Diem Thi-Xuan Danga,
Yen Thi-Hai Pham†
a,
Da Zhou
bc,
Dai-Nam Le
a,
Mauricio Terrones
bc,
Manh-Huong Phan
*a and
Lilia M. Woods
*a
aDepartment of Physics, University of South Florida, Tampa, Florida 33620, USA. E-mail: phanm@usf.edu; lmwoods@usf.edu
bDepartment of Physics, The Pennsylvania State University, University Park, Pennsylvania 16802, USA
cCenter for 2-Dimensional and Layered Materials, The Pennsylvania State University, University Park, Pennsylvania 16802, USA
First published on 30th July 2025
Magnetism in two-dimensional materials is of great importance in discovering new physical phenomena and developing new devices at the nanoscale. In this paper, first-principles simulations are used to calculate the electronic and magnetic properties of heterostructures composed of graphene and MoS2 considering the influence of point defects and vanadium doping. It is found that the concentration of the dopants and the types of defects can result in induced magnetic moments leading to ferromagnetically polarized systems with sharp interfaces. This provides a framework for interpreting the experimental observations of enhanced ferromagnetism in both MoS2/graphene and V-doped MoS2/graphene heterostructures. The computed electronic and spin polarizations give a microscopic understanding of the origin of ferromagnetism in these systems and illustrate how doping and defect engineering can lead to targeted property tunability. Our work has demonstrated that through defects engineering, ferromagnetism can be achieved in V-doped MoS2/graphene heterostructures, providing a potential way to induce magnetization in other TMDC/graphene materials and opening new opportunities for their applications in nano-spintronics.
Typically, TMDC/graphene HSTs are non-magnetic. Nevertheless, the development of methods to effectively generate and control magnetism in 2D materials is critical for the establishment of new magnetic HSTs with various applications. An effective method to induce magnetic properties in 2D semiconductors involves doping and defect engineering.5,6 Based on previous studies, transition metal substitutional doping can effectively generate tunable magnetism in TMDC systems.7,8 Among them, Vanadium doping can realize not only p-type transport behavior in TMDCs but also room-temperature ferromagnetism, making V-doped TMDCs a favorable platform for electronic and spintronic devices.9,10
Recently, V-doped MoS2 has been intensively studied. Zhang et al. demonstrated that V-doped MoS2 increases radiative recombination of excitons relative to trions by reducing background electrons in MoS2.11 Sahoo et al. reported the presence of time-reversal asymmetry between the two K-valleys with a 32 meV relative energy shift between them.12 Maity et al. observed a semiconductor-to-metal transition in V-doped MoS2 alloy when the doping concentration increased above 5%.13 The magnetic performance of the V-doped MoS2 can be tuned by varying the doping concentration. In particular, enhanced ferromagnetism is observed for up to 3% V doping concentration,14 which is strongly inhibited by raising doping beyond 8%.14 However, due to the low doping concentration, the net magnetization of these monolayers remains relatively weak limiting their suitability for spintronic and magnetic sensor applications.
Constructing a HST of a doped TMDC exhibiting ferromagnetic behavior and a nonmagnetic 2D system such as graphene brings forward new possibilities for magnetic HSTs with sharp interfaces. Of special interest is to see if graphene can become magnetic as a result of interactions and proximity effects with the ferromagnetic TMDC.15,16 In this study, we broaden our understanding of doped TMDCs and their magnetism by investigating V-doped MoS2/graphene HSTs by taking into account various types of defects and doping concentrations. In graphene, mono- and divacancies, Stone–Wales defects, etc. have been studied quite extensively.17,18 The fabrication of TMDCs can also result in various types of vacancies,19,20 including V-doped MoS2 monolayer obtained by chemical vapor deposition methods.14 These vacancies are known to have dramatic effects on the electronic, magnetic, and optical properties of TMDC systems.14,21,22 Due to their importance, in this paper, we have systematically studied the role of point defects in the properties of the V-doped MoS2/graphene HST. We present various types of defect formations, the electronic structures of the defective materials, and the defect signatures of the magnetic properties of the different HSTs. Our work shows pathways to improve the saturation magnetization of TMDC/graphene HSTs, which can be useful for their application in spintronics.
