems. However, it is slightly antiferromagnetic and within the energy resolution of the calculation, which suggests that an RKKY interaction could govern it. Regardless of the nature of the mechanism, the Mn-doped system exhibits a dramatic drop-off in magnetic exchange energy as the separation increases.
To gain further insight, we re-examine the density of states. As shown in Figure 4, there are no conduction electrons at the Fermi level, and a clear bandgap exists, which does not support the idea of an RKKY-like interaction. Furthermore, looking at the projected density of states for the Mn-doped system (Figure 5i–k) reveals that the Mn-d orbitals are not greatly in play, which is likely why the exchange interaction drops off so dramatically. As such, further investigation is required before a dominant exchange interaction mechanism can be formally suggested for this system.

4. Conclusions

Magnetic exchange pathways are computationally investigated through transition metal doping within a monolayer TiS2 lattice. Given the non-magnetic response within a pristine sample, directly substituting various transition metals (V, Cr, and Mn) into Ti locations within the sample allows for a controlled method of isolating magnetic atoms and subsequent analysis. Four different systems are calculated for each dopant, each with a distinct atomic separation. Magnetic exchange interactions for each system are computed by setting two calculations, one with initial spin parameters corresponding to a ferromagnetic alignment and one with an antiferromagnetic alignment. The alignment that results in a lower total energy is then used and tested with the opposite alignment, without geometry optimization, to isolate any geometric factors contributing to the total energy difference.
V-substituted monolayer TiS2 is shown to demonstrate a standard Heisenberg exchange, as the relationship between the atomic separation of magnetic atoms and the magnetic exchange energy is a characteristic decaying exponential. The magnetic moment of V dopants is calculated to be 1 μ B , corresponding with the expected value. Cr-substitutions result in an oscillating magnetic exchange based on the separation of dopants, meaning that the magnetic alignment of Cr atoms alternates between ferromagnetic and antiferromagnetic as a function of separation. While this is characteristic of an RKKY interaction, there are no states available within the conduction band, requiring a more in-depth investigation to understand the underlying mechanisms at work. Cr dopants have a magnetic moment of 2.9 μ B , which is slightly higher than the expected value but agrees with spin-only considerations. Mn dopants technically result in an oscillating magnetic exchange interaction. However, the antiferromagnetic alignments are very small in energy, at less than 0.1 eV. In addition, there are no available conduction electrons at the Fermi level. Mn dopants are predicted to have a magnetic moment of 3.1 μ B , in agreement with the expected three electrons per dopant. Analysis of the density of states helps confirm these exchange mechanisms.
Overall, the presence of magnetic dopants in the Ti sites of TiS2 does produce magnetism in the system. Due to the weakly semiconducting nature of TiS2, the addition of extra electrons into the system appears to be able to create either a standard exchange or an RKKY exchange mechanism, depending on the presence of metallicity in the system. As such, careful consideration is required when analyzing magnetic exchange interactions within the TiS2 lattice, highlighting the potential for further investigation into robust magnetic coupling in these systems. As of the writing of this manuscript, there are no experimental realizations of these substitutions. However, this study does provide insight into how magnetic exchange couples through the lattice and at what distances this occurs, which can have implications on the magnetism caused by substitutional dopants, intercalated dopants, vacancies, or defects.

Author Contributions

Conceptualization, P.M.C. and J.T.H.; methodology, P.M.C. and J.T.H.; software, J.T.H.; validation, P.M.C. and J.T.H.; formal analysis, P.J.K., P.M.C. and J.T.H.; investigation, P.J.K.; resources, P.J.K., P.M.C. and J.T.H.; data curation, P.J.K.; writing—original draft preparation, P.J.K.; writing—review and editing, P.J.K., P.M.C. and J.T.H.; visualization, P.J.K. and J.T.H.; supervision, P.M.C. and J.T.H.; project administration, J.T.H.; funding acquisition, J.T.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Institute for Materials Science at Los Alamos National Laboratory under subcontract No. C5496.

