Skip to main content

ORIGINAL RESEARCH article

Front. Chem., 25 August 2022
Sec. Theoretical and Computational Chemistry
Volume 10 - 2022 | https://doi.org/10.3389/fchem.2022.1003027

A novel two-dimensional transition metal dichalcogenide as water splitting photocatalyst with excellent performances

www.frontiersin.orgFang Wang1,2 www.frontiersin.orgZishuang Cheng1,3* www.frontiersin.orgXiaoming Zhang3 www.frontiersin.orgChunxiao Xie1,4 www.frontiersin.orgFucai Liu2 www.frontiersin.orgChuntao Chang1* www.frontiersin.orgGuodong Liu3
  • 1School of Mechanical Engineering, Neutron Scattering Technical Engineering Research Center, Dongguan University of Technology, Dongguan, China
  • 2School of Optoelectronic Science and Engineering, University of Electronic Science and Technology of China, Chengdu, China
  • 3School of Materials Science and Engineering, Hebei University of Technology, Tianjin, China
  • 4Guangdong-Taiwan College of Industrial Science & Technology, Dongguan University of Technology, Dongguan, China

With the rising demand for renewable energy, photocatalysts are considered the most promising solution to harness solar energy, and the search for photocatalysts with excellent performances remains an urgent task. Here, based on density functional theory (DFT), the photocatalytic properties of MoWS4 are systematically investigated. The MoWS4 monolayer and bilayer are demonstrated as semiconductors with indirect band gaps of 2.01 and 1.48 eV. Moreover, they exhibit high and anisotropic light absorption coefficients of up to ∼105 cm−1 in the visible-ultraviolet region. The intrinsic band edge positions could fully satisfy the redox potentials of water without any external adjustment. The electron mobility of MoWS4 monolayer is 557 cm2 V−1s−1, which is seven times higher than MoS2 monolayer. Hence, MoWS4 can be regarded as a promising 2D photocatalyst candidate for water splitting.

Introduction

With the depletion of fossil energy and the increasing pollution of the natural environment, the demands for renewable energy become critical and urgent for sustainable development of global economy. In 1972, Fujishima and Honda discovered that TiO2 can split water to produce hydrogen and oxygen in the presence of sunlight, making photocatalysis one of the most noteworthy solutions to harness solar energy (Fujishima and Honda, 1972). Afterwards, great efforts have been made to develop effective photocatalysts, including transition metal oxides, sulfides, nitrides, and so forth (Tsuji et al., 2005; Suntivich et al., 2011; Han et al., 2018). However, the low quantum efficiency derived from charge recombination on the surface and in the bulk of these photocatalysts could not meet the criteria of favorable photocatalyst for sunlight driven water splitting (Fujishima et al., 2000).

Due to the interesting structures and corresponding electronic properties, two-dimensional (2D) materials have been widely used in various fields and also provide new research directions for efficient photocatalysis (Singh et al., 2015). In recent years, 2D photocatalysts showed the greater advantages over their bulk phase counterparts in terms of photocatalytic performance. For example, Both SnS2 monolayer and ZnSe nanosheet exhibited higher photocurrent density than their bulk materials (Sun et al., 2012a; Sun et al., 2012b). In addition, various 2D materials have also been theoretically and experimentally demonstrated to be used as photocatalysts for water splitting, such as 2D transition metal dichalcogenides, g-C3N4, phosphorene, and so on (Wang et al., 2009; Zhuang and Hennig, 2013a; Rahman et al., 2016; Phuc et al., 2018; Zhang et al., 2022). However, few of these photocatalysts can simultaneously satisfy high visible light absorption, high carrier mobility and perfect band edge positions. For instance, MoS2 monolayer showed lower carrier mobility, and MoTe2 monolayer can’t perfectly meet the redox potential of water (Wang et al., 2009; Zhuang and Hennig, 2013a; Cai et al., 2014; Rahman et al., 2016; Zhang et al., 2022). Both GaS and GaSe monolayers demonstrated low visible light absorption due to the large band gaps (Zhuang and Hennig, 2013b). Therefore, it is still a challenge to develop water splitting photocatalysts with excellent performances.

