Cuifang Jia,
Bo Zhou*,
Qi Song,
Xiaodong Zhang and
Zhenyi Jiang
Institute of Modern Physics, Shaanxi Key Laboratory for Theoretical Physics Frontiers, Northwest University, Xi'an 710069, People's Republic of China. E-mail: zhoubo@nwu.edu.cn
First published on 23rd May 2018
In this work, density functional theory is adopted to study the electronic and magnetic properties of MoS2 monolayers combined with a single S vacancy defect and a group VIII (G8) atom dopant, in which the dopant is incorporated via Mo substitution. The calculated results show that the magnetic properties of monolayer MoS2 can be tuned by changing the distribution of the G8 atom and S vacancy. The S vacancy tends to decrease the net magnetic moment of the doped system when these two defects are in their closest configuration. By adjusting the distance between the dopant and the S vacancy, the doped MoS2 monolayer may show a variable net magnetic moment. In particular, all of the Ni-doped MoS2 monolayers show zero magnetic moment with or without an S vacancy. The mean-field approximation is used to estimate the Curie temperature (TC). Our results show that Fe, Co, Ru, Rh, Os and Ir-doped MoS2 monolayers are potential candidates for ferromagnetism above room temperature. The density of states calculations provide further explanations as to the magnetic behavior of these doped systems. These results provide a new route for the potential application of atomically thin dilute magnetic semiconductors in spintronic devices by employing monolayer MoS2.
Recently, many studies, theoretical and experimental, have focused on the magnetic properties of 1H-MoS2.11–13 Magnetic doping through either adsorption or substitution is found to be an efficient method to introduce magnetism into MoS2.5–7 Magnetic interactions can be tuned by carriers and strain. At the same time, sulfur vacancies have also been found to be related to the magnetic properties of MoS2.14 Vacancies not only influence the magnetic properties, but also change the carrier density of the semiconductor. Obviously, S vacancies can influence the magnetism in the transition metal doping case. Lots of studies have been done on each factor. However, it is not clear how these two factors interplay and then what the influence of this interplay is on the magnetic properties.
In this paper, theoretical methods are used to study the effect of substitutional group VIII (G8) atom doping in 1H-MoS2 with or without an S vacancy. The distributions of the S vacancy and G8 elements have also been considered. This work may bring a new insight into the preparation of ferromagnetic materials in DMSs.
(1) |
Doped-atom | Configuration | dRMSD (Å) | Eform (eV) | Mtot (μB) | Q (e) | Eg (eV) | |
---|---|---|---|---|---|---|---|
Mo-rich | S-rich | ||||||
Fe | Mo15FeS32 | 0.121 | 2.64 | −0.06 | 2.00 | 6.95 | 0.30 |
Mo15FeS31-A | 0.152 | 6.61 | 5.26 | 0 | 7.11 | 0.33 | |
Mo15FeS31-B | 0.117 | 7.79 | 6.44 | 2.00 | 6.99 | 0.19 | |
Mo15FeS31-C | 0.113 | 7.80 | 6.45 | 2.00 | 7.00 | 0.27 | |
Mo15FeS31-D | 0.118 | 7.76 | 6.41 | 2.00 | 6.98 | 0.37 | |