Data Availability Statement

All data is available upon reasonable request to the authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Krause, M.J.; Tolaymat, T. Quantification of energy and carbon costs for mining cryptocurrencies. Nat. Sustain. 2018, 1, 711–718. [Google Scholar] [CrossRef]
  2. Kristiansen Nøland, J.; Hjelmeland, M.; Korpås, M. Will Energy-Hungry AI Create a Baseload Power Demand Boom? IEEE Access 2024, 12, 110353–110360. [Google Scholar] [CrossRef]
  3. García-Martín, E.; Rodrigues, C.F.; Riley, G.; Grahn, H. Estimation of energy consumption in machine learning. J. Parallel Distrib. Comput. 2019, 134, 75–88. [Google Scholar] [CrossRef]
  4. Rimmler, B.H.; Pal, B.; Parkin, S.S.P. Non-collinear antiferromagnetic spintronics. Nat. Rev. Mater. 2025, 10, 109–127. [Google Scholar] [CrossRef]
  5. Al-Douri, Y.; Ameri, M. Physical studies of spintronics-based Heusler alloys. Crit. Rev. Solid State Mater. Sci. 2025, 50, 189–238. [Google Scholar] [CrossRef]
  6. Xie, Y.; Zhang, S.Y.; Yin, Y.; Zheng, N.; Ali, A.; Younis, M.; Ruan, S.; Zeng, Y.J. Emerging ferromagnetic materials for electrical spin injection: Towards semiconductor spintronics. npj Spintron. 2025, 3, 10. [Google Scholar] [CrossRef]
  7. Slaughter, J.M.; Dave, R.W.; DeHerrera, M.; Durlam, M.; Engel, B.N.; Janesky, J.; Rizzo, N.D.; Tehrani, S. Fundamentals of MRAM Technology. J. Supercond. 2002, 15, 19–25. [Google Scholar] [CrossRef]
  8. Huai, Y. Spin-transfer torque MRAM (STT-MRAM): Challenges and prospects. AAPPS Bull. 2008, 18, 33–40. [Google Scholar]
  9. Bhatti, S.; Sbiaa, R.; Hirohata, A.; Ohno, H.; Fukami, S.; Piramanayagam, S. Spintronics based random access memory: A review. Mater. Today 2017, 20, 530–548. [Google Scholar] [CrossRef]
  10. Na, T.; Kang, S.H.; Jung, S.O. STT-MRAM sensing: A review. IEEE Trans. Circuits Syst. II Express Briefs 2020, 68, 12–18. [Google Scholar] [CrossRef]
  11. Glavin, N.R.; Rao, R.; Varshney, V.; Bianco, E.; Apte, A.; Roy, A.; Ringe, E.; Ajayan, P.M. Emerging applications of elemental 2D materials. Adv. Mater. 2020, 32, 1904302. [Google Scholar] [CrossRef]
  12. Feng, Y.P.; Shen, L.; Yang, M.; Wang, A.; Zeng, M.; Wu, Q.; Chintalapati, S.; Chang, C.R. Prospects of spintronics based on 2D materials. WIREs Comput. Mol. Sci. 2017, 7, e1313. [Google Scholar] [CrossRef]
  13. Yu, X.; Xu, L.; Shi, W.; Meng, X.; Huang, X.; Peng, Z.; Tong, L.; Sun, H.; Miao, X.; Ye, L. 2D materials-based flash memory device: Mechanism, structure, application. Mater. Horiz. 2025. Advance Article. [Google Scholar] [CrossRef]
  14. Aggarwal, S.; Almasi, H.; DeHerrera, M.; Hughes, B.; Ikegawa, S.; Janesky, J.; Lee, H.; Lu, H.; Mancoff, F.; Nagel, K.; et al. Demonstration of a reliable 1 Gb standalone spin-transfer torque MRAM for industrial applications. In Proceedings of the 2019 IEEE International Electron Devices Meeting (IEDM), San Francisco, CA, USA, 7–11 December 2019; IEEE: Piscataway, NJ, USA, 2019; pp. 2.1.1–2.1.4. [Google Scholar]
  15. Yakout, S.M. Spintronics: Future Technology for New Data Storage and Communication Devices. J. Supercond. Nov. Magn. 2020, 33, 2557–2580. [Google Scholar] [CrossRef]