In addition to a HST composed of pristine monolayers, several point-like defects in both monolayers are considered, because isolated point defects are unavoidable in the synthesis process as discussed earlier. Specifically, a C monovacancy (Gr + VC), C divacancies (Gr + V2C) and Stone–Wales (Gr + VSW) defect are constructed in the Gr part of the supercell. The defective TMDC monolayers are also generated in the MoS2 part of the supercell and they are: MoS2 + VS (with a S monovacancy), MoS2 + VS–S (with two adjacent S vacancies on the same S layer), MoS2 + V2S (with two S vacancies on the same Mo atom but in opposite S layers), MoS2 + VMo (with a Mo monovacancy) and MoS2 + VMo+2S (with a Mo vacancy and two S vacancies on opposite layers). To probe the magnetism in the system, Vanadium substitutional dopants at the Mo sites in the TMDC layer are simulated by taking V0.02Mo0.98S2 (2% V doping) and V0.04Mo0.96S2 (4% V doping) monolayers. Various combinations for the combined effect of point defects and dopants within the MoS2 layer are also taken by constructing V0.02Mo0.98S2 + VS (2% V doping and a S monovacancy), V0.02Mo0.98S2 + VS–S (2% V doping and two adjacent S vacancies on the same layer), V0.02Mo0.98S2 + V2S (2% V doping and two adjacent S vacancies on the different layer), V0.02Mo0.98S2 + VMo (2% V doping and a Mo monovacancy), and V0.02Mo0.98S2 + VMo+2S (with 2% V doping, a Mo vacancy and two adjacent S vacancies on the different layers) monolayers.
In this study, we simulate HSTs composed of various combinations of Gr and TMDC sheets discussed above. We find that the magnetic properties of HSTs composed of different TMDCs and pristine Gr or Gr + V2C or Gr + VSW are very similar due to the similar nonmagnetic nature of these Gr sheets. Thus, in what follows, we only consider HSTs composed of different TMDCs and pristine Gr or Gr + VC. The full list of the simulated HSTs composed of various combinations of defective and/or doped monolayers is given in Table 1, and several representative examples of the lattice configurations are shown in Fig. 1. The numerous combinations (26 structures altogether) help us to understand the complex interplay between the types of point defects in a specific monolayer and the magnetic dopants in the Gr/MoS2 HSTs.
Name | Gr strain (%) | TMDC strain (%) | Name | Gr + VC strain (%) | TMDC strain (%) |
---|---|---|---|---|---|
HST1: Gr/MoS2 | −0.234 | 0.498 | HST14: (Gr + VC)/MoS2 | −0.225 | 0.530 |
HST2: Gr/(MoS2 + VS) | −0.283 | 0.644 | HST15: (Gr + VC)/(MoS2 + VS) | −0.288 | 0.654 |
HST3: Gr/(MoS2 + VS–S) | −0.337 | 0.824 | HST16: (Gr + VC)/(MoS2 + VS–S) | −0.351 | 0.833 |
HST4: Gr/(MoS2 + V2S) | −0.342 | 0.778 | HST17: (Gr + VC)/(MoS2 + V2S) | −0.342 | 0.782 |
HST5: Gr/(MoS2 + VMo) | −0.220 | 0.489 | HST18: (Gr + VC)/(MoS2 + VMo) | −0.197 | 0.534 |
HST6: Gr/(MoS2 + VMo+2S) | −0.382 | 0.806 | HST19: (Gr + VC)/(MoS2 + VMo+2S) | −0.697 | 0.708 |
HST7: Gr/V0.02Mo0.98S2 | −0.247 | 0.426 | HST20: (Gr + VC)/V0.02Mo0.98S2 | −0.234 | 0.462 |
HST8: Gr/V0.04Mo0.96S2 | −0.229 | 0.389 | HST21: (Gr + VC)/V0.0Mo0.96S2 | −0.499 | 0.407 |
