Hui Wana,
Liang Xua,
Wei-Qing Huang*a,
Jia-Hui Zhoua,
Chao-Ni Hea,
Xiaofan Li*a,
Gui-Fang Huang*a,
P. Pengb and
Zheng-Gui Zhoua
aDepartment of Applied Physics, School of Physics and Electronics, Hunan University, Changsha 410082, China. E-mail: wqhuang@hnu.edu.cn; wqhuang2000@yahoo.com; yueyanglxf@hnu.edu.cn; gfhuang@hnu.edu.cn
bSchool of Materials Science and Engineering, Hunan University, Changsha 410082, China
First published on 5th December 2014
The monolayer MoS2, possessing an advantage over graphene in that it exhibits a band gap whose magnitude is appropriate for solar applications, has attracted increasing attention because of its possible use as a photocatalyst. Herein, we propose a codoping strategy to tune the band structure of monolayer MoS2 aimed at enhancing its photocatalytic activity using first-principles calculation. The monodoping (halogen element, Nd) introduces impurity states in the gap, thus decreasing the photocatalytic activity of MoS2. Interestingly, the NbMoFS codoping reduces the energy cost of doping as a consequence of the charge compensation between the niobium (p-dopant) and the fluorine (n-dopant) impurities, which eliminates the isolated levels (induced by monodopant) in the band gap. Most importantly, the NbMoFS codoped MoS2 has more active sites for photocatalysis. These results show the proposed NbMoFS codoped monolayer MoS2 is a promising photocatalyst or photosensitizer for visible light in the heterogeneous semiconductor systems.
MoS2 is formed by a graphene-like hexagonal arrangement of Mo and S atoms stacked together to give S–Mo–S sandwiches which are bonded together by weak van-der-Waals forces14 and coordinated in a triangular prismatic fashion. The electronic properties of bulk MoS2 show a strong dependence on the layer thickness.15–17 In this context, 1H-MoS2, one kind of transition-metal dichalcogenides, is of particular interest.17–25 Recently, two-dimensional (2D) 1H-MoS2 has been successfully synthesized,26–28 displaying unique properties. These findings have triggered many theoretical studies29–36 to tune its properties, such as decorating the pristine surface with adatoms or molecules to tune the electronic structure (also add catalytic sites) of 1H-MoS2. Among them, the photocatalytic performance of MoS2 has drawn particular attention due to its large surface area and catalytically active sites. Of critical importance, the band gap around 2.0 eV of 1H-MoS2 is preferable in order to utilize the maximum portion of the solar visible light, thus making it to be a promising catalyst in various fields including hydrodesulfurization (HDS), hydrogen evolution reaction (HER), hydrogen generation and photocatalytic degradation of organic pollutants.37,38 It is found that the catalytic activity of MoS2 plates is derived from the under coordinated sulfur edge sites, while their basal planes remain catalytically inert. Moreover, because of its layered nature and high surface energy of the two-dimensional structure, MoS2 nanomaterials are prone to stacking together through π–π interaction during the preparation process, making a substantial amount of catalytic edge sites on MoS2 blocked, and delivering higher resistance for the electron transfer and diffusion of reactant molecules, and retarding the catalytic reaction. Therefore, the effective strategy to enhance the photocatalytic performance of MoS2 by increasing edge catalytic sites or making their basal planes to be catalytic active is being actively pursued.
Introducing a small amount of foreign elements into the semiconductors with wide band gap is an effective method frequently used. For example, Komsa et al.39 demonstrated the vacancy formation in 1H-MoS2 and then the filling of vacancies with substitutional impurity atoms, verifying the above expectation and appears to be a promising way to extend the applications of MoS2. Actually, the doping strategy has been widely used in many other systems, such as, N-doped Cu2O,40 graphene,41 boron nitride42 and related structures.43 There are so far many reports available on specific dopants for 1H-MoS2.44–47 However, it should be pointed out that some major issues may arise from monodoping. It has widely been recognized that monodoping of semiconductors with other elements produces localized electronic midgap states associated with dopants, which is the main reason of deteriorating their properties, such as photocatalytic properties. For examples, these localized states can boost electron–hole recombination, thereby diminishing the photocatalytic efficiency. Another key issue, particularly in the case of charge imbalance, is the formation of vacancies, which are efficient charge carrier traps. To overcome these problems, we employ the codoping strategy, which has been found to be very successful in modifying the band structure of semiconductor photocatalysts.48–50 However, there are few studies reporting the significance of codoping in improving the photoactivity of 1H-MoS2.
