Based on earlier results on the photocatalytic properties of MoS2, the 1T form of MoSe2, prepared by lithium intercalation and exfoliation of bulk MoSe2, has been employed for the visible-light induced generation of hydrogen. 1T-MoSe2 is found to be superior to both 2H and 1T MoS2 as well as 2H-MoSe2 in producing hydrogen from water, the yield being in the 60–75 mmol h−1 g−1 range with a turn over frequency of 15–19 h−1. First principles calculations reveal that 1T-MoSe2 has a lower work function than 2H-MoSe2 as well as 1T and 2H-MoS2, making it easier to transfer an electron from 1T-MoSe2 for the production of H2.
Artificial photosynthesis has been recognized as a potential means of water splitting. Various strategies which can be employed for this purpose like dye sensitization of semiconductors or use of light-harvesting semiconducting nanostructures. TiO2 was the first material to be used as the photocatalyst for water-splitting.1 Since then several inorganic catalysts have been used for photocatalytic, photoelectrochemical, and electrocatalytic production of H2 from water. Semiconducting oxide nanoparticles are one of the most common photocatalysts used for this purpose2,3 and are preferred because they are chemically robust and stable against photocorrosion during water splitting. The intrinsic limitation of oxides is that they generally have a highly positive valence band (O 2p), making it difficult to find a material which has both sufficiently negative conduction band to reduce H2O to H2 along with a sufficiently small bandgap to absorb visible light.4,5 Metal sulfides and selenides, on the other hand, have less positive valence bands making them visible light active. Majority of the metal sulfides however undergo photocorrosion during the hydrogen evolution reaction (HER).
Transition metal dichalcogenides of lamellar structure have gained attention recently because of their interesting electronic properties and easy availability.6,7 Exfoliation of these materials into single or few-layers often brings about drastic changes in the electronic structure as compared to the bulk species. Dichalcogenides of MoS2 and WS2 generally occur in the 2H form with the trigonal prismatic arrangement of hexagonal S–M–S (M = Mo/W) triple layers are among the most studied of the layered metal chalcogenides. While the 2H forms of these metal dichalcogenides are semiconducting, the 1T forms are metallic.8–10 MoS2 has been widely used as a catalyst for electrochemical, photoelectrochemical, and photocatalytic H2 generation from water11–14 in consequence of having the conduction band minimum well above the H2O reduction potential.15–17 Nanoparticles of 2H-MoS2 as well as composites of 2H-MoS2 with graphene and other materials have been employed as catalysts yielding 0.05–10 mmol h−1 g−1 of H2 with a turn over frequency anywhere between 0.2 and 6 h−1.13,18–21 It has been shown recently that the composite of MoS2 with heavily nitrogen-doped graphene has an activity of 10.8 mmol h−1 g−1 and a turn over frequency of 2.9 h−1 under a 100 W halogen lamp.21 This was further improved by using 1T-MoS2 prepared by the exfoliation of bulk MoS2 by Li-intercalation. The metallic nature of 1T MoS2 is expected to be responsible for H2 evolution.21–23 2H-MoSe2 with an indirect bandgap of 1.05 eV has its conduction band minimum 0.37 eV higher than 2H-MoS2, and well above the water reduction potential,15,16 thereby making it an ideal catalyst for H2 evolution. The 1T form of MoSe2 is also metallic and could be expected to be a better catalyst than its 2H analogue for water reduction. Based on these findings, we have exfoliated bulk 2H-MoSe2 after Li-intercalation to obtain few-layer 1T-MoSe2 with the octahedral co-ordination of Mo22 and employed it for HER by under visible light. We find that 1T-MoSe2 is superior to 2H-MoSe2 as well as 1T-MoS2.