  16. Butler, S.Z.; Hollen, S.M.; Cao, L.; Cui, Y.; Gupta, J.A.; Gutiérrez, H.R.; Heinz, T.F.; Hong, S.S.; Huang, J.; Ismach, A.F.; et al. Progress, Challenges, and Opportunities in Two-Dimensional Materials Beyond Graphene. ACS Nano 2013, 7, 2898–2926. [Google Scholar] [CrossRef] [PubMed]
  17. Bertolazzi, S.; Bondavalli, P.; Roche, S.; San, T.; Choi, S.Y.; Colombo, L.; Bonaccorso, F.; Samorì, P. Nonvolatile memories based on graphene and related 2D materials. Adv. Mater. 2019, 31, 1806663. [Google Scholar] [CrossRef]
  18. Sethulakshmi, N.; Mishra, A.; Ajayan, P.; Kawazoe, Y.; Roy, A.K.; Singh, A.K.; Tiwary, C.S. Magnetism in two-dimensional materials beyond graphene. Mater. Today 2019, 27, 107–122. [Google Scholar] [CrossRef]
  19. Min, J.; Kim, J.H.; Kang, J. Chalcogen Vacancy Engineering of Two-Dimensional Transition Metal Dichalcogenides for Electronic Applications. ACS Appl. Nano Mater. 2024, 7, 26377–26396. [Google Scholar] [CrossRef]
  20. Zhang, X.; Gao, L.; Yu, H.; Liao, Q.; Kang, Z.; Zhang, Z.; Zhang, Y. Single-Atom Vacancy Doping in Two-Dimensional Transition Metal Dichalcogenides. Acc. Mater. Res. 2021, 2, 655–668. [Google Scholar] [CrossRef]
  21. Hawkins, C.G.; Whittaker-Brooks, L. Controlling Sulfur Vacancies in TiS2-x Cathode Insertion Hosts via the Conversion of TiS3 Nanobelts for Energy-Storage Applications. ACS Appl. Nano Mater. 2018, 1, 851–859. [Google Scholar] [CrossRef]
  22. Chen, K.; Song, M.; Sun, Y.Y.; Xu, H.; Qi, D.C.; Su, Z.; Gao, X.; Xu, Q.; Hu, J.; Zhu, J.; et al. Defects controlled doping and electrical transport in TiS2 single crystals. Appl. Phys. Lett. 2020, 116, 121901. [Google Scholar] [CrossRef]
  23. Loh, L.; Zhang, Z.; Bosman, M.; Eda, G. Substitutional doping in 2D transition metal dichalcogenides. Nano Res. 2021, 14, 1668–1681. [Google Scholar] [CrossRef]
  24. Zhao, Y.; Xu, K.; Pan, F.; Zhou, C.; Zhou, F.; Chai, Y. Doping, Contact and Interface Engineering of Two-Dimensional Layered Transition Metal Dichalcogenides Transistors. Adv. Funct. Mater. 2017, 27, 1603484. [Google Scholar] [CrossRef]
  25. Chen, S.; Yang, L.; Wang, D. Photoelectric Properties of Single Layer TiS2 Modified by Non-Metal Doping. Russ. J. Phys. Chem. A 2022, 96, 3031–3037. [Google Scholar] [CrossRef]
  26. Chen, H.; Yin, J.; Yang, J.; Zhang, X.; Liu, M.; Jiang, Z.; Wang, J.; Sun, Z.; Guo, T.; Liu, W.; et al. Transition-metal dichalcogenides heterostructure saturable absorbers for ultrafast photonics. Opt. Lett. 2017, 42, 4279–4282. [Google Scholar] [CrossRef]
  27. Selamneni, V.; Sahatiya, P. Mixed dimensional Transition Metal Dichalcogenides (TMDs) vdW Heterostructure based Photodetectors: A review. Microelectron. Eng. 2023, 269, 111926. [Google Scholar] [CrossRef]
  28. Choi, W.; Choudhary, N.; Han, G.H.; Park, J.; Akinwande, D.; Lee, Y.H. Recent development of two-dimensional transition metal dichalcogenides and their applications. Mater. Today 2017, 20, 116–130. [Google Scholar] [CrossRef]
  29. Zhao, B.; Shen, D.; Zhang, Z.; Lu, P.; Hossain, M.; Li, J.; Li, B.; Duan, X. 2D Metallic Transition-Metal Dichalcogenides: Structures, Synthesis, Properties, and Applications. Adv. Funct. Mater. 2021, 31, 2105132. [Google Scholar] [CrossRef]