In this work, we propose, for the first time, the charge compensated codoping of Nd and F to modify the band structure of 1H-MoS2. Basically, the introduction of Nd (4s2 4p6 4d4 5s1) atom in place of Mo (4s2 4p6 4d5 5s1) atom leads to one electron deficiency, which can be compensated by replacing one S (3s2 3p4) by F (2s2 3p5). Being very similar in atomic size, Nd can be easily fitted into the Mo lattice site without any large distortion in the crystal structure. Another reason for choosing Nd is due to that the Nd-doped 1H-MoS2 has been prepared successfully by vapor-phase reaction.51 We perform a detailed systematic calculation exploring the role of the codoping and individual dopant elements, as well as halogen element in place of S atom for comparison using density functional theory, which is a very popular tool for describing the electronic structures of semiconductor materials. It is found that monodoping will cause acceptor or donor levels in the gap region. The charge compensated dopants, Nd and F, exhibit strong interactions through significant lattice relaxation. The stability of the codoped system has been assessed by calculating the defect-pair binding energy. Because the impurity levels are passivated, they will not be effective as carrier recombination centers. Furthermore, the NbMoFS–1H-MoS2 has more active sites for photocatalytic reaction. This work is an example of heteroatom substitution in the case of 1H-MoS2, opening up a wide range of possibilities including varying the electronic and optical behaviors of the monolayer MoS2, in catalysis and in solar cells.
The first-principles calculations are performed using density-function theory (DFT) as implemented in the CASTEP code.52 The exchange–correlation interactions are treated by the generalized gradient approximation (GGA) with Perdew–Burke–Ernzerhof (PBE) functional,53 which are based on a total energy pseudo-potential plane wave method.54 The vanderbilt ultrasoft pseudopotential is used to deal with the interaction between valence electrons and the ionic core, and the valence atomic configurations are Mo: 4s2 4p6 4d5 5s1, S: 3s2 3p4, Nb: 4s2 4p6 4d4 5s1, F: 2s2 2p5, Cl: 3s2 3p5, Br: 4s2 4p5, and I: 5s2 5p5, respectively. The cutoff energy is set to 400 eV, the k-point mesh of 9 × 9 × 1 and the self-consistent field (SCF) of 1 × 10−6 eV per atom is used for geometry optimizations using the Monk horst–Pack scheme.55 Geometry optimization is carried out before single point energy calculation and the force on the atoms is less than 0.03 eV Å−1, the stress on the atoms is less than 0.05 GPa, the atomic displacement is less than 1.0 × 10−3 eV Å−1, and the energy change per atom is less than 1.0 × 10−5 eV. Periodic boundary conditions are applied and a vacuum layer of at least 10 Å is placed above the monolayer to minimize the interaction between the adjacent periodic images. To check the reliability of the results, a test calculation with higher plane-wave cutoff energy and more k-points is conducted. Compared with the cutoff energy of 400 eV, negligible changes are obtained for both structural and electronic structures and the difference between the total energies is less than 0.03%. So the calculations in the present work are valid.
The changes in lattice parameters and bond lengths of doped 1H-MoS2 are summarized in Table 1. One can see that the lattice constant change of NbMo–1H-MoS2 is the biggest of all. This is due to the length of Nb–S bond (2.455 Å) longer than that of Mo–S bond (2.409 Å) in pure 1H-MoS2 because the electronegativity of Nb atom (1.6) is smaller than that of Mo atom (2.16), although the radius of Nb (1.45) and Mo (1.39) atoms have a few differences. For the NMS–1H-MoS2 case, the lattice constant changes are different with increasing their atomic numbers. The optimized lattice parameter of FS–1H-MoS2 is decreased, owing mainly to that the larger electronegativity of F atom (3.98) leads to a shorter length of F–Mo bond (2.275 Å). While the lattice parameters of ClS–1H-MoS2, BrS–1H-MoS2, and IS–1H-MoS2 are increased due to the Cl–Mo, Br–Mo, and I–Mo bond lengths elongating to 2.517, 2.671, and 2.897 Å, respectively, which can largely be attributed to the much bigger atomic radius of Cl (1.62 Å), Br (1.15 Å), and I (1.4 Å) than S (1.04 Å) atom. Interestingly, the lattice parameters of codoping 1H-MoS2 have changed very little, only a reduction of 0.13% (9.530 Å vs. 9.542 Å). Two important factors are responsible for the small change in the lattice parameters of codoping 1H-MoS2. The first is the discrepancies in both the radius and the electronegativity of host atoms and substitutional atoms. The other factor directly impacting the lattice parameters of codoping 1H-MoS2 is the coupling between the substitutional metal and nonmetal atoms.