In this work, based on the first-principles calculations, we propose a novel 2D transition metal dichalcogenide namely MoWS4 and systemically investigate its photocatalytic properties. Firstly, the stability of MoWS4 monolayer is confirmed by calculating its phonon spectra and ab initio molecular dynamics (AIMD) simulations. Secondly, we calculate the band structures of MoWS4 monolayer and bilayer, and show their semiconductive characteristic. Then, relevant photocatalytic properties are systematically investigated. It is found that MoWS4 monolayer and bilayer can nicely meet the redox potentials without strain engineering, and their light absorption coefficients reach ∼105 cm−1 in the visible-ultraviolet region. Moreover, the electronic mobility of MoWS4 monolayer is as high as 557 cm2 V−1s−1.

Computational details

For geometric and electronic structures, all calculations are performed by using the Vienna ab into simulation package (VASP) based on density functional theory (DFT) (Kresse and Furthmuller, 1996). We choose the generalized gradient approximation (GGA) of the Perdew–Burke–Ernzerhof (PBE) functional as the exchange-correlation functional to perform these calculations (Perdew et al., 1996). For the 2D monolayer structure, the vacuum layer thickness is set as about 20 Å to avoid layer-to-layer effects. The kinetic energy cutoff is set as 500 eV. The Brillouin zone is regulated with 10 × 6 × 1. During the calculations, the DFT-D2 method with Grimme correction is used to describe the long-range van der Waals interactions (Grimme, 2006). Besides, all atoms are fully relaxed, and the energy and force convergence criteria are set as 10–6 eV and 0.01 eV Å−1, respectively. Except for PBE functional, to obtain the more accurate results, the HSE06 functional is also adopted to calculate the band structures and the band edge positions (Deák et al., 2010). To identify structural stability of MoWS4 monolayer, its phonon spectra are calculated by using the PHONOPY code (Gonze and Lee, 1997). A 6 × 3 supercell structure of MoWS4 monolayer is used in ab initio molecular dynamics (AIMD) simulations (NVT ensemble), which is carried out for 3 ps with a time step of 1 fs at 500 K (Cimas et al., 2014).

The optical properties can be defined by the complex dielectric function (frequency) for characterization:

ε(ω)=ε1(ω)+iε2(ω)(1)

where ε1(ω) and ε2(ω) represent the real and imaginary parts, respectively. Based on the Kramers–Kronig transformation (Kuzmenko, 2005), the real part [ε1(ω)] can be expressed as follows:

ε1(ω)=1+(2π)p0dω(ω)2ε2(ω)(ω)2(ω)2(2)

where p is the integral principal value. In addition, the imaginary part [ε2(ω)] can be described as (Saha et al., 2000):

ε2(ω)=4πe2m2ω2i,f2d3k(2π)3|ik|P|fk|2Fik(1Ffk)δ(EfkEikE)(3)

where ω, E, F, P, ||ik, and |fk represent the incident photon frequency, the incident photon energy, the Fermi function, the transition matrix, the CB state and VB state, respectively. Finally, the absorption coefficient [α(ω)] can be calculated by (Cheng et al., 2022a):

α(ω)=2ωc[ε12(ω)+ε22(ω)ε1(ω)]1/2(4)

where c denotes the speed of light in vacuum.

The carrier mobility is calculated by the following formula (Qiao et al., 2014):

μ=eħ3C2DkBTmemdEd2(5)

where C2D is the elastic modulus, which can be expressed by C2D=1S02Eδ2, here E is the total energy of monolayer after deformation, S0 is the lattice area of monolayer under equilibrium and δ is the uniaxial strain. Besides, kB is the Boltzmann constant, T is the temperature. me is the effective mass along the transport direction and md is the average effective mass determined by md=mxmy. Ed represents the deformation potential constant, which is expressed by Ed=Eedgeδ, here Eedge represents the shift of the band edge position with respect to the uniaxial strain δ.

Results and discussions

Crystal structures, stability, and electronic properties

Figure 1A shows the top and side views of 4 × 2 supercell of MoWS4 monolayer, which has a space group of Pmm2. The light green area is the unit cell of MoWS4 monolayer, which contains four atoms. The pink, blue, and yellow balls represent Mo, W, and S atoms, respectively. MoWS4 monolayer is a honeycomb structure from the top view, while sandwich structure from the side view, which is very similar to the structure of MoS2 monolayer (Kan et al., 2014). After fully geometric optimization, the lattice constants are a = 3.188 Å and b = 5.526 Å with a layer thickness of 3.119 Å.