HST9: Gr/(V0.02Mo0.98S2 + VS) | −0.288 | 0.535 | HST22: (Gr + VC)/(V0.02Mo0.98S2 + VS) | −0.306 | 0.512 |
HST10: Gr/(V0.02Mo0.98S2 + VS–S) | −0.351 | 0.723 | HST23: (Gr + VC)/(V0.02Mo0.98S2 + VS–S) | −0.373 | 0.723 |
HST11: Gr/(V0.02Mo0.98S2 + V2S) | −0.346 | 0.714 | HST24: (Gr + VC)/(V0.02Mo0.98S2 + V2S) | −0.364 | 0.695 |
HST12: Gr/(V0.02Mo0.98S2 + VMo) | −0.234 | 0.631 | HST25: (Gr + VC)/(V0.02Mo0.98S2 + VMo) | −0.184 | 0.480 |
HST13: Gr/(V0.02Mo0.98S2 + VMo+2S) | −0.292 | 0.800 | HST26: (Gr + VC)/(V0.02Mo0.98S2 + VMo+2S) | −0.306 | 0.809 |
Table 1 shows the lattice mismatch of each HST by giving the strain each layer experiences in the supercell after full relaxation. It appears that upon the formation of the different HSTs, the Gr layer is shrunk while the TMDC layer is expanded, but in all cases the effect is less than a percent. For example, V0.02Mo0.98S2 sheet in Gr/V0.02Mo0.98S2 HST is stretched by 0.426%, whereas the Gr sheet is compressed by 0.247% from the optimized pristine cell parameters. V0.02Mo0.98S2 sheet in (Gr + VC)/V0.02Mo0.98S2 HST is stretched by 0.462%, while the Gr sheet is compressed by 0.234% from the optimized parameters of the pristine cell. These rather small values for the built-in stress (less than 1% overall) do not change the properties of the monolayers and their HSTs, in agreement with other similar systems.23
To gain some insight into the interface properties of the HSTs, we compute the interlayer binding energy using the relation Eb = (EHST – EL1 – EL2)/N, where EHST is the total energy of each HST, EL1,L2 are the total energies for the individual monolayers, and N is the number of atoms in the system. Fig. 2 displays Eb versus the interlayer separation d for all systems. The results show that the binding energy is in the range Eb = (−30.152, −26.772) meV per atom, while the interlayer separation varies in d = (2.885, 3.334) Å range. The stability of all HSTs is reflected in the negative values of Eb, and we find that those systems containing MoS2 with a S monovacancy or divacancies have the highest binding energy. On the other hand, the lowest binding energy is found for systems containing MoS2 with a Mo + 2S vacancy complex or V dopants, indicating that such HSTs have a better structural stability compared to the rest of the systems. These Eb values are typical for HSTs in which van der Waals (vdW) interactions play an important role in their structural stability.24
![]() | ||
Fig. 2 Interlayer binding energy Eb (eV) as a function of the interlayer separation d (Å) of HSTs: (a) Gr/TMDCs and (b) (Gr + VC)/TMDCs. |
The interlayer separation in all cases is also in the typical vdW range, since for all systems d is found to be greater than the sum of the covalent radii of C and S atoms (0.76 and 1.05 Å, respectively). For Gr/V0.04Mo0.96S2 HST, d is very similar to the case of the pristine Gr/V0.02Mo0.98S2 HST. However, it appears that d between Gr and defective V0.02Mo0.98S2 monolayers is smaller as compared to the one for the pristine Gr/V0.02Mo0.98S2 HST. This effect becomes more pronounced for the defective Gr.