| No. | Modela | Structure parameter (Å) | Bond length (Å) | |||
|---|---|---|---|---|---|---|
| a(= b) | c | NM–Mo | M–S | NM–Mb | ||
| a In pure 1H-MoS2: a(= b) = 9.542 Å, c = 13.196 Å, the bond length Mo–S = 2.409 Å.b NM and M means the substitutional nonmetal atoms and metal atom, respectively. | ||||||
| (a) | NbMo | 9.631 | 13.159 | 2.455 | ||
| (b) | FS | 9.511 | 13.202 | 2.275 | ||
| (c) | ClS | 9.580 | 13.240 | 2.517 | ||
| (d) | BrS | 9.592 | 13.289 | 2.671 | ||
| (e) | IS | 9.598 | 13.392 | 2.879 | ||
| (f) | NbMoFS | 9.530 | 13.276 | 2.220 | 2.449 | 2.224 |
| 2.436(2) | ||||||
| 2.432(2) | ||||||
To understand the charge redistribution induced by Nb or NM dopant, we calculate the electron density difference for the doped 1H-MoS2 supercells, as shown in Fig. 2. The population of the selected atoms is also given by the numbers. The maps of pure phase are also plotted for comparison. It can be seen from parts (b) and (c) of Fig. 2 that when the Nb or F atom substitutes the lattice Mo or S atom, after charge redistribution, the interactions of Nb–S or Mo–F bond becomes stronger than the original Mo–S bonds, and the maps indicate that Nb atom lose charges and F atom withdraws charges from Mo atoms. Furthermore, Mulliken charge population analyses are calculated and listed in Table 2. We can find some useful information from the relative values of Mulliken population. The Nb atom is deprived of 0.34e, while the charge of F atom is −0.42e, capturing from the adjacent Mo cations, which are much more than that on the Mo atom (0.05e) and S atom (−0.02e) in the pure phase. Thus, we contribute the stronger Nb–S and Mo–F bonds to the more charge transfer from adjacent atoms. The smaller interaction of Cl–Mo, Br–Mo, and I–Mo (Fig. 2) is mainly due to the weak ability of Cl atom (−0.04e) to withdraw charge, while Br and I atoms deprive charges by adjacent Mo atom. The dopant–dopant coupling is significant in the NbMoFS–1H-MoS2, in which the substitutional Nb and F could form a strong Nb–F bond because F atom withdraws charges from Nb atom along with the direction of Nb–F bond, as shown in Fig. 2(g).
| Model | Species | s | p | d | Total | Charge (e) |
|---|---|---|---|---|---|---|
| Pure | S | 1.86 | 4.16 | — | 6.02 | −0.02 |
| Mo | 2.46 | 6.45 | 5.04 | 13.95 | 0.05 | |
| NbMo | Nb | 2.43 | 6.28 | 3.94 | 12.66 | 0.34 |
| S | 1.86 | 4.21 | — | 6.07 | −0.07 | |
| Mo | 2.46 | 6.44 | 5.03 | 13.93 | 0.07 | |
| FS | F | 1.96 | 5.46 | — | 7.42 | −0.42 |
| S | 1.86 | 4.15 | — | 6.01 | −0.01 | |
| Mo | 2.44 | 6.39 | 4.97 | 13.81 | 0.19 | |
| ClS | Cl | 1.93 | 5.11 | — | 7.04 | −0.04 |
| S | 1.86 | 4.19 | — | 6.05 | −0.05 | |
| Mo | 2.47 | 6.45 | 5.04 | 13.96 | 0.05 | |
| BrS | Br | 1.93 | 4.99 | — | 6.92 | 0.08 |
| S | 1.86 | 4.19 | — | 6.05 | −0.05 | |
| Mo | 2.46 | 6.47 | 5.05 | 13.98 | 0.02 | |
| IS | I | 1.93 | 4.81 | — | 6.73 | 0.27 |
| S | 1.86 | 4.17 | — | 6.04 | −0.04 | |
| Mo | 2.46 | 6.44 | 5.03 | 13.94 | 0.06 | |
| NbMoFS | F | 1.95 | 5.47 | — | 7.42 | −0.42 |
