Next Article in Journal
Fabrication and Optimization of a Lipase Immobilized Enzymatic Membrane Bioreactor based on Polysulfone Gradient-Pore Hollow Fiber Membrane
Next Article in Special Issue
Photocatalytic Hydrogen Evolution
Previous Article in Journal
Screening of Additives to Ni-Based Methanation Catalyst for Enhanced Anti-Sintering Performance
Previous Article in Special Issue
Synthesis of Spherical TiO2 Particles with Disordered Rutile Surface for Photocatalytic Hydrogen Production
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photoelectrochemical Hydrogen Evolution and CO2 Reduction over MoS2/Si and MoSe2/Si Nanostructures by Combined Photoelectrochemical Deposition and Rapid-Thermal Annealing Process

Department of Chemistry, Chungnam National University, Daejeon 34134, Korea
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(6), 494; https://doi.org/10.3390/catal9060494
Submission received: 17 April 2019 / Revised: 24 May 2019 / Accepted: 25 May 2019 / Published: 28 May 2019
(This article belongs to the Special Issue Photocatalytic Hydrogen Evolution)

Abstract

:
Diverse methods have been employed to synthesize MoS2 and MoSe2 catalyst systems. Herein, a combined photoelectrochemical (PEC) deposition and rapid-thermal annealing process has first been employed to fabricate MoS2 and MoSe2 thin films on Si substrates. The newly developed transition-metal dichalcogenides were characterized by scanning electron microscopy, Raman spectroscopy and X-ray photoelectron spectroscopy. PEC hydrogen evolution reaction (HER) was demonstrated in an acidic condition to show a PEC catalytic performance order of MoOx/Si < MoS2/Si << MoSe2/Si under the visible light-on condition. The HER activity (4.5 mA/cm2 at −1.0 V vs Ag/AgCl) of MoSe2/Si was increased by 4.8× compared with that under the dark condition. For CO2 reduction, the PEC activity was observed to be in the order of MoS2/Si < MoOx/Si << MoSe2/Si under the visible light-on condition. The reduction activity (0.127 mA/cm2) of MoSe2/Si was increased by 9.3× compared with that under the dark condition. The combined electrochemical deposition and rapid-thermal annealing method could be a very useful method for fabricating a thin film state catalytic system perusing hydrogen production and CO2 energy conversion.

Graphical Abstract

1. Introduction

Transition-metal dichalcogenides (TMDCs) have widely been studied for applications to energy and environment such as hydrogen evolution and CO2 reduction [1,2,3,4,5,6,7,8,9,10]. Especially, molybdenum disulfide and diselenide (MoS2 and MoSe2) materials with two-dimensional character have been synthesized using diverse synthesis methods for their applications [11]. Ye et al. employed a chemical vapor deposition (CVD) method to synthesize monolayer MoS2 followed by oxygen plasma treatment or hydrogen annealing. They showed that hydrogen evolution reaction (HER) activity (e.g