FIGURE 1
www.frontiersin.org

FIGURE 1. (A) Top and side views of MoWS4 monolayer. The pink, blue, and yellow balls represent Mo, W, and S atoms, respectively. The light green area is the unit cell of MoWS4 monolayer (B) Phonon spectra of MoWS4 monolayer (C) Evolution of the total energy and a snapshot of MoWS4 monolayer after a 3,000 fs AIMD simulations in vacuum at 500K

To confirm the stability of MoWS4 monolayer, the formation energy (Eform), defined as: Eform=[EMoWS4EMoEW4ES]/6, is calculated, where EMoWS4 presents the total energy of the unit cell of MoWS4 monolayer, and EMo, EW, and ES present the energies of each Mo, W, and S atom in itself bulk phase, respectively. The calculated formation energy of MoWS4 monolayer is −1.1 eV/atom (<0), implying the procedure of synthesizing MoWS4 monolayer is exothermic and favorable. Besides, the dynamic and thermal stabilities of MoWS4 monolayer are further investigated. As shown in Figure 1B, the calculated phonon spectra exhibit no imaginary frequency, indicating high dynamic stability of MoWS4 monolayer. The thermal stability is further studied via AIMD calculations. As illustrated in Figure 1C, the total energy of the MoWS4 monolayer remains essentially stable and no deformation occurs in its final structure. These results suggest that MoWS4 monolayer exhibits good dynamic and thermal stability, which deduces the possibility of experimental synthesis of MoWS4 monolayer.

Figure 2 shows the top and side views of MoWS4 bilayer. Three stacking patterns (AA, AB, and AC) are considered in the bilayer structures of MoWS4. The results of the calculated total energy are −94.2727 eV, −94.4175 eV, and −94.2726 eV for AA, AB, and AC stacking patterns, respectively. AA and AC stacking structures have the same layer spacing of 6.76 Å, and AB stacking has a spacing of 6.13 Å. Obviously, the AB stacking structure is most stable energetically.

FIGURE 2
www.frontiersin.org

FIGURE 2. Top and side views of MoWS4 bilayer with three stacking patterns (AA, AB, and AC).

Figures 3A,B shows the electronic band structures of MoWS4 monolayer and bilayer obtained by using the PBE functional (blue) and the HSE06 functional (red). The Fermi level and the high symmetry path are set as 0 eV and Γ-X-M-Y-Γ, respectively. The MoWS4 monolayer and bilayer are indirect band gap semiconductors and their band gaps calculated by HSE06/PBE functional are 2.01/1.67 eV and 1.48/1.14 eV, respectively. The band gaps obtained by the HSE06 functional are larger than that obtained by the PBE functional since the PBE functional tends to underestimate the band gap. Figures 3C,D shows the maps of charge density for MoW4 monolayer at the conduction band minimum (CBM) and the valence band maximum (VBM), and it can be concluded that the CBM is mainly contributed by the Mo atoms, whereas the VBM comes from a combination of Mo, W, and S atoms.

FIGURE 3
www.frontiersin.org

FIGURE 3. Band structures of (A) MoWS4 monolayer and (B) MoWS4 bilayer. The red and blue represent the HSE06 functional and PBE functional, respectively. Top and side views of the charge density at the (C) CBM, and (D) VBM. The isosurface value is set as 0.012 e Å−3.

Photocatalytic water splitting and optical properties

Considering the excellent semiconductor properties of MoWS4 monolayer and bilayer, we systematically study their feasibility as photocatalysts for water splitting. It is well known that a photocatalytic candidate should meet the following conditions: Firstly, its band gap should exceed the free energy of water splitting (1.23 eV). Obviously, the band gaps of MoWS4 monolayer and bilayer both exceed 1.23 eV; Secondly, its band edges must cross the redox potentials of water. The CBM energy should be higher than the reduction potential of H+/H2 (−4.44 eV), and the VBM energy should be lower than the oxidation potential of O2/H2O (−5.67 eV) (Abe, 2010; Sun et al., 2019; Cheng et al., 2022b). For 2D materials, the band edges with respect to the vacuum level (ECBM/VBMVac) can be obtained by: ECBM/VBMVac=ECBM/VBMDFTVvacuum, where ECBM/VBMDFT represents the value of CBM/VBM obtain by DFT and Vvacuum represents the electrostatic potential in the vacuum region. We plot the map of band edges relative to the vacuum level of MoWS4 monolayer and bilayer in Figure 4A, and find that the band edge positions of MoWS4 monolayer and bilayer can perfectly satisfy the redox potential for the water splitting reaction at pH = 0. Specifically, the band edges of CBM/VBM for MoWS4 monolayer and bilayer are −4.202/−6.213 eV and −4.230/−5.712 eV, respectively. Thus, the results suggest that the MoWS4 monolayer and bilayer can be promising candidates for water splitting photocatalysts.