| S | 1.86 | 4.22 | — | 6.08 | −0.08 | |
| Nb | 2.41 | 6.24 | 3.87 | 12.52 | 0.48 | |
| Mo | 2.46 | 6.42 | 4.96 | 13.83 | 0.17 |
| ΔEform = Etot(α) − Etot(MoS2) − ∑nAμA + ∑nBμB + q(EVBM + εF + ΔV) | (1) |
| Eform(MoS2) = μMoS2 − μMo − 2μS | (2) |
The value of μB in eqn (1), and therefore also Eform, largely depends on the experimental growth conditions. As limiting cases we consider the Mo-rich and S-rich situations. Under Mo-rich conditions the Mo chemical potential is equal to the bulk Mo value, μMo-richMo = μMo. In thermodynamic equilibrium one can assume that μMoS2 = μMo + 2μS, so that for Mo-rich conditions the S chemical potential is
. Hence, by using eqn (2), the chemical potentials for the Mo-rich limit can then be written as
| μMo-richMo = μMo | (3) |
![]() | (4) |
Analogously, for S-rich conditions one obtains
| μS-richS = μMo + Eform[MoS2] | (5) |
| μS-richS = μS | (6) |
Table 3 lists the formation energies of all the impurities studied in this paper. Impurity formation energies in both S-rich and Mo-rich conditions, respectively, are shown in Fig. 3. As is known, the smaller defect formation energy is, the more preferable to incorporate impurity atoms into the host material. As one halogen element NM occupies a sulfur site (NMS), it can release an additional valence electron to the conduction band, thus causing NMS to act as a donor. It can be observed from Table 3(b)–(e), the Eform gradually increase, and the defect formation energy of FS is smaller than others. This is mainly due to the stronger interaction of FS than ClS, BrS, and IS and the latter becomes smaller, indicating that FS is easier to form. The results are similar with the previous study of F doped in MoS2.47 In Fig. 3, the defect formation energy NbMo has the lowest formation energy, suggesting a favorable replacement of Mo atom with Nb, which is in good agreement with previous calculations.20 This can be interpreted by nearly the same radius of Mo and Nb atoms, indicating smaller distortion and fewer changes of the equilibrium structure and system energy. Interestingly, the defect formation energy of NbMoFS–1H-MoS2 are smallest, no matter in S-rich or Mo-rich, manifesting that simultaneous substitutional Nb and F atom codoping is stable and more favored than F and Nb atom monodoping in 1H-MoS2. Most importantly, the Nb and F is a charge-compensated donor–acceptor pair, which will be more favorable to enhance visible light photoactivity. We can see from Fig. 3 that all doped systems, except NbMo, in S-rich conditions have higher formation energy than in Mo-rich conditions. This reflects the fact that the formation energy is largely dependent on the growth conditions. The large variation of the formation energies for S substitutions in 1H-MoS2 with respect to different values of μS has also been reported in ref. 39. The negative formation energies indicate the corresponding substitutional reactions could happen. Thus, the NbMo and NbMoFS doped 1H-MoS2 are possible to be formed in the experiments.