  30. Liu, H.; Yang, L.; Zhao, Y.; Sun, S.; Wei, X. Photoelectric and Magnetic Variation of Transition Metal-Doped Monolayer TiS2: A First-Principles Calculation. J. Supercond. Nov. Magn. 2024, 37, 639–655. [Google Scholar] [CrossRef]
  31. Let, A.L.; Mainwaing, D.; Rix, C.; Murugaraj, P. Synthesis and Optical Properties of TiS2 Nanoclusters. Rev. Roum. Chim. 2007, 52, 235–241. [Google Scholar]
  32. Zeb, J.; Zhao, X.; Ullah, S.; Menezes, M.G.; Zhang, W. Tunable optoelectronic properties in multilayer 1T-TiS2: The effects of strain and an external electric field. J. Mater. Sci. 2021, 56, 6891–6902. [Google Scholar] [CrossRef]
  33. El-Kouch, H.; Farh, L.E. Electronic and Optical Properties of TiS2 Determined from Modified Becke–Johnson GGA Potential. Rev. Roum. Chim. 2018, 63, 873–880. [Google Scholar]
  34. Wang, H.; Qiu, Z.; Xia, W.; Ming, C.; Han, Y.; Cao, L.; Lu, J.; Zhang, P.; Zhang, S.; Xu, H.; et al. Semimetal or Semiconductor: The Nature of High Intrinsic Electrical Conductivity in TiS2. J. Phys. Chem. Lett. 2019, 10, 6996–7001. [Google Scholar] [CrossRef]
  35. Huang, X.; Tang, J.; Luo, B.; Knibbe, R.; Lin, T.; Hu, H.; Rana, M.; Hu, Y.; Zhu, X.; Gu, Q.; et al. Sandwich-Like Ultrathin TiS2 Nanosheets Confined within N, S Codoped Porous Carbon as an Effective Polysulfide Promoter in Lithium-Sulfur Batteries. Adv. Energy Mater. 2019, 9, 1901872. [Google Scholar] [CrossRef]
  36. Sun, K.; Fu, M.; Xie, Z.; Su, D.; Zhong, H.; Bai, J.; Dooryhee, E.; Gan, H. Improvement of Li-S battery electrochemical performance with 2D TiS2 additive. Electrochim. Acta 2018, 292, 779–788. [Google Scholar] [CrossRef]
  37. Huckaba, A.J.; Gharibzadeh, S.; Ralaiarisoa, M.; Roldán-Carmona, C.; Mohammadian, N.; Grancini, G.; Lee, Y.; Amsalem, P.; Plichta, E.J.; Koch, N.; et al. Low-Cost TiS2 as Hole-Transport Material for Perovskite Solar Cells. Small Methods 2017, 1, 1700250. [Google Scholar] [CrossRef]
  38. Alias, N.; Ali Umar, A.; Malek, N.A.A.; Liu, K.; Li, X.; Abdullah, N.A.; Rosli, M.M.; Abd Rahman, M.Y.; Shi, Z.; Zhang, X.; et al. Photoelectrical Dynamics Uplift in Perovskite Solar Cells by Atoms Thick 2D TiS2 Layer Passivation of TiO2 Nanograss Electron Transport Layer. ACS Appl. Mater. Interfaces 2021, 13, 3051–3061. [Google Scholar] [CrossRef]
  39. Mahuli, N.; Sarkar, S.K. Atomic layer deposition of titanium sulfide and its application in extremely thin absorber solar cells. J. Vac. Sci. Technol. A 2014, 33, 01A150. [Google Scholar] [CrossRef]
  40. Toularoud, S.A.; Hadipour, H.; Soleimani, H.R. Engineering the electronic and magnetic properties of monolayer TiS2 through systematic transition-metal doping. Phys. B Condens. Matter 2024, 694, 416413. [Google Scholar] [CrossRef]
  41. QuantumATK Version R-2020.09, Synopsys QuantumATK, Sunnydale CA, USA, 2020. Available online: https://www.synopsys.com/quantumatk (accessed on 17 September 2025).
  42. Smidstrup, S.; Markussen, T.; Vancraeyveld, P.; Wellendorff, J.; Schneider, J.; Gunst, T.; Verstichel, B.; Stradi, D.; Khomyakov, P.A.; Vej-Hansen, U.G.; et al. QuantumATK: An integrated platform of electronic and atomic-scale modelling tools. J. Phys. Condens. Matter 2019, 32, 015901. [Google Scholar] [CrossRef] [PubMed]