FIGURE 4
www.frontiersin.org

FIGURE 4. (A) The locations of VBM and CBM of MoWS4 monolayer and bilayer, which are obtained by the HSE06 functional. The redox potentials of water splitting at pH = 0 are shown for comparison. All the energy levels are based on the vacuum level, and the vacuum level is set as 0 eV. The optical absorption coefficients for the different directions of (B) MoWS4 monolayer and (C) MoWS4 bilayer.

Besides, efficient light absorption is also an important feature for water splitting photocatalysts. Figures 4B,C shows the light absorption spectra of the MoWS4 monolayer and bilayer within the visible-ultraviolet light range. The result can be summarized as follows: 1) Their light absorption coefficients in the ultraviolet light range are higher than visible light region (up to ∼106 cm−1); 2) Their light absorption coefficients of MoWS4 bilayer overall are higher than that of MoWS4 monolayer; 3) Their light absorption coefficients are significantly anisotropic: that is higher in the z direction, but have a wider range of light absorption spectra in the x and y directions. It is known that the high light absorption coefficients can guarantee the effective use of solar energy, which is very favorable for water splitting photocatalysts (Zhao et al., 2018; Fan et al., 2021). Therefore, both MoWS4 monolayer and bilayer can be promising potential candidates as photocatalysts.

Strain engineering and high carrier mobility

We further investigate the effect of strain engineering on the photocatalytic performance of MoWS4 monolayer. Figure 5 shows the band structures of MoWS4 monolayer under the in-plane biaxial strain from −2 to 2%. Its band gap increases when compressive strain increases or tensile strain decreases, and its band gap varies in the range of 1.508–2.378 eV. In addition, when it is subjected to compressive strain, the indirect bandgap changes to be a direct bandgap; while when it is subjected to tensile strain, it remains an indirect bandgap semiconductor. As illustrated in Figure 6, we study the changes of its band edge positions under the applied in-plane biaxial strain. All of VBM positions are lower than the oxidation potential of O2/H2O under the −2∼2% in-plane biaxial strain, which indicates that MoWS4 monolayer always serves as a potential photocatalyst to generate oxygen. Besides, when the applied tensile strain reaches 2%, its CBM positions become lower than the reduction potential of H+/H2. Thus, excessive tensile strain will lead to its inability to produce hydrogen. In general, compressive strain does not cause MoWS4 monolayer to deviate from the basic requirements for water splitting photocatalyst, but tensile strain can easily affect its hydrogen production performance.

FIGURE 5
www.frontiersin.org

FIGURE 5. Band structures of MoWS4 monolayer under the biaxial strain from −2 to 2% are calculated by HSE06 functional.

FIGURE 6
www.frontiersin.org

FIGURE 6. Strain effects on band edge positions of MoWS4 monolayer with respect to the vacuum level (0 eV). The redox potentials of water splitting at pH = 0 are shown for comparison.