| No. | Model | Eform (eV) | |
|---|---|---|---|
| Mo-rich | S-rich | ||
| (a) | NbMo | −0.25 | −3.094 |
| (b) | FS | −2.318 | −0.898 |
| (c) | ClS | −0.102 | 1.318 |
| (d) | BrS | 0.336 | 1.756 |
| (e) | IS | 0.912 | 2.332 |
| (f) | NbMoFS | −4.321 | −5.743 |
![]() | ||
| Fig. 4 Band structures of (a) pure, (b) NbMo, (c) NbMoFS, (d) FS, (e) ClS, (f) BrS, and (g) IS doped 1H-MoS2. The Fermi level indicated by dashed line is set at 0 eV. | ||
| No. | Model | Ev (eV) | Ec (eV) | Eg (eV) |
|---|---|---|---|---|
| (a) | Pure | 0 | 1.652 | 1.652 |
| (b) | NbMo | 0.097 | 1.457 | 1.360 |
| (c) | FS | −1.516 | −0.035 | 1.481 |
| (d) | ClS | −1.615 | −0.115 | 1.500 |
| (e) | BrS | −1.565 | −0.075 | 1.490 |
| (f) | IS | −1.470 | 0 | 1.470 |
| (g) | NbMoFS | 0 | 1.381 | 1.381 |
As is shown in Fig. 4(a), pure 1H-MoS2 is a kind of direct band gap semiconductor, with the computed band gap of 1.652 eV, which agrees well with the MoS2 photoluminescence experiment17 and previous theoretical studies.18,59 The bottom CB of pure 1H-MoS2 (Fig. 5(a1)) is mainly constituted of Mo 4d states, while the upper VB is composed of Mo 4d and S 3d states. This is one of the most important factors for the lower photocatalytic properties of pure 1H-MoS2. For the NbMo–1H-MoS2 (Fig. 4(b) and 5(b)), the Nd 4d states appear in the upper VB. Thus the VB extends forward the CB and the band gap is decreased to 1.36 eV. Considering the Nb atom is one d-electron less than the Mo atom, the Nb atom will act as hole doping in the MoS2 to form a p-type doping. More recently, an experimental study51 shows Nb-substituted (concentration range 15–25%) MoS2 nanoparticles exhibit p-type character, which coincides with our theoretical calculation. Because of one electron removal, the effect on the Fermi energy below the valence band maximum (VBM) is notable which mainly originate from hybridized d orbitals of Nb and Mo (see Fig. 5(b) and (b1)). The newly created defect states are rather delocalized and the charge excess spreads out up to the third nearest neighbor Mo atoms (Fig. 6(a)). And also the valence band (VB) looks sufficiently dispersive in our DFT band structure, and therefore the mobility of holes is expected to be rather large. On the other hand, some Nd 4d states are mixed with Mo 4d states in the CB of the NbMo–1H-MoS2. This is similar with the case of Nb doped into TiO2.60
We also consider substitutional doping at the S site with elements taken from the halogen family, namely F, Cl, Br, and I (Fig. 4(d)–(g) and 5(d)–(g)). These systems are expected to act as a source of n-type doping for 1H-MoS2, since they have one additional p-electron with respect to S. As a representative, we discuss in detail the results of F doping, since the other halogens present a similar electronic structure. For the FS–1H-MoS2 (Fig. 4(d)), the upper VB still consists of Mo 4d and S 3d states, while the F 2p states hybrid with Mo 4d states forming an occupied defect level below the CBM about 0.01 eV. These defect states originate from the hybridization between the F 2p and the Mo 4d states (see Fig. 5(d) and (d1)). This is very similar to the case of FS–ZnS.61 As a result, the band gap of the FS–1H-MoS2 decreases to 1.481 eV from 1.652 eV. For other halogen dopants (Cl, Br, and I), similar results have been obtained (Fig. 4 and 5, and Table 4).
It is very interesting that, different with the impurity states in the NbMo–1H-MoS2 and FS–1H-MoS2, there are no defect states in the NbMoFS–1H-MoS2 (Fig. 4(c)). According to the results discussed above, this can be attributed to the charge compensated codoping because NbMo and FS are the donor–acceptor passivated dopants. Fig. 5(c) and(c1) show that the Nb 4d orbits would hybridize with S 3p states at the upper VB, while the hybridization of F 2p orbits with Nb 4d states is obvious at the bottom CB in the NbMoFS–1H-MoS2. This hybridization will significantly strengthen the delocalization of Mo 4d, as shown in Fig. 6(h). Furthermore, the resonance of the dopants states with host phase states leads to about 16.4% reduction of the energy gap (1.381 vs. 1.652 eV), indicating codoping could effectively narrow the band gap, which is similar with the results of N + F-codoped TiO2.62 The NbMoFS–1H-MoS2 has a band structure similar to that of pure 1H-MoS2. The results are very crucial for 1H-MoS2 as a promising photocatalyst or photosensitizers for visible light in the heterogeneous semiconductor systems because this ensures improved charge carrier mobility in the codoped system. It should be pointed out that it is very difficult to precisely control the dopant sites in the fabricated samples. Thus, we also consider another case: Nb and F atoms are separated each other. The calculated results show that the codopant sites have little influence on the calculated electronic structure, especially in the vicinity of Fermi level. More importantly, the calculated defect formation energy (−3.683 eV) indicates that Nb and F atoms are more inclined to form covalent bond when they are codoped into 1H-MoS2. Therefore, we focus on the case that Nb atom is bonded to F atom in the 1H-MoS2 in this work.