Bulk MoSe2 in the 2H form was intercalated with lithium using n-butyl lithium and exfoliation carried out by reacting the intercalated product with water.24–26 The high resolution transmission electron microscope image in Fig. 1(a) shows that the exfoliated sample corresponds to the 1T phase with an octahedral (Oh) or trigonal antiprismatic symmetry. The 1T form has the √3a × a arrangement which is related to its electronic structure.25,27 The shifting of the atoms from their equilibrium positions, probably arises because of the Jahn-Teller instability, resulting in chain clusterization of the metal atoms with the formation of a superlattice.28 This distorted phase is stable as a dispersion in water even after Li is removed, but restacks when dried,21 to transform into thermodynamically stable 2H phase. The Mo atoms in the 2H form of MoSe2 have trigonal prismatic coordination as is evident from the high resolution TEM image in Fig. 1(b). The packing of atoms in 2H MoSe2 is AbA type while in the 1T form it is AbC type. The point group of the trigonal prismatic 2H-MoSe2 is D3h while the 1T polytype belongs to the D3d point group.24 The 1T phases of both MoSe2 and MoS2 exhibit a Raman spectra which are distinctly different from those of the 2H-phases. In Fig. 2, we show the Raman spectra of the 1T phases of MoSe2 and MoS2 and compare them with the spectra of the 2H phases. We list the band positions of these phases in Table I.
Raman modes . | 2H MoSe2 (cm−1) . | 2H MoS2 (cm−1) . | 1T-MoSe2 (cm−1) . | 1T-MoS2 (cm−1) . | MoSe2 bulk (cm−1) . |
---|---|---|---|---|---|
J1 | … | … | 106.4 | 165.4 | … |
J2 | … | … | 150.7 | 236.6 | … |
J3 | … | … | 221.4 | 339.3 | … |
A1g | 236.2 | 400 | … | 414.3a | 240.1 |
E1g | 165 | … | 292.4 | 166.7 | |
E12g | 375.9 | 289.4 | 391.3 | … |
Raman modes . | 2H MoSe2 (cm−1) . | 2H MoS2 (cm−1) . | 1T-MoSe2 (cm−1) . | 1T-MoS2 (cm−1) . | MoSe2 bulk (cm−1) . |
---|---|---|---|---|---|
J1 | … | … | 106.4 | 165.4 | … |
J2 | … | … | 150.7 | 236.6 | … |
J3 | … | … | 221.4 | 339.3 | … |
A1g | 236.2 | 400 | … | 414.3a | 240.1 |
E1g | 165 | … | 292.4 | 166.7 | |
E12g | 375.9 | 289.4 | 391.3 | … |
Some 2H contribution.
Bulk (2H) MoSe2 has d2 electronic configuration and hexagonal (D3h) symmetry which would induce splitting of the 4d orbitals into three orbitals of closely spaced energies (shown in Fig. 3(a) (i)). The Mo 4dz2 level is occupied and spin paired forming the valence band minimum (VBM), while the other four orbitals form the empty conduction band. During lithium intercalation, a structural re-orientation occurs to the stable half-filled d-orbital configuration in the octahedral geometry (D3d) as shown in Fig. 3(a) (ii) which is responsible for the Mo atoms to go from the prismatic co-ordination to the anti-prismatic co-ordination in the 1T-form. Addition of water causes an exothermic reaction because of which the MoSe2 sheets get separated and remain in this metastable 1T-form. The crystal-field splitting of Mo 4d under the octahedral Oh field generates two set of degenerate orbitals as shown in Fig. 3(a) (iii). The incompletely filled dyx, dzx, dzy orbital gives rise to the metallic properties of 1T-MoSe2. The Fermi level in 1T-MoSe2 therefore lies in the Mo 4d making it metallic. Based on the electronic configuration of 1T and 2H phases of MoSe2 it is clear that when an extra electron is added to 2H-MoSe2, it resides in the degenerate dyx, dx2-y2 states and destabilizes the lattice, while in case of 1T-MoSe2 the extra electron induces half-filled configuration of dyx, dzx, dzy and increases the stability of the 1T phase.