As we know, the fast carrier migration capability is necessary for high performance photocatalysts. Thus, the PBE functional is used to calculate the carrier mobility of MoWS4 monolayer according to Eq. 5. Subsequently, we calculate its carrier effective masses (m), in-plane stiffness (C2D) and deformation potential constants (Ed). Therein, to obtain Ed of MoWS4 monolayer, the linear fitting maps of band edge positions are plotted as the function of the applied uniaxial strain δ along the x and y directions (Bardeen and Shockley, 1950; Zhang et al., 2021a), as illustrated in Figure 7 and the result is summarized in Table 1. It is notable that the values of Ed have a small difference along the x and y directions, which shows that the scattering ability of its carriers in different directions is also similar. In addition, the carrier effective masses of MoWS4 monolayer have a small difference along the x and y directions. The electron effective masses of MoWS4 monolayer are much lower than its hole effective masses (me = 0.377m0 and 0.374m0 along the x and y directions), which are smaller than that of many 2D photocatalytic materials, such as Penta-PdSSe (me = 2.16m0), Penta-PdSe2 (me = 1.88m0), δ-SnS (me = 1.01m0), SnP3 (me = 0.9m0) (Long et al., 2018; Sun et al., 2018; Zhang et al., 2021b; Xiao et al., 2021). The smaller carrier effective masses benefit the transfer rate of photogenerated carriers in the photocatalytic process. What’s more, the calculated electron mobility along x and y directions are 529 cm2V−1s−1 and 557 cm2V−1s−1, respectively, whereas the calculated hole mobility along x and y directions are both 38 cm2 V−1s−1. The reason for this difference is that the hole effective mass is much larger than the electron effective mass. More importantly, as shown in Figure 8, the electron mobility of MoWS4 monolayer can significantly exceed that of many other 2D photocatalytic materials such as MoS2 (µe = 72.16 cm2 V−1 s−1, µh = 200.52 cm2 V−1 s−1), Penta-PdS2 (µe = 40.97 cm2 V−1 s−1, µh = 339.25 cm2 V−1 s−1), Penta-PdSe2 (µe = 29.40 cm2 V−1 s−1, µh = 534.55 cm2 V−1 s−1), SnP2S6 monolayer (µe = 148.48 cm2 V−1 s−1, µh = 143.12 cm2 V−1 s−1), As2S3 monolayer (µe = 253.11 cm2 V−1 s−1, µh = 10.85 cm2 V−1 s−1) (Cai et al., 2014; Wang et al., 2015; Long et al., 2018; Jing et al., 2019; Liu et al., 2021). It is well known that materials with high carrier mobility can effectively reduce their photogenerated electron and hole recombination rates and increase the participation rate in redox reactions, which are beneficial for photocatalytic processes. Thus, our results suggest that MoWS4 monolayer is a promising 2D photocatalyst candidate for water splitting.

FIGURE 7
www.frontiersin.org

FIGURE 7. Linear fitting maps of VBM and CBM locations of MoWS4 monolayer under the strain along x and y directions.

TABLE 1
www.frontiersin.org

TABLE 1. Carrier effective masses m, In-plane stiffness C2D, Deformation potential constants Ed, and carrier mobility μ for MoWS4 monolayer along the x and y directions.

FIGURE 8
www.frontiersin.org

FIGURE 8. Carrier mobility of MoWS4 compared with other typical 2D photocatalysts.

Conclusion

Based on the first-principles calculations, we systematically investigate photocatalytic properties of two-dimensional MoWS4. It is found that both the MoWS4 monolayer and bilayer are semiconductors with indirect band gaps and show high and anisotropic light absorption coefficients in the visible-ultraviolet range. The band edge positions of the materials can satisfy the redox potentials perfectly and the electron mobility of MoWS4 monolayer is up to 557 cm2 V−1 s−1, which outperforms many other 2D photocatalytic materials, such as MoS2 monolayer, Penta-PdS2, Penta-PdSe2, and SnP2S6 monolayer. These results indicate that MoWS4 can be a promising photocatalyst for water splitting with outstanding performances.

Data availability statement

The original contributions presented in the study are included in the article/Supplementary Material, further inquiries can be directed to the corresponding authors.

Author contributions

FW: Methodology, Investigation, Writing-original draft. ZC: Conceptualization, Investigation, Methodology, Writing—original draft. XZ: Investigation, Writing-original draft. CX: Validation, Resources. FL: Data curation, Validation. CC: Conceptualization, Writing—original draft, Supervision, Funding acquisition. GL: Validation, Data curation, Supervision.

Funding

The authors are grateful for support from the National Natural Science Foundation of China (52101192 and 51871056), Foundation for Basic and Applied Basic Research-Joint Foundation of Dongguan-Guangdong Province (2020A1515111109), and Department of Education of Guangdong in China (2018KZDXM069 and 2021 KCXTD050).