Let us first examine the isolated MoS2 monolayer. From the calculations, we find that isolated S monovacancy and divacancies, and Mo monovacancy do not induce magnetic ordering in the isolated TMDC monolayer. This agrees with previous studies on defective and single-doped MoS2.25 The total magnetic moment is 2 μB for MoS2 + VMo+2S. On the other hand, Mtot = 1 μB for a V-doped MoS2 with or without S monovacancy, S divacancies, Mo monovacancy (V0.02Mo0.98S2, V0.02Mo0.98S2 + VS, V0.02Mo0.98S2 + VS–S, V0.02Mo0.98S2 + V2S, V0.02Mo0.98S2 + VMo). Increasing of the doping concentration raises Mtot to 2 μB for V0.04Mo0.96S2, while Mtot = 3 μB for V0.02Mo0.98S2 + VMo+2S. Such synergistic effects have been observed experimentally.26
This situation changes upon the HST formation by adding a Gr layer. In fact, Fig. 3 shows that there is a tendency of reducing the total magnetization when a pristine Gr is stacked above a MoS2 with mono or divacancies despite their V dopants. It is interesting to observe that Mtot = 1 μB of the stand-alone V0.02Mo0.98S2 becomes Mtot = 0 for Gr/V0.02Mo0.98S2. In other words, a ferromagnetic standalone MoS2 with a 2% V doping becomes nonmagnetic upon bringing Gr together (HST7). Fig. 3 also shows that this trend holds for Gr/(V0.02Mo0.98S2 + VS), Gr/(V0.02Mo0.98S2 + VS–S), Gr/(V0.02Mo0.98S2 + V2S) (HST9, 10, 11), thus S mono and divacancies in the 2% doped TMDC do not affect the overall lack of magnetic moment in the HSTs. Moreover, Mtot = 2 μB of the stand-alone V0.04Mo0.96S2 becomes Mtot = 0.736 μB for Gr/V0.04Mo0.96S2. This reduction in the total magnetic moment is directly associated with charge transfer from Gr to V-doped MoS2 layer regardless of point defect presence, as discussed below. Nevertheless, there are situations where the Gr layer enhances the FM properties of the standalone TMDC. Such is the case of a V-doped MoS2 with a Mo monovacancy or a Mo + 2S complex defect. The Gr monolayer enhances the Mtot by about 88% and 33% when compared to the isolated V0.02Mo0.98S2 + VMo or V0.02Mo0.98S2 + VMo+2S. Fig. 3b and c show that in all HSTs containing a pristine Gr monolayer, the magnetic properties are carried out by the TMDC monolayer, and Gr remains nonmagnetic.
The magnetic properties of the HSTs with a defective Gr layer are overall greatly enhanced compared with those of the standalone TMDC layer and the HST with the pristine Gr layer. We find that the total magnetic moment of Gr with C monovacancy is 1.271 μB. The FM state of (Gr + VC) agrees with the well-known Lieb theorem that C monovacancy induces a magnetic moment on Gr while other defects do not.27,28 We further find that (Gr + VC)/TMDC with the S point defects shows a prominent ferromagnetism with Mtot > 1 μB unlike the nonmagnetic TMDC monolayer or the weakly ferromagnetic Gr/TMDC system. This enhancement effect is even stronger for the (Gr + VC)/(MoS2 + VMo) and (Gr + VC)/(MoS2 + VMo+2S). The magnetic moments for (Gr + VC)/V0.02Mo0.98S2 and (Gr + VC)/V0.04Mo0.96S2 are 1.157 and 1.513 μB, respectively. It is also interesting to note that the defective Gr has elevated Mtot for its HSTs formed with a doped TMDC with (Gr + VC)/(V0.02Mo0.98S2 + VMo) and (Gr + VC)/(V0.02Mo0.98S2 + VMo+2S) defects, for which Gr/TMDC showed weak ferromagnetism. It appears that in such cases, the magnetization is carried out primarily by the defective Gr layer (Fig. 3b and c). Introducing V dopants, however, enhances the magnetization of the TMDC layer itself, while Gr + VC remains ferromagnetic as well. Thus, these results show that the C vacancy plays a key role in the overall magnetic properties of the entire HST because it enhances the magnetization of the Gr layer itself. Such enhancement due to the ferromagnetism of the defective Gr layer is consistent with the enhanced magnetism found in other HSTs involving TMDC monolayers and ferromagnetic layers.15,29,30