The hydrogen evolution activity of 1T-MoSe2 was studied using Eosin Y as the sensitizer and triethanolamine as the sacrificial electron donor. The reaction of dye-sensitized H2 evolution over MoSe2 involves photosensitization of Eosin Y followed by formation of Eosin Y anion (EY−). EY− being highly reactive donates this electron to MoSe2, which then catalyzes the reduction of proton to H2 as shown in Fig. 3(b). Fig. 4 shows the time course of hydrogen evolution by 1T MoSe2. The yields are in range of the 60–75 mmol g−1 h−1 and remains the same at least up to 5 cycles, i.e., 30 h, with 0.014 mM of dye being added after each cycle (see Fig. S2 of the supplementary material).29 The turn over frequencies (TOF) are in the range 15–19 h−1. The catalytic activity of the 1T form of MoSe2 is nearly few hundred times higher than that of the 2H form (see inset of 4(a)). It is noteworthy that the yield of H2 and TOF with 1T-MoSe2 is superior even to those found with 1T MoS2. The 2H form of MoSe2 too shows better activity than that of 2H-MoS2 (yield of 0.05 mmol g−1 h−1 and TOF of 0.008 h−1). In Table II, we have compared the hydrogen evolution activity and TOF of different transition metal chalcogenides. 1T-MoSe2 shows higher activity compared to these transition metal chalcogenides and their composites. 1T forms of other metallic transition metal sulfides like TaS2 and TiS2 have earlier been shown to be active co-catalysts for H2 evolution.30 However, the activity for H2 evolution is much lower in these chalcogenides as compared to 1T-MoS221 and 1T-MoSe2. Greater stability afforded by the extra electron to Mo 4d level by inducing a half-filled configuration of dyx, dzx, dzy, as compared to Ta 5d (with 1 electron) and Ti 3d (with no electron) is probably the reason for higher activity of 1T-MoS2 and 1T-MoSe2.
Compounds . | Light source . | Activity (mmol h−1 g−1) . | TOF (h−1) . | Reference . |
---|---|---|---|---|
MoS2/CdS | 300 W Xe lamp | 5.3 | ∼0.7 | 20 |
TaS2/CdS | 400 W Xe lamp (λ > 399 nm) | 2.32 | 0.57 | 30 |
2H MoS2a | 100 W halogen lamp | 0.05 | 0.008 | 21 |
NRGO-MoS2a | 100 W halogen lamp | 10.8 | 2.9 | 21 |
1T MoS2a | 100 W halogen lamp | 30 | 6.5 | 21 |
1T-MoSe2a | 100 W halogen lamp | 62 ± 5 | 15.5 ± 2 | Present work |
Few layer 2H MoSe2a | 100 W halogen lamp | 0.08 | 0.02 | Present work |
Compounds . | Light source . | Activity (mmol h−1 g−1) . | TOF (h−1) . | Reference . |
---|---|---|---|---|
MoS2/CdS | 300 W Xe lamp | 5.3 | ∼0.7 | 20 |
TaS2/CdS | 400 W Xe lamp (λ > 399 nm) | 2.32 | 0.57 | 30 |
2H MoS2a | 100 W halogen lamp | 0.05 | 0.008 | 21 |
NRGO-MoS2a | 100 W halogen lamp | 10.8 | 2.9 | 21 |
1T MoS2a | 100 W halogen lamp | 30 | 6.5 | 21 |
1T-MoSe2a | 100 W halogen lamp | 62 ± 5 | 15.5 ± 2 | Present work |
Few layer 2H MoSe2a | 100 W halogen lamp | 0.08 | 0.02 | Present work |
Eosin Y dye sensitized.
It is worthwhile to note that MoSe2 shows H2 evolution on sensitization with dye. It may therefore not be considered to be a photocatalyst similar to TiO2, where photo-excited electrons directly reduce H2O. MoSe2 is still different from co-catalysts like Pt which show H2 evolution in the presence of another catalyst like TiO2 and cannot evolve H2 from water even on sensitization with a dye. Conduction Band Minimum of MoSe2 is favourable for HER but cannot generate enough photoelectrons by absorption of visible light (the bandgap being very small). The dye donates an electron to MoSe2 and induces the H2 evolution reaction. We must point out that the exfoliated sample on restacking loses activity due to partial conversion to 2H-form. On annealing it transforms completely to the 2H form, thereby losing its hydrogen evolution activity (see Fig S3 of the supplementary material).29