  43. van Setten, M.; Giantomassi, M.; Bousquet, E.; Verstraete, M.; Hamann, D.; Gonze, X.; Rignanese, G.M. The PseudoDojo: Training and grading a 85 element optimized norm-conserving pseudopotential table. Comput. Phys. Commun. 2018, 226, 39–54. [Google Scholar] [CrossRef]
  44. Sokolovskiy, V.; Baigutlin, D.; Miroshkina, O.; Buchelnikov, V. Meta-GGA SCAN Functional in the Prediction of Ground State Properties of Magnetic Materials: Review of the Current State. Metals 2023, 13, 728. [Google Scholar] [CrossRef]
  45. Bartók, A.P.; Yates, J.R. Regularized SCAN functional. J. Chem. Phys. 2019, 150, 161101. [Google Scholar] [CrossRef] [PubMed]
  46. Bihlmayer, G. Density Functional Theory for Magnetism and Magnetic Anisotropy. In Handbook of Materials Modeling: Methods: Theory and Modeling; Andreoni, W., Yip, S., Eds.; Springer International Publishing: Cham, Switzerland, 2018; pp. 1–23. [Google Scholar] [CrossRef]
  47. Kovács, P.; Tran, F.; Blaha, P.; Madsen, G.K.H. Comparative study of the PBE and SCAN functionals: The particular case of alkali metals. J. Chem. Phys. 2019, 150, 164119. [Google Scholar] [CrossRef]
  48. Blundell, S. Magnetism in Condensed Matter; Oxford University Press: Oxford, UK, 2001. [Google Scholar]
  49. Yosida, K. Magnetic Properties of Cu-Mn Alloys. Phys. Rev. 1957, 106, 893–898. [Google Scholar] [CrossRef]
  50. Kasuya, T. A Theory of Metallic Ferro- and Antiferromagnetism on Zener’s Model. Prog. Theor. Phys. 1956, 16, 45–57. [Google Scholar] [CrossRef]
  51. Ruderman, M.A.; Kittel, C. Indirect Exchange Coupling of Nuclear Magnetic Moments by Conduction Electrons. Phys. Rev. 1954, 96, 99–102. [Google Scholar] [CrossRef]
  52. Wang, Y.; Puggioni, D.; Rondinelli, J.M. Assessing exchange-correlation functional performance in the chalcogenide lacunar spinels GaM4Q8 (M = Mo, V, Nb, Ta; Q = S, Se). Phys. Rev. B 2019, 100, 115149. [Google Scholar] [CrossRef]
  53. Baltzer, P.K.; Wojtowicz, P.J.; Robbins, M.; Lopatin, E. Exchange Interactions in Ferromagnetic Chromium Chalcogenide Spinels. Phys. Rev. 1966, 151, 367–377. [Google Scholar] [CrossRef]
  54. Miessler, G.L.; Fischer, P.J.; Tarr, D.A. Inorganic Chemistry, 5th ed.; Pearson: London, UK, 2013. [Google Scholar]
  55. Atkins, P.; Overton, T. Shriver and Atkins’ Inorganic Chemistry, 5th ed.; Oxford University Press: Oxford, UK, 2010. [Google Scholar]
  56. Cotton, F.A.; Wilkinson, G.; Murillo, C.A.; Bochmann, M. Advanced Inorganic Chemistry, 6th ed.; Wiley-Interscience: New York, NY, USA, 1999. [Google Scholar]
  57. Ghosh, R. Bose–Stoner Formula in Magnetochemistry. Resonance 2022, 27, 1131–1133. [Google Scholar] [CrossRef]
  58. Dresselhaus, M.S. Solid State Physics Part III: Magnetic Properties of Solids. Lecture Notes from MIT Course 6.732. 2001. Available online: https://web.mit.edu/6.732/www/6.732-pt3.pdf (accessed on 17 September 2025).