Mo15FeS31-E | 0.119 | 7.82 | 6.47 | 2.00 | 6.98 | 0.27 | |
Co | Mo15CoS32 | 0.120 | 3.91 | 1.21 | 3.00 | 8.23 | 0.18 |
Mo15CoS31-A | 0.169 | 7.19 | 5.84 | 1.00 | 8.32 | 0.21 | |
Mo15CoS31-B | 0.122 | 9.05 | 7.70 | 3.00 | 8.21 | 0.24 | |
Mo15CoS31-C | 0.444 | 8.66 | 7.29 | 1.00 | 8.29 | 0.28 | |
Mo15CoS31-D | 0.134 | 8.97 | 7.68 | 1.00 | 8.24 | 0.22 | |
Mo15CoS31-E | 0.118 | 9.09 | 7.74 | 3.00 | 8.22 | 0.11 | |
Ni | Mo15NiS32 | 0.029 | 4.72 | 2.22 | 0 | 9.35 | 0.28 |
Mo15NiS31-A | 0.175 | 7.82 | 6.47 | 0 | 9.37 | 0.43 | |
Mo15NiS31-B | 0.034 | 9.75 | 8.74 | 0 | 9.37 | 0.15 | |
Mo15NiS31-C | 0.030 | 9.57 | 8.57 | 0 | 9.36 | 0.30 | |
Mo15NiS31-D | 0.411 | 9.70 | 8.35 | 0 | 9.35 | 0.27 | |
Mo15NiS31-E | 0.031 | 9.88 | 8.74 | 0 | 9.35 | 0.27 | |
Ru | Mo15RuS32 | 0.032 | 3.05 | 0.35 | 2.00 | 7.14 | 0.17 |
Mo15RuS31-A | 0.067 | 6.54 | 5.19 | 0 | 7.32 | 0.57 | |
Mo15RuS31-B | 0.039 | 7.90 | 6.60 | 0 | 7.19 | 0.34 | |
Mo15RuS31-C | 0.037 | 7.96 | 6.61 | 0 | 7.20 | 0.26 | |
Mo15RuS31-D | 0.046 | 7.92 | 6.57 | 0 | 7.18 | 0.38 | |
Mo15RuS31-E | 0.048 | 8.04 | 6.69 | 0 | 7.18 | 0.23 | |
Rh | Mo15RhS32 | 0.033 | 4.17 | 1.47 | 1.00 | 8.39 | 0.03 |
Mo15RhS31-A | 0.060 | 7.22 | 5.87 | 1.00 | 8.52 | 0.19 | |
Mo15RhS31-B | 0.293 | 8.98 | 7.63 | 1.00 | 8.47 | 0.28 | |
Mo15RhS31-C | 0.415 | 8.72 | 7.23 | 1.00 | 8.49 | 0.20 | |
Mo15RhS31-D | 0.038 | 9.09 | 7.74 | 1.00 | 8.40 | 0.15 | |
Mo15RhS31-E | 0.035 | 9.21 | 7.86 | 1.00 | 8.41 | 0.10 | |
Pd | Mo15PdS32 | 0.298 | 5.03 | 2.33 | 0 | 9.58 | 0.30 |
Mo15PdS31-A | 0.058 | 8.28 | 6.93 | 0 | 9.60 | 0.49 | |
Mo15PdS31-B | 0.347 | 10.08 | 8.73 | 2.00 | 9.59 | 0.19 | |
Mo15PdS31-C | 0.355 | 9.86 | 8.51 | 0 | 9.59 | 0.42 | |
Mo15PdS31-D | 0.356 | 10.08 | 8.73 | 0 | 9.58 | 0.28 | |
Mo15PdS31-E | 0.358 | 10.12 | 8.77 | 0 | 9.57 | 0.27 | |
Os | Mo15OsS32 | 0.031 | 3.59 | 0.89 | 2.00 | 7.06 | 0.12 |
Mo15OsS31-A | 0.065 | 7.08 | 5.73 | 0 | 7.30 | 0.58 | |
Mo15OsS31-B | 0.035 | 8.26 | 6.91 | 0 | 7.14 | 0.40 | |
Mo15OsS31-C | 0.026 | 8.43 | 7.08 | 2.00 | 7.15 | 0.21 | |
Mo15OsS31-D | 0.043 | 8.36 | 7.01 | 0 | 7.11 | 0.35 | |
Mo15OsS31-E | 0.044 | 8.50 | 7.15 | 0 | 7.11 | 0.16 | |
Ir | Mo15IrS32 | 0.032 | 4.38 | 1.68 | 1.00 | 8.44 | — |
Mo15IrS31-A | 0.059 | 7.46 | 6.11 | 1.00 | 8.60 | 0.23 | |
Mo15IrS31-B | 0.048 | 9.21 | 7.86 | 1.00 | 8.48 | 0.35 | |
Mo15IrS31-C | 0.032 | 9.19 | 7.84 | 1.00 | 8.50 | 0.19 | |
Mo15IrS31-D | 0.038 | 9.22 | 7.87 | 1.00 | 8.46 | 0.11 | |
Mo15IrS31-E | 0.040 | 9.36 | 8.01 | 1.00 | 8.47 | 0.09 | |
Pt | Mo15PtS32 | 0.322 | 4.98 | 2.28 | 0 | 9.68 | 0.32 |
Mo15PtS31-A | 0.055 | 8.14 | 6.79 | 0 | 9.72 | 0.58 | |
Mo15PtS31-B | 0.373 | 9.85 | 8.50 | 0 | 9.70 | 0.36 | |
Mo15PtS31-C | 0.372 | 9.77 | 8.42 | 0 | 9.71 | 0.45 | |
Mo15PtS31-D | 0.392 | 9.93 | 8.58 | 0 | 9.68 | 0.27 | |
Mo15PtS31-E | 0.393 | 9.95 | 8.60 | 0 | 9.67 | 0.28 |
In order to inspect the stability and feasibility of all optimized geometrical structures, the formation energies were obtained utilizing the following formula:23–25
Eform = Edoped − Epure + (μMo − μG8) + μS | (2) |
Eform(MoS2) = μMoS2 − μ0Mo − 2μ0S | (3) |