In order to understand the higher activity of 1T-MoSe2 in comparison to 2H-MoSe2 and the 1T and 2H forms of MoS2, we have carried out first-principles calculations based on density functional theory as implemented in Quantum ESPRESSO package,31 in which the ionic and core-valence electron interactions are modeled with ultrasoft pseudopotentials.32 The exchange-correlation energy of electrons is treated within a Generalized Gradient Approximation (GGA) functional as parametrized by Perdew, Burke, and Ernzerhof.33 We use an energy cutoff of 35 Ry to truncate the plane wave basis used in representing the Kohn-Sham wave functions, and an energy cutoff of 280 Ry to represent the charge density. Structures are relaxed till the Hellman-Feynman forces on each atom are less than 0.02 eV/Å. We have used a periodic supercell geometry to simulate a 2D sheet, including vacuum of 15 Å to separate the adjacent periodic images of the sheet. For self-consistent Kohn-Sham (scf) calculations, configurations of √3 × √3 and √3 × 1 supercells, the BZ integrations are sampled over uniform meshes of 7 × 7 × 1 and 12 × 7 × 1 k-points, respectively. Since KS-DFT typically underestimates electronic bandgaps (a known limitation), we employ hybrid functional based on Hartree-Fock-Exchange (HSE)34 with screened Coulomb potential to estimate the bandgaps more accurately. The calculations were based on first-principles DFT using Projector Augmented Wave (PAW) method35,36 as implemented in the VASP (Vienna Ab-initio Simulations Package).37 We have studied two superstructures of 1T-MoX2 (where X = S and Se), √3 × √3 and √3 × 126 (Fig. 5). Among the two superstructures, √3 × 1 is metallic and shows dimerization of Mo atoms, and √3 × √3 is semiconducting with trimerized Mo atoms. From phonon dispersion, we find that, both MoS2 and MoSe2 are stable in the √3 × √3 and √3 × 1 superstructures. However, MoS2 is energetically more stable in the √3 × √3 compared to √3 × 1 by 27 meV/f.u., while the √3 × 1 super-structure of MoSe2 is energetically more stable than the √3 × √3 super-structure by 33 meV/f.u.
Experimentally, MoSe2 is found to be in the √3 × 1 super-structure, in agreement with our first-principles results. Henceforth, we shall consider the √3 × √3 superstructure for MoS2 and √3 × 1 superstructure for MoSe2. To determine the efficiency of MoX2 in reducing a proton to hydrogen as observed in experiments, we have estimated their electron affinities (EA) and work function (φ). For metallic states, the relevant property here is the work function. The EA is estimated as the difference between the vacuum potential (Evac) and the lowest energy conduction band (ECB). Since DFT is a ground state theory, estimation of the bandgap and hence the ECB is not accurate. Hence, we replace the ECB with EVB + Eg, where EVB is the energy of the highest energy valance band and Eg is the bandgap. Since Eg is grossly underestimated in DFT calculations, we use the HSE corrections (using VASP) to determine Eg accurately. For the monolayered MoS2, experimental value of bandgap (1.8 eV6) is available.
Comparison of the experimental bandgap with calculated bandgap for 2H-MoS2 reveals that Kohn-Sham bandgap is underestimated by 7.2% and the HSE bandgap is overestimated by 17.7% (see Table III), in agreement with Ahuja et al.38 It is thus clear that the HSE method overestimates the experimental bandgap, whereas the KS-DFT calculation (GGA) yields a better estimate. We use estimates of Eg obtained from KS-DFT calculations in this work. The work function for metals and semiconductors is calculated as φ = Evac − EF (where EF = Fermi energy) and φ = Evac − EVB respectively. We find that (a) the 2H and 1T-polytypes of MoS2 have a greater φ than that of the respective structure of MoSe2 (refer to Table IV). This implies that it is easier to extract an electron from MoSe2 compared to that of MoS2 in both 1T and 2H polytypes. (b) The 1T polytype has a lower φ than that of 2H, which means that it is easier for the 1T to donate electron compared to the 2H-structure. This explains why the 1T-polytype of MoSe2 produces hydrogen more efficiently than the 2H-polytype as observed in experiments. The electron affinities of both 1T and 2H polytypes indicate that MoS2 has a stronger electron affinity (indicating a higher tendency to attract electrons) than that of MoSe2 (refer to Table IV), and the work function is also larger for MoS2. Thus, though MoS2 more readily attracts/accepts electrons, it does not donate it that easily. Hence, MoSe2 is efficient in hydrogen evolution as compared to that of MoS2 as observed in experiments here.