Further insight of the ferromagnetism of these systems can be gained by examining the electron spin density property. In Fig. 4, we show some representative cases by giving the electron spin density distribution with TMDC on top of the Gr layer. Gr/V0.02Mo0.98S2 is non-magnetic and no spin density appears between atoms in this system (Fig. 4a). The nearest Mo-site spins are antiferromagnetically coupled to the V-site spin, while the nearest S-site spins are ferromagnetically coupled to the V-site spin in V0.04Mo0.96S2 monolayer. However, both the nearest Mo and S-site spins couple antiferromagnetically with the V-site spin in V0.04Mo0.98S2 layer of Gr/V0.04Mo0.96S2 HST (Fig. 4b). Although the Gr/(V0.02Mo0.98S2 + VS–S) HST has a total magnetic moment of zero, the antiferromagnetic orders appeared around the V atom (0.807 μB) in the TMDC layer (Fig. 4c). In V0.02Mo0.98S2 + VMo+2S sheet of Gr/(V0.02Mo0.98S2 + VMo+2S) HST, the antiferromagnetic orders appeared on the nearest Mo and S atoms around V atom, while the ferromagnetic orders appeared on the nearest Mo and S atoms around vacancy. In this case, we can clearly see that the spin density is located not only on the atoms but also on the empty spaces (Fig. 4d). Spin density distributions are not found in the Gr sheets of these HSTs, which are consistent with the values shown in Fig. 3b. For defective Gr sheet of (Gr + VC)/TMDC HSTs, the C monovacancy produces two pentagons while the FM order is observed in one pentagon and the AFM order is observed in another pentagon (Fig. 4e–h). In (Gr + VC)/V0.02Mo0.98S2 HST, no spin density is observed in V0.02Mo0.98S2 component (Fig. 4e). In (Gr + VC)/V0.04Mo0.96S2 HST, both V atoms have negative magnetic moments, but only one AFM order is observed for the nearest Mo and S atoms (Fig. 4f). Interestingly, spin density distributions of V0.02Mo0.98S2 + VS–S or V0.02Mo0.98S2 + VMo+2S are almost unchanged in HSTs with the pristine or defective Gr sheets (Fig. 4g and h).
We also examined the magnetic moments of V0.02Mo0.98S2 with various vacancies placed on Gr with C divacancies or SW defect and found that the magnetic phenomena in these systems are similar to those observed in pristine Gr. This is because the three systems pristine Gr, Gr + V2C and Gr + VSW are nonmagnetic.
![]() | ||
Fig. 5 Band structures of (a) Gr/MoS2, (b) Gr/V0.02Mo0.98S2, (c) Gr/V0.04Mo0.96S2 and (d) Gr/(V0.02Mo0.98S2 + VMo+2S) HSTs where the dotted lines represent the Fermi level. |
Fig. 5b shows that V doping is responsible for the electronic state of Gr/V0.02Mo0.98S2 HST becoming metallic. There is an upward shift of the electron states occupied in the valence band of the TMDC layer compared to that of the pristine HST, confirming the electron depletion from the Gr to the V0.02Mo0.98S2 sheet. Increasing the concentration of V doping causes the electron states occupied in the valence band to continue to shift upward, resulting in an impurity state near the Fermi level (Fig. 5c). For Gr/V0.02Mo0.98S2 with vacancies, the defective states appear mainly in the gap region of the V0.02Mo0.98S2 system (Fig. 5d).