The values of μMo and μS in eqn (2) depend on the experimental growth conditions. For the Mo-rich case, the Mo chemical potential is equal to the bulk Mo value, μMo-richMo = μ0Mo, and the S chemical potential can be obtained from μMoS2 = μMo + 2μS on the basis of thermodynamic equilibrium. Hence, combined with eqn (3), the chemical potentials for the Mo-rich limit can then be written as:
μMo-richMo = μ0Mo, | (4) |
(5) |
Likewise, under S-rich conditions, the values are:
μS-richMo = μ0Mo + Eform(MoS2), | (6) |
μS-richS = μ0S. | (7) |
All of the formation energies are listed in Table 1.
Monolayer MoS2 with a single S vacancy (VS) or Mo vacancy (VMo) was fully relaxed. The optimized structure shows that the neighboring Mo and S atoms have slight displacements with respect to the vacancy site VS and VMo, which is different from the obvious reconstruction in a graphene sheet with a single C vacancy.27 For VS-MoS2, this is more likely to occur under S-rich conditions and the formation energy is 6.57 eV, which in agreement with the previous values of 5.89 (ref. 28) and 5.72 eV,7 is much lower than the formation energy of the Mo vacancy VMo (14.09 eV) in Mo-rich conditions. The previous studies have reported that the substitution of an Mo site is more stable than that of an S site.7,29 Experimentally, S vacancies are more common than Mo vacancies.30 Therefore, in this paper, the co-doped configurations mainly consist of an S vacancy and G8 impurity substitution of an Mo site. The C3v symmetry of pristine monolayer MoS2 is destroyed after G8 element doping, and the distances between the impurities and the nearest S atom change with different distributions of the dopants and the vacancy. All the calculated formation energies are summarized in Table 1. The distances between the dopants atoms and the S vacancy are taken from their original positions in pristine 1H-MoS2. Fig. 2 shows our calculated formation energies as a function of the distance between the S vacancy and dopant atoms, which varies from 2.42 to 7.53 Å. The first data “0” represents the doped MoS2 without an S vacancy. According to our theoretical results, Fe doping is most favorable energetically among all considered impurities. For all of the G8 elements, the formation energies from the doped defective 1H-MoS2 configurations of the second-, third-, fourth-, and fifth-nearest neighbor are very close, and are 1.50 eV larger than the corresponding nearest neighboring case. The relevant data for the Co atom are consistent with ref. 31.
Fig. 2 Formation energies as a function of the distance between the S vacancy and impurity atoms in S-rich and Mo-rich conditions. |
The positive formation energy indicates that the formation of the vacancy defect and the doping of the transition metal are endothermic processes. The equilibrium concentrations of the vacancies are usually very low because of their high formation energies. Nonetheless, new techniques have been developed to create these defects. The generation of nanomesh size vacancies in graphene has been reported.32,33 Vacancy engineering of a doped MoS2 monolayer can also be achieved.