Figure 1. (Left to right): 1T-TiS2 supercells demonstrating transition-metal dopants directly substituted adjacent to each other and then with one, two, and three Ti atom locations mediating the exchange interaction along with surrounding S atoms. Here, grey atoms are Ti, yellow atoms are S, and the red atoms are the transition-metal atoms.
Figure 1. (Left to right): 1T-TiS2 supercells demonstrating transition-metal dopants directly substituted adjacent to each other and then with one, two, and three Ti atom locations mediating the exchange interaction along with surrounding S atoms. Here, grey atoms are Ti, yellow atoms are S, and the red atoms are the transition-metal atoms.
Nanomaterials 15 01435 g001
Figure 2. The electron density difference for each separation of the vanadium, chromium, and manganese systems. The red-colored density indicates spin up, and the blue indicates spin down.
Figure 2. The electron density difference for each separation of the vanadium, chromium, and manganese systems. The red-colored density indicates spin up, and the blue indicates spin down.
Nanomaterials 15 01435 g002
Figure 3. Energy difference (eV) as a function of transition metal (TM) separation (Å) for (a) vanadium, (b) chromium, and (c) manganese. The energy difference is the change in energy between respective antiferromagnetic and ferromagnetic configurations of the transition-metal dopants. This is consistent with the electron density difference calculations shown in Figure 2.
Figure 3. Energy difference (eV) as a function of transition metal (TM) separation (Å) for (a) vanadium, (b) chromium, and (c) manganese. The energy difference is the change in energy between respective antiferromagnetic and ferromagnetic configurations of the transition-metal dopants. This is consistent with the electron density difference calculations shown in Figure 2.
Nanomaterials 15 01435 g003
Figure 4. Density of states plots for each transition metal-doped system. V- and Mn-doped systems exhibit semiconducting behavior with Fermi levels located within the bandgap, whereas the Cr-doped system has states available at the Fermi level. In the legend, l is the orbital angular momentum, where 1 corresponds to p-orbitals and 2 corresponds to d-orbitals. The green dotted line indicates the Fermi level.
Figure 4. Density of states plots for each transition metal-doped system. V- and Mn-doped systems exhibit semiconducting behavior with Fermi levels located within the bandgap, whereas the Cr-doped system has states available at the Fermi level. In the legend, l is the orbital angular momentum, where 1 corresponds to p-orbitals and 2 corresponds to d-orbitals. The green dotted line indicates the Fermi level.
Nanomaterials 15 01435 g004
Figure 5. The projected density of states for the pristine TiS2 for the S-p-orbitals (a) and Ti-d-orbitals (b), as well as the S-p-orbitals, Ti-d-orbitals, and TM-d-orbitals for the double V- (ce), Cr- (fh), Mn- (ik) doped systems, respectively. In the legend, l is the orbital angular momentum, where 1 corresponds to p-orbitals and 2 corresponds to d-orbitals. The m is the z-component of orbital angular momentum corresponding to the individual d-orbitals. The green dotted lines denote the Fermi level.