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

Abe, R. (2010). Recent progress on photocatalytic and photoelectrochemical water splitting under visible light irradiation. J. Photochem. Photobiol. C Photochem. Rev. 11, 179–209. doi:10.1016/j.jphotochemrev.2011.02.003

CrossRef Full Text | Google Scholar

Bardeen, J., and Shockley, W. (1950). Deformation potentials and mobilities in non-polar crystals. Phys. Rev. 80, 72–80. doi:10.1103/physrev.80.72

CrossRef Full Text | Google Scholar

Cai, Y. Q., Zhang, G., and Zhang, Y. W. (2014). Polarity-reversed robust carrier mobility in monolayer MoS2 nanoribbons. J. Am. Chem. Soc. 136, 6269–6275. doi:10.1021/ja4109787

PubMed Abstract | CrossRef Full Text | Google Scholar

Cheng, Z. S., Zhang, X. M., Zhang, H., Liu, H. Y., Dai, X. F., Liu, G. D., et al. (2022). Binary pentagonal auxetic materials for photocatalysis and energy storage with outstanding performances. Nanoscale 14, 2041–2051. doi:10.1039/d1nr08368f

PubMed Abstract | CrossRef Full Text | Google Scholar

Cheng, Z. S., Zhang, X. M., Zhang, H., Liu, H. Y., Dai, X. F., Liu, G. D., et al. (2022). Two-dimensional auxetic pentagonal materials as water splitting photocatalysts with excellent performances. J. Mat. Sci. 57, 7667–7679. doi:10.1007/s10853-022-07130-x

CrossRef Full Text | Google Scholar

Cimas, Á., Tielens, F., Sulpizi, M., Gaigeot, M. P., and Costa, D. (2014). The amorphous silica-liquid water interface studied by ab initio molecular dynamics (AIMD): Local organization in global disorder. J. Phys. Condens. Matter 26, 244106. doi:10.1088/0953-8984/26/24/244106

PubMed Abstract | CrossRef Full Text | Google Scholar

Deák, P., Aradi, B., Frauenheim, T., Janzén, E., and Gali, A. (2010). Accurate defect levels obtained from the HSE06 range-separated hybrid functional. Phys. Rev. B 81, 153203. doi:10.1103/physrevb.81.153203

CrossRef Full Text | Google Scholar

Fan, Y. S., Xi, X. L., Liu, Y. S., Nie, Z. R., Zhao, L. Y., and Zhang, Q. H. (2021). Regulation of morphology and visible light-driven photocatalysis of WO3 nanostructures by changing pH. Rare Met. 40, 1738–1745. doi:10.1007/s12598-020-01490-6

CrossRef Full Text | Google Scholar

Fujishima, A., and Honda, K. (1972). Electrochemical photolysis of water at a semiconductor electrode. Nature 238, 37–38. doi:10.1038/238037a0

PubMed Abstract | CrossRef Full Text | Google Scholar

Fujishima, A., Rao, T. N., and Tryk, D. A. (2000). Titanium dioxide photocatalysis. J. Photochem. Photobiol. C Photochem. Rev. 1, 1–21. doi:10.1016/s1389-5567(00)00002-2

CrossRef Full Text | Google Scholar

Gonze, X., and Lee, C. Y. (1997). Dynamical matrices, born effective charges, dielectric permittivity tensors, and interatomic force constants from density-functional perturbation theory. Phys. Rev. B 55, 10355–10368. doi:10.1103/physrevb.55.10355

CrossRef Full Text | Google Scholar

Grimme, S. (2006). Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 27, 1787–1799. doi:10.1002/jcc.20495

PubMed Abstract | CrossRef Full Text | Google Scholar

Han, N., Liu, P. Y., Jiang, J., Ai, L. H., Shao, Z. P., and Liu, S. M. (2018). Recent advances in nanostructured metal nitrides for water splitting. J. Mat. Chem. A 6, 19912–19933. doi:10.1039/c8ta06529b

CrossRef Full Text | Google Scholar

Jing, Y., Zhou, Z. P., Zhang, J., Huang, C. B., Li, Y. F., and Wang, F. (2019). SnP2S6 monolayer: A promising 2D semiconductor for photocatalytic water splitting. Phys. Chem. Chem. Phys. 21, 21064–21069. doi:10.1039/c9cp04143e

PubMed Abstract | CrossRef Full Text | Google Scholar

Kan, M., Wang, J. Y., Li, X. W., Zhang, S. H., Li, Y. W., Kawazoe, Y., et al. (2014). Structures and phase transition of a MoS2 monolayer. J. Phys. Chem. C 118, 1515–1522. doi:10.1021/jp4076355

CrossRef Full Text | Google Scholar

Kresse, G., and Furthmuller, J. (1996). Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 54, 11169–11186. doi:10.1103/physrevb.54.11169

PubMed Abstract | CrossRef Full Text | Google Scholar

Kuzmenko, A. B. (2005). Kramers-Kronig constrained variational analysis of optical spectra. Rev. Sci. Instrum. 76, 083108. doi:10.1063/1.1979470