Fig. 3 Ligand field picture and corresponding d-orbitals for a trigonal prismatic coordination (D3h symmetry) in a d-metal dichalcogenide with a d2 configuration of the metal atom. |
The Fe case, which corresponds to the d4 configuration, is likely to have a net magnetic moment of 2μB, except for the A case. In the nearest configuration, the interaction between the Fe and the S vacancy (the unsaturated Mo atoms) leads to strong crystal field splitting between the dz2, dxy, and dx2−y2 orbitals, which causes the system to become a small band gap semiconductor. Among all of the five pristine dopant-vacancy configurations, the B and E cases have more symmetry and the vertical mirror plane is preserved. It is important for the system to keep the energy levels of the hybridized states below the Fermi level in the close energy range. This can explain the net magnetic moment of the B and E cases of Fe, and Co. On the other side, the C and D cases are more likely to have zero magnetic moment. Our results also indicate that local stress appears to be a crucial factor in the development of magnetism as Jahn–Teller distortions that destroy the C3v lattice symmetry lead to the disappearance of magnetism. For the Fe and Co cases, the C and D configurations may also have a net magnetic moment because of the coupling between the defective d orbital of the unsaturated Mo atoms and the lower dxy, dx2−y2, and dz2 orbitals of the impurity atom.
For 4d or 5d dopants such as the Ru-doped cases, only the structure of mono-doped Mo15RuS31 generates a 2μB magnetic moment. The origin of the magnetism is attributed to the near degeneracy of the Ru dx2−y2 and dz2 orbitals. The Os-doping has similar characteristics to the Ru doped cases. For the Pd-doping, only the Mo15PdS31-B structure generates a 2μB magnetic moment. For the Rh- and Ir-doping, all of the structures generate a 1μB magnetic moment. For the Pt-doping structures, all are nonmagnetic. Taken as a whole, the net magnetic moments from the 3d elements are larger than those from the 4d and 5d elements, because strongly localized 3d orbitals are more likely to induce net magnetic moments due to strong Hund coupling, which competes with the ligand field energy splitting.
Furthermore, the ligand field energy splitting is connected with the distortion of the local structure around the dopants. The root-mean-square deviations of the X–S bonds have been obtained to quantify the local deformation caused by the dopants and are listed in Table 1. The results clearly show that the RMSDs of the models with smaller magnetic moments are much larger, which means a larger deviation from C3v symmetry. This rule holds for the Fe, Co and Os cases.
To verify our results, the DFT+U method was also used. It was found that the magnetic properties with the GGA+U method are consistent with the results without the Hubbard-U parameter. The results with the corresponding U values are shown in ESI Table S1.† There is no relevant influence on our conclusions and therefore we discuss in the following the results without on-site interaction.
The system is stabilized by charge transfer. The Bader charge has also been obtained to analyze the charge transfer between the G8 atoms and MoS2, as shown in Table 1. In a perfect 1H-MoS2, the formal valences of Mo and S are +4 and −2, respectively. Since a Mo has six S neighbors, it contributes 2/3 electrons to each Mo–S covalent bond. Therefore, the charge transfer of the S and Mo atoms is approximately +0.67e and −1.33e, respectively. Taking Fe-doped 1H-MoS2 as an example, in the monodoping case, the charge number of Fe will be 8 − 2 × 2/3 = 6.67 (8 is the valence electron number of Fe treated by the PBE pseudopotential). It will be 8 − 5/3 × 2/3 = 6.89 when one S atom is removed. The theoretical charge numbers of Ru and Os are equal to that of Fe due to their similar valence electron numbers. For the other G8 elements, the two values for mono- and co-doping are 7.67 and 7.89 for Co, Rh and Ir, and 8.67 and 8.89 for Ni, Pd and Pt. The more the Bader charge deviates from the ideal charge number, the more electron transfer there is from the host ions to the dopant ion (Fig. 4).