Figure 5. The projected density of states for the pristine TiS2 for the S-p-orbitals (a) and Ti-d-orbitals (b), as well as the S-p-orbitals, Ti-d-orbitals, and TM-d-orbitals for the double V- (ce), Cr- (fh), Mn- (ik) doped systems, respectively. In the legend, l is the orbital angular momentum, where 1 corresponds to p-orbitals and 2 corresponds to d-orbitals. The m is the z-component of orbital angular momentum corresponding to the individual d-orbitals. The green dotted lines denote the Fermi level.
Nanomaterials 15 01435 g005
Table 1. Magnetic analysis for 8 × 8 × 1 monolayer 1T-TiS2 systems with transition-metal dopants directly substituted into two Ti sites. Column 1 corresponds with Figure 1 and indicates which doped system is being considered. Column 2 provides the distance between dopants. Column 3 represents the difference in ground-state total energies for antiferromagnetic and ferromagnetic systems. Columns 4 and 5 indicate the magnetic moment corresponding to the dopant atoms. Column 6 provides the total magnetic moment for each system.
Table 1. Magnetic analysis for 8 × 8 × 1 monolayer 1T-TiS2 systems with transition-metal dopants directly substituted into two Ti sites. Column 1 corresponds with Figure 1 and indicates which doped system is being considered. Column 2 provides the distance between dopants. Column 3 represents the difference in ground-state total energies for antiferromagnetic and ferromagnetic systems. Columns 4 and 5 indicate the magnetic moment corresponding to the dopant atoms. Column 6 provides the total magnetic moment for each system.
SystemDopant Separation (Å)EA-EF (eV) M TM 1 ( μ B ) M TM 2 ( μ B )MTotal ( μ B )
1V1.0911.073
2V13.0790.230961.2311.1492.077
2V26.6190.053781.0951.1222.122
2V310.2180.01031.0821.0952.086
2V413.66701.1441.0752.179
1Cr2.9722.914
2Cr13.4530.02602.9392.945.502
2Cr26.829−0.00552.943−2.945−0.013
2Cr310.258−0.000992.948−2.9620
2Cr413.650.00012−2.9532.9545.645
1Mn3.1022.995
2Mn13.3880.464383.1363.1395.989
2Mn26.779−0.0066−3.1153.104−0.002
2Mn310.250−0.00028−3.1203.109−0.003
2Mn413.654−0.00032−3.1193.0108−0.002
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

Keeney, P.J.; Coelho, P.M.; Haraldsen, J.T. Understanding the Magnetic Exchange Pathways of Transition-Metal-Doped Monolayer TiS2 Using First-Principles Calculations. Nanomaterials 2025, 15, 1435. https://doi.org/10.3390/nano15181435

AMA Style

Keeney PJ, Coelho PM, Haraldsen JT. Understanding the Magnetic Exchange Pathways of Transition-Metal-Doped Monolayer TiS2 Using First-Principles Calculations. Nanomaterials. 2025; 15(18):1435. https://doi.org/10.3390/nano15181435

Chicago/Turabian Style

Keeney, P. J., P. M. Coelho, and J. T. Haraldsen. 2025. "Understanding the Magnetic Exchange Pathways of Transition-Metal-Doped Monolayer TiS2 Using First-Principles Calculations" Nanomaterials 15, no. 18: 1435. https://doi.org/10.3390/nano15181435

APA Style

Keeney, P. J., Coelho, P. M., & Haraldsen, J. T. (2025). Understanding the Magnetic Exchange Pathways of Transition-Metal-Doped Monolayer TiS2 Using First-Principles Calculations. Nanomaterials, 15(18), 1435. https://doi.org/10.3390/nano15181435

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