. | Bandgap (eV) . | ||
---|---|---|---|
Compounds . | KS-DFT (VASP) . | HSE (VASP) . | Expt. . |
2H-MoS2 | 1.67 | 2.12 | 1.834 |
2H-MoSe2 | 1.45 | 1.88 | |
1T-MoS2 (√3 × √3) | 0.76 | 1.28 | |
1T-MoSe2(√3 × √3) | 0.64 | 1.16 |
. | Bandgap (eV) . | ||
---|---|---|---|
Compounds . | KS-DFT (VASP) . | HSE (VASP) . | Expt. . |
2H-MoS2 | 1.67 | 2.12 | 1.834 |
2H-MoSe2 | 1.45 | 1.88 | |
1T-MoS2 (√3 × √3) | 0.76 | 1.28 | |
1T-MoSe2(√3 × √3) | 0.64 | 1.16 |
. | 1T-form . | 1T-form . | . | |||||
---|---|---|---|---|---|---|---|---|
Superstructure . | √3 × √3 . | √3 × 1 . | 2H-form . | |||||
Compound | MoS2 | MoSe2 | MoS2 | MoSe2 | MoS2 | MoSe2 | ||
EA (eV) | 4.95 | 4.42 | … | … | 4.22 | 3.78 | ||
WF (eV) | 5.68 | 5.20 | 5.63 | 5.00 | 5.86 | 5.35 |
. | 1T-form . | 1T-form . | . | |||||
---|---|---|---|---|---|---|---|---|
Superstructure . | √3 × √3 . | √3 × 1 . | 2H-form . | |||||
Compound | MoS2 | MoSe2 | MoS2 | MoSe2 | MoS2 | MoSe2 | ||
EA (eV) | 4.95 | 4.42 | … | … | 4.22 | 3.78 | ||
WF (eV) | 5.68 | 5.20 | 5.63 | 5.00 | 5.86 | 5.35 |
To connect the hydrogen (H) binding energy to the catalytic activity of the MoX2 compounds, we have determined the binding energy of hydrogen (H) to √3 × 1 1T superstructure of MoS2 and MoSe2. This superstructure is relevant to the experiments reported here, and has the lowest work function.
The hydrogen binding energy is calculated as
where n is the number of H atoms considered in the simulation.
Bulk MoS2 and MoSe2 do not absorb hydrogen (Eads > 0). It has been reported that edges of these dichalcogenides are catalytically active in hydrogen adsorption.22 We have therefore simulated ribbons of MoX2 with two different types of edges (Mo terminated edge and X terminated edge), and their interaction with H (with 100% H coverage at the edges). The hydrogen binding energies at Mo sites at the edges of MoS2 and MoSe2 are −33.8 meV/f.u. and −32.3 meV/f.u., respectively. The respective Mo–H bond lengths are 1.72 Å and 1.73 Å. The hydrogen binding energies at the S/Se edges of MoS2 and MoSe2 are −34.6 meV/f.u. and −13.1 meV/f.u. respectively. The corresponding X–H bond lengths are 1.35 Å and 1.48 Å. The binding energy of hydrogen at the metal edge is about the same in the two compounds but the Se edge shows weaker binding with hydrogen than the S edge. According to the volcano plot,39,40 this suggests a higher exchange current for hydrogen evolution over MoSe2 compared to MoS2. This is consistent with our analysis based on the work functions. Since MoSe2 has a lower work function than MoS2, its Fermi energy (EF) lies closer to the normal hydrogen electrode (ENHE), which allows an easy exchange of an electron with MoSe2 (H atom has a weaker binding at the Se edge) as compared to MoS2. Thus, MoSe2 is more efficient in facilitating the hydrogen evolution reaction.
In conclusion, metallic 1T-MoSe2 prepared by Li intercalation followed by exfoliation of bulk 2H-MoSe2 shows excellent H2 evolution activity in comparison to few-layered semiconducting 2H-MoSe2. Interestingly, 1T-MoSe2 shows better H2 evolution activity than 1T-MoS2 as well. Our first-principles analysis reveals that MoSe2 has a lower work function as compared to MoS2, and that the 1T-structure exhibits lower work function than the 2H-structure for both MoX2 (X = S, Se). This results in easy transfer of electron from the MoSe2 for the reduction to hydrogen, and hence MoSe2 is more efficient for hydrogen evolution reaction compared to MoS2, and in agreement with the experimental results.