In order to understand the characteristics of the magnetic moment induced by impurity states in the G8-doped 1H-MoS2 systems, the spin density distribution is plotted to visualize the distribution of the magnetic moments of the doped 1H-MoS2 systems in Fig. 5. Taking Fe-doped 1H-MoS2 as an example, Fig. 5(a)–(f) show that the spatial extensions of the spin polarizations have reached the first-nearest S atoms and the second-nearest Mo atoms. The structure Mo15FeS32 displays weak ferromagnetic and antiferromagnetic coupling between Fe and three neighboring S atoms, and ferromagnetic coupling among the six second-nearest Mo atoms. Fig. 5(b) shows the nearest configuration of the co-doping system, which is typical for the no magnetism case. The antiferromagnetic alignment of d electrons is observed.
Mo15XS32 | Mo15XS31-A | Mo15XS31-B | Mo15XS31-C | Mo15XS31-D | Mo15XS31-E | ||
---|---|---|---|---|---|---|---|
Fe | Δ | 0.209 | — | 0.138 | 0.037 | 0.053 | 0.175 |
TC | 1614 | — | 1106 | 284 | 410 | 1352 | |
Co | Δ | 0.110 | 0.060 | 0.114 | 0.072 | 0.060 | 0.027 |
TC | 850 | 462 | 882 | 556 | 464 | 208 | |
Ru | Δ | 0.125 | — | — | — | — | — |
TC | 966 | — | — | — | — | — | |
Rh | Δ | 0.026 | 0.099 | 0.070 | 0.050 | 0.039 | 0.026 |
TC | 200 | 766 | 542 | 386 | 302 | 200 | |
Pd | Δ | — | — | 0.010 | — | — | — |
TC | — | — | 78 | — | — | — | |
Os | Δ | 0.088 | — | — | 0.032 | — | — |
TC | 680 | — | — | 248 | — | — | |
Ir | Δ | 0.022 | 0.053 | 0.090 | 0.055 | 0.031 | 0.017 |
TC | 170 | 410 | 696 | 426 | 240 | 132 |
Fig. 6 Total DOS and partial DOS for pristine 1H-MoS2 with and without S vacancies. The dashed line indicates the Fermi level at 0 eV. |
The density of states of all the Fe-doped MoS2 models are shown in Fig. 7. It can be found that the impurity states presented in the band gap region are mainly contributed by the Mo 4d, Fe 3d, and S 3p orbitals. All of these systems are ferromagnetic with net magnetic moments of 2μB, except for the structure Mo15FeS31-A, which is a nonmagnetic semiconductor with a 0.3 eV band gap. In configuration A, the largest deformation of the Fe atom can be found from visualizing the optimized structure, which results in the removal of the degeneracy of the 3d orbitals and leads to the formation of the localized hydride states near the valance bands. As for the five magnetic systems, as expected, the Fe dopant is a main contributor to the total magnetic moment. This feature applies to all G8 impurity-doped configurations in this calculation.
Fig. 7 Total DOS and partial DOS for each atomic species in the all Fe-doped 1H-MoS2. The dashed line indicates the Fermi level at 0 eV. |
For the Co-doped cases, these systems have odd electrons. They are all ferromagnetic with a magnetic moment of 3 or 1μB. Unlike the Fe case, the correlation between the Co atom dopant and the S vacancy causes the impurity states to spread in the whole band gap region. Especially for the configurations Mo15CoS31-A and Mo15CoS31-D, the strong interaction between the Co–Mo–S pairs leads to the formation of a localized impurity state near the valance bands. The decrease in the magnetic moment can be attributed to the pairing of the electrons in this mixed orbital. In other words, the magnetic properties of the co-doped system depend on the competition between the ligand field splitting and the Hund coupling (Fig. 8).