As is known, the photocatalytic activity is not only determined by the quantity of available photons, but also influenced by the quantum efficiency. During the photocatalytic reaction process, the recombination of photogenerated carriers easily occurs, thus decreasing the quantum efficiency. For the doped semiconductor photocatalysts, the impurity levels are either recombination or separation center of photogenerated carriers, depending on the property of impurity levels. The latter role of the impurity levels can be understood as follows. Under irradiation, the electron in the impurity level moves to the CBM by capturing a photon with enough energy, and a hole is left at the same time. Consequently, the electron at the VBM can leap into the impurity level by absorbing another appropriate photon. Thus, the hole in the impurity level is eliminated and a new hole forms in the VBM. As a result, the impurity level does not change, while an electron and a hole is generated in the CB and VB, respectively, indicating the separation of photogenerated carriers. But this process is largely determined by the mobility of the carriers, which can be indirectly characterized by the efficient masses of carriers. The efficient masses of carriers can be calculated using the equation, m*e/h = ±(h/(2π2))/(d2E/dk2)2, where m*e/h is the electron/hole efficient mass, h is the Planck constant, k is the wave vector and E is the energy of k. The positive and negative signs correspond to the electron and hole, respectively. Obviously, the effective mass depends on the curvature of the parabolic portions of the bands near the CBM and VBM. To obtain the m*, one has to calculate the coefficient of the second-order term in the quadratic fit of E(k). A large coefficient of the second-order term is corresponding to the low effective mass of the charge carrier, which implies a high mobility. Especially, a larger value of m*e/m*h, suggests a greater variation of the electron–hole mobility and thus a lower recombination rate of the electron–hole pairs.
Fig. 4 shows that the levels at the CBM and the VBM in pure 1H-MoS2 and NbMo–1H-MoS2 are relatively dispersive, indicating the smaller value of m*e/m*h. Interestingly, the level just at the CBM in the NbMoFS–1H-MoS2 is fairly flat, whereas that below the Fermi level is much dispersive, as is shown in Fig. 4(c). Thus, the value of m*e/m*h in the NbMoFS–1H-MoS2 is larger than that in pure 1H-MoS2. The larger value of m*e/m*h also appears in NMS–1H-MoS2 (Fig. 4(d)–(g)). The quantitative results from Fig. 4 indicate that the recombination rate of the electron–hole pairs in both NbMoFS–1H-MoS2 and NMS–1H-MoS2 is low. On the other hand, Fig. 6(c)–(f) show that the level just near the Fermi level in the NMS–1H-MoS2 are localized around the substitutional NM atom and three nearest Mo atoms. On the contrary, the lowest level of CB of the NbMo–1H-MoS2 is evenly distributed among the nearest six Mo atoms (Fig. 6(a)), while the top level of VB on the nearest six Mo atoms and Nd atom (Fig. 6(b)). The distribution of these states are one of the most basic and the most important factors for the enhanced photocatalytic activity of NbMo–1H-MoS2, as found in the experiment.51
For the NbMoFS–1H-MoS2, the top level of VB is rather delocalized around the Nd atom (Fig. 6(h)), while the level at the bottom CB, mainly consisted of Mo 4d states mixing with small F 2p orbitals, in the NbMoFS–1H-MoS2 are distributed on the two Mo atoms connected with the dopant F, which can be clearly seen from Fig. 6(g). This further indicates that the value of m*e/m*h in the NbMoFS–1H-MoS2 is large, which is beneficial for the separation of the photogenerated charge carriers. Most importantly, the electrons in the VBM leap into the two Mo atoms by capturing a photon with enough energy (>1.38 eV). Therefore, the two Mo atoms connected with the dopant F become the active sites during the photocatalytic reaction process. This is, therefore, the first to reveal that the photocatalytic activity of the 1H-MoS2 can be enhanced by the charge compensated codoping, such as NbMo and FS, because the catalytic activity of the codoped 1H-MoS2 is not only derived from the under coordinated sulfur edge sites, but also from some new added active Mo sites in the basal planes which remain catalytically inert in the pure 1H-MoS2.
| This journal is © The Royal Society of Chemistry 2015 |