CrossRef Full Text | Google Scholar

Liu, X. F., Zhang, Z. F., Ding, Z., Lv, B., Luo, Z. J., Wang, J. S., et al. (2021). Highly anisotropic electronic and mechanical properties of monolayer and bilayer As2S3. Appl. Surf. Sci. 542, 148665. doi:10.1016/j.apsusc.2020.148665

CrossRef Full Text | Google Scholar

Long, C., Liang, Y., Jin, H., Huang, B. B., and Dai, Y. (2018). PdSe2: Flexible two-dimensional transition metal dichalcogenides monolayer for water splitting photocatalyst with extremely low recombination rate. ACS Appl. Energy Mat. 2, 513–520. doi:10.1021/acsaem.8b01521

CrossRef Full Text | Google Scholar

Perdew, J. P., Burke, K., and Ernzerhof, M. (1996). Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868. doi:10.1103/physrevlett.77.3865

PubMed Abstract | CrossRef Full Text | Google Scholar

Phuc, H. V., Hieu, N. N., Hoi, B. D., Hieu, N. V., Thu, T. V., Hung, N. M., et al. (2018). Tuning the electronic properties, effective mass and carrier mobility of MoS2 monolayer by strain engineering: First-principle calculations. J. Electron. Mat. 47, 730–736. doi:10.1007/s11664-017-5843-8

CrossRef Full Text | Google Scholar

Qiao, J. S., Kong, X. H., Hu, Z. X., Yang, F., and Ji, W. (2014). High-mobility transport anisotropy and linear dichroism in few-layer black phosphorus. Nat. Commun. 5, 4475–4477. doi:10.1038/ncomms5475

PubMed Abstract | CrossRef Full Text | Google Scholar

Rahman, M. Z., Kwong, C. W., Davey, K., and Qiao, S. Z. (2016). 2D phosphorene as a water splitting photocatalyst: Fundamentals to applications. Energy Environ. Sci. 9, 709–728. doi:10.1039/c5ee03732h

CrossRef Full Text | Google Scholar

Saha, S., Sinha, T. P., and Mookerjee, A. (2000). Electronic structure, chemical bonding, and optical properties of paraelectric BaTiO3. Phys. Rev. B 62, 8828–8834. doi:10.1103/physrevb.62.8828

CrossRef Full Text | Google Scholar

Singh, A. K., Mathew, K., Zhuang, H. L., and Hennig, R. G. (2015). Computational screening of 2D materials for photocatalysis. J. Phys. Chem. Lett. 6, 1087–1098. doi:10.1021/jz502646d

PubMed Abstract | CrossRef Full Text | Google Scholar

Sun, Y. F., Cheng, H., Gao, S., Sun, Z. H., Li, Q. H., Liu, Q., et al. (2012). Freestanding tin disulfide single‐layers realizing efficient visible‐light water splitting. Angew. Chem. Int. Ed. Engl. 51, 8857–8861. doi:10.1002/ange.201204675

PubMed Abstract | CrossRef Full Text | Google Scholar

Sun, Y. F., Sun, Z. H., Gao, S., Cheng, H., Liu, Q. H., Piao, J. Y., et al. (2012). Fabrication of flexible and freestanding zinc chalcogenide single layers. Nat. Commun. 3, 1057–7. doi:10.1038/ncomms2066

PubMed Abstract | CrossRef Full Text | Google Scholar

Sun, S. S., Meng, F. C., Wang, H. Y., Wang, H., and Ni, Y. X. (2018). Novel two-dimensional semiconductor SnP3: High stability, tunable bandgaps and high carrier mobility explored using first-principles calculations. J. Mat. Chem. A 6, 11890–11897. doi:10.1039/c8ta02494d

CrossRef Full Text | Google Scholar

Sun, S. S., Meng, F. C., Xu, Y. F., He, J., Ni, Y. X., and Wang, H. Y. (2019). Flexible, auxetic and strain-tunable two dimensional penta-X2C family as water splitting photocatalysts with high carrier mobility. J. Mat. Chem. A 7, 7791–7799. doi:10.1039/c8ta12405a

CrossRef Full Text | Google Scholar

Suntivich, J., May, K. J., Gasteiger, H. A., Goodenough, J. B., and Yang, S. H. (2011). A perovskite oxide optimized for oxygen evolution catalysis from molecular orbital principles. Science 334, 1383–1385. doi:10.1126/science.1212858