For Ni-doped 1H-MoS2 systems, there are 4 unpaired electrons remaining, which would lead to about a 4μB magnetic moment. But from our calculations, all of the ground states of the Ni-doped models are nonmagnetic semiconductors where the spin-up and spin-down channels are completely symmetrical and they possess tiny band gaps of 0.28, 0.43, 0.15, 0.30, 0.27, and 0.27 eV, respectively. The strong crystal field of neighboring Mo and S atoms leads to large energy splitting of the defective orbital. Therefore, four outer electrons form two electron pairs without the existence of isolated electrons. During our calculations, we found that the Mo15NiS32 and Mo15NiS31-B, C and E configurations would relax to transition states that have high magnetic order (net magnetic moment 4μB). These transition states are local minima which are 199, 341, 355, and 208 meV above the global minima. In a recent experimental study,9 the author reported that 4% Ni doped MoS2 has a paramagnetic phase at room temperature, and the paramagnetic phase may dominate at low temperature. Our theoretical results of the transition states with an S vacancy provide an explanation for this phenomenon (Fig. 9).
For the Ru-doped cases, substitution of Mo atoms by Ru increases the degree of p–d hybridization leading to a shift of the majority spin below the Fermi level. All configurations with the existence of S vacancies have semiconductor character with a band gap shown in Table 1, while the structure Mo15RuS32 is a magnetic semiconductor with a net magnetic moment of 2μB. The magnetism of Mo15RuS32 can be attributed to the formation of degenerated bands mainly consisting of the Ru dx2−y2 and dz2, Mo 4d and S 3p orbitals (Fig. 10).
Rh-doping provides 3 more valence electrons than the host Mo atom and all structures generate a 1μB magnetic moment. In addition, the mono-doped structure is a half-metal and the others are magnetic semiconductors (Fig. 11).
For Pd-doped 1H-MoS2, the structure Mo15PdS31-B is a magnetic semiconductor with a 2μB magnetic moment, although Pd is a nonmagnetic element. The other systems are all nonmagnetic semiconductors and generate tiny band gaps of 0.30, 0.58, 0.21, 0.35, and 0.16 eV for Mo15PdS32, Mo15PdS31-A, Mo15PdS31-C, Mo15PdS31-D, and Mo15PdS31-E, respectively. The reason for the magnetism is the Hund coupling between the two hybridized states from the Mo 4d, S 3p and Pd 4d orbitals (Fig. 12).
Os is a 5d element with less localized d orbitals than the 3d and 4d elements discussed above. For the Os-doped 1H-MoS2, the impurity states are located near the conduction bands. The Mo15OsS32 and Mo15OsS31-C configurations are magnetic, and both have a net magnetic moment of 2 μB. In the Mo15OsS31-C configuration, the magnetism is attributed to the near degeneracy of two defective bands composed of the Co dx2−y2 and dz2 orbitals and the d orbital of the nearby Mo ions. The Mo15OsS31-A, Mo15OsS31-B, Mo15OsS31-D and Mo15OsS31-E configurations are nonmagnetic semiconductors with small band gaps of 0.57, 0.40, 0.35 and 0.16, respectively (Fig. 13).
For Ir-doping, the DOS are very similar to the cases of Rh-doping, since the Ir atom belongs to the same column of the periodic table and has the same valence electron configuration as Rh. The difference is that the structure Mo15IrS32 is unambiguously half-metallic and can be utilized as a spin filter. Similarly, all structures possess a 1μB net magnetic moment (Fig. 14).
For the Pt-doped cases, all of the structures are nonmagnetic semiconductors, i.e., they all have symmetrical spin-up and spin-down channels, and generate a series of energy gaps of 0.32, 0.58, 0.36, 0.45, 0.27, and 0.28 eV corresponding to Mo15PtS32, Mo15PtS31-A, Mo15PtS31-B, Mo15PtS31-C, Mo15PtS31-D, and Mo15PtS31-E, respectively. From the DOS figures, we can find that the d bands of the Pt ion mainly appear above the Fermi level and the strong hybridization of the S 3p and Pt 4d orbital, and the electrons tend to occupy the delocalized defective bands that are caused by the S vacancy (Fig. 15).
The magnetic properties of the co-doped system depend on the competition between the Hund coupling and the ligand field splitting. Our results suggest that the co-doping of G8 atoms and S vacancies is an efficient way to modulate the magnetic properties. The main obstacle is the control of the distribution. Our further work will focus on methods to decrease the formation energy of the defects and the maintenance of the magnetic moment.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c8ra01644e |
This journal is © The Royal Society of Chemistry 2018 |