The band structures of the defective Gr in (Gr + VC)/TMDC HSTs show that the bands for both spin carriers cross the Fermi energy (near the Γ point) (Fig. 6). Similarly to Gr, the band structures of Gr + VC in these HSTs show upward shifts of the Dirac points. The position of the Dirac points above the Fermi level reflects the number of electrons transferred from the Gr + VC to the TMDC monolayer. A notable difference between the band structures of Gr + VC in these HSTs is the different locations of the σ bands around the energy range of −1 to −0.5 eV. The band structures of the TMDC components in (Gr + VC)/TMDC HSTs are almost unchanged compared to those in Gr/TMDC HSTs.
Upon the introduction of the C monovacancy in Gr as part of each HST, one of linear π–π* crossing bands breaks its degeneracy turning it into a quadratically dispersing one (Fig. 6). The spin-splitting gap between these bands depends on the V doping and vacancy concentrations in the TMDC layer. In particular, the introduction of V dopants at 2% concentration (Fig. 6b) increases the spin splitting gap while 4% concentration of V dopants (Fig. 6c) reduces the spin splitting gap when compared to the pristine TMDC (Fig. 6a). In the (Gr + VC)/V0.02Mo0.98S2 HST (Fig. 6b), the Fermi level crosses the spin-up π band but lies completely below the entire spin-down π band, which causes a slight increase in the defective Gr magnetism. On the other hand, in the (Gr + VC)/V0.04Mo0.96S2 HST (Fig. 6c), the Fermi level lies completely below both spin-up and -down π bands and leads to a small reduction in the defective Gr magnetism. Furthermore, when (Mo + 2S) vacancies are present in the V dopants with 2% concentration (Fig. 6d), the splitting of π and π* bands is similar to that in defect-free HST (Fig. 6b). However, the (Mo + 2S) vacancies in TMDC layer reverses the spin polarizations of the π and π* bands. This reversal reduces the magnetism in the defective Gr layer in (Gr + VC)/(V0.02Mo0.98S2 + VMo+2S) HST compared to the corresponding defect-free (Gr + VC)/V0.02Mo0.98S2 HST.
The charge transfer in each HST is closely related to its electronic structure; thus, we also calculate the 3D charge density differences in each case. The results are shown in Fig. 7. We find that generally electrons are transferred from Gr or (Gr + VC) to the TMDC monolayer. No significant charge transfer is observed between Gr and MoS2 (Fig. 7a), which is also found experimentally.31 Significant charge transfer is observed between the Gr and V0.02Mo0.98S2 (Fig. 7b). The creation of vacancies or the increase in the doping concentration of V also significantly increases the charge transfer between the Gr and the TMDC layers (Fig. 7c and d). This is consistent with previous theoretical and experimental papers in other HSTs involving V-doped TMDCs (Fe/W0.967V0.033Se2/Pt, Bi2O2Se/WS2–V).29,32 The charge transfers between the defective Gr and TMDCs are comparatively higher than those observed between the pristine Gr and TMDCs (Fig. 7e–h). We can see that a significant portion of the charge density is localized on the C atoms near the defect sites, which results in a stronger interaction between the defective Gr and TMDC HSTs.
Notably, when the V-doped MoS2 monolayer is stacked with graphene, the MS increases dramatically in the HST compared to the V-doped monolayer alone (Fig. 8a and b). At room temperature MS increases from ∼5.67 μemu cm−2 in the doped TMDC to ∼55.83 μemu cm−2 in the HST. As shown in Fig. 8c, this enhancement persists across the full temperature range of 10–300 K. Moreover, the MS of the V-doped MoS2/Gr system (Fig. 8a) is nearly doubled of the undoped MoS2/Gr (Fig. 8d). Compared to the isolated MoS2 or V-doped MoS2 monolayers, the increase in magnetization upon interfacing with graphene in our HSTs is also consistent with reports on TMDC/Gr systems.15,16 For example, Cai et al.16 suggest that covalent interactions at the MoS2 Moiré superlattice/graphene oxide interface, such as the formation of Mo–S–C bonds, lead to spin polarization of Mo 4d electrons near the Fermi level, thereby inducing high-temperature FM ordering in the HST. However, such interfacial effects alone cannot explain the substantial MS enhancement observed in our MoS2/Gr systems (Fig. 8(a and d)). The study15 has demonstrated that vacancy defects play a critical role in inducing ferromagnetic ordering in the MoS2 monolayer as part of Mn-doped MoS2/Gr systems.
A comparison between Fig. 8a and d suggests that the graphene interfacing effect contributes more to the overall magnetization than V-doping alone. The combined effect of V-doping and MoS2/Gr interfacing results in the highest observed MS in the V-doped MoS2/Gr HST (see the inset of Fig. 8c), surpassing both the V-doped MoS2 monolayer and the MoS2/Gr HST. These experimental observations are fully supported by our DFT simulations, which also predict that the presence of (C, S, Mo) vacancies in both graphene and TMDC layers can enhance the overall magnetization, particularly when these defects are close and located near the interface. Electron transfer from MoS2 to graphene further strengthens the FM coupling between V magnetic ions via p–d orbital hybridization (Fig. 7b), contributing to the observed magnetization enhancement. Compared to the undoped HST, the charge transfer effect appears significantly stronger in the V-doped MoS2/Gr system, especially in those with VMo and VMo+2S vacancies. Our combined theoretical and experimental findings not only validate the magnetization enhancement predicted for Mn-doped MoS2/Gr heterostructures,15 but also provide new insights into the underlying physical mechanisms driving FM ordering in the 2D TMDC/Gr systems adding new functionalities to 2D van der Waals heterostructure-based nanodevices.35,36
We further find that Gr/MoS2 is nonmagnetic when doped with 2% V and becomes ferromagnetic with a total magnetic moment of 3 μB per cell as the V doping concentration increases. The V-doped MoS2/Gr HST with 4% V doping experiences stronger charge transfer as well as an upshift of the Dirac point. The effects of various defects, such as C monovacancy and Mo, S vacancies on the structural, electronic, and magnetic properties of V-doped MoS2/Gr HSTs with 2% V doping are also studied. The overall geometry and stability of HSTs change insignificantly in the presence of defects. However, the defects in both Gr and V-doped MoS2 control the magnetic properties of the resulting HSTs. Pristine Gr remains nonmagnetic in all studied HSTs, while defective Gr displays strong magnetic properties. On the other hand, S mono and divacancies have secondary effects on the overall magnetization in the Gr/TMDC and (Gr + VC)/TMDC HSTs, respectively. In contrast, the Mo monovacancy and (Mo + 2S) complex defect in the 2% doped TMDC noticeably affect the overall magnetization in the Gr/TMDC and (Gr + VC)/TMDC HSTs, respectively.
This study sheds light on the complex magnetism in sharp interfaced HSTs composed of graphene and a TMDC monolayer when doping and defects are taken into account. Tuning ferromagnetism in 2D systems by varying doping concentration and defect engineering is an attractive pathway for targeted applications in spintronics and other areas.
The structural properties of both pristine and V-doped TMDC monolayers, as well as their corresponding TMDC/graphene HSTs, were characterized using scanning transmission electron microscopy (STEM) and atomic force microscopy (AFM). The optical responses were examined via Raman and photoluminescence (PL) spectroscopy. Magnetic properties were measured using a vibrating sample magnetometry (VSM) probe integrated into a physical property measurement system (PPMS), under magnetic fields up to 9 T and across abroad temperature range from 2 K to 400 K.
Footnote |
† Current address: Department of Physics and Astronomy, George Mason University, Fairfax, Virginia 22030, USA. |
This journal is © The Royal Society of Chemistry 2025 |