PubMed Abstract | CrossRef Full Text | Google Scholar

Tsuji, I., Kato, H., and Kudo, A. (2005). Visible‐light‐induced H2 evolution from an aqueous solution containing sulfide and sulfite over a ZnS-CuInS2-AgInS2 solid‐solution photocatalyst. Angew. Chem. Int. Ed. Engl. 117, 3631–3634. doi:10.1002/ange.200500314

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, X. C., Maeda, K., Thomas, A., Takanabe, K., Xin, G., Carlsson, J. M., et al. (2009). A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nat. Mat. 8, 76–80. doi:10.1038/nmat2317

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, Y., Li, Y., and Chen, Z. F. (2015). Not your familiar two dimensional transition metal disulfide: Structural and electronic properties of the PdS2 monolayer. J. Mat. Chem. C 3, 9603–9608. doi:10.1039/c5tc01345c

CrossRef Full Text | Google Scholar

Xiao, F., Lei, W., Wang, W., Xu, L. L., Zhang, S. L., and Ming, X. (2021). Pentagonal two-dimensional noble-metal dichalcogenide PdSSe for photocatalytic water splitting with pronounced optical absorption and ultrahigh anisotropic carrier mobility. J. Mat. Chem. C 9, 7753–7764. doi:10.1039/d1tc01245b

CrossRef Full Text | Google Scholar

Zhang, W., Chai, C. C., Fan, Q. Y., Sun, M. L., Song, Y. X., Yang, Y. T., et al. (2021). Two-dimensional tetrahex-GeC2: A material with tunable electronic and optical properties combined with ultrahigh carrier mobility. ACS Appl. Mat. Interfaces 13, 14489–14496. doi:10.1021/acsami.0c23017

CrossRef Full Text | Google Scholar

Zhang, Q., Wang, X., and Yang, S. L. (2021). δ-SnS: An emerging bidirectional auxetic direct semiconductor with desirable carrier mobility and high-performance catalytic behavior toward the water-splitting reaction. ACS Appl. Mat. Interfaces 13, 31934–31946. doi:10.1021/acsami.1c03650

CrossRef Full Text | Google Scholar

Zhang, Y. F., Wei, J. M., Liu, T. Y., Zhong, Z., Luo, Z. J., Xiao, W. J., et al. (2022). Tunable properties of ZnSe/graphene heterostructure as a promising candidate for photo/electro-catalyst applications. Appl. Surf. Sci. 574, 151679. doi:10.1016/j.apsusc.2021.151679

CrossRef Full Text | Google Scholar

Zhao, P., Ma, Y. D., Lv, X. S., Li, M. M., Huang, B. B., and Dai, Y. (2018). Two-dimensional III2-VI3 materials: Promising photocatalysts for overall water splitting under infrared light spectrum. Nano Energy 51, 533–538. doi:10.1016/j.nanoen.2018.07.010

CrossRef Full Text | Google Scholar

Zhuang, H. L., and Hennig, R. G. (2013). Computational search for single-layer transition-metal dichalcogenide photocatalysts. J. Phys. Chem. C 117, 20440–20445. doi:10.1021/jp405808a

CrossRef Full Text | Google Scholar

Zhuang, H. L., and Hennig, R. G. (2013). Single-layer group-III monochalcogenide photocatalysts for water splitting. Chem. Mat. 25, 3232–3238. doi:10.1021/cm401661x

CrossRef Full Text | Google Scholar

Keywords: two-dimensional materials, transition metal dichalcogenides, water splitting photocatalyst, high mobility, density functional theory

Citation: Wang F, Cheng Z, Zhang X, Xie C, Liu F, Chang C and Liu G (2022) A novel two-dimensional transition metal dichalcogenide as water splitting photocatalyst with excellent performances. Front. Chem. 10:1003027. doi: 10.3389/fchem.2022.1003027

Received: 25 July 2022; Accepted: 03 August 2022;
Published: 25 August 2022.

Edited by:

Guangzhao Wang, Yangtze Normal University, China

Reviewed by:

Junli Chang, Southwest University, China
Liu Xuefei, Guizhou Normal University, China

Copyright © 2022 Wang, Cheng, Zhang, Xie, Liu, Chang and Liu. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Zishuang Cheng, czs19950627@163.com; Chuntao Chang, changct@dgut.edu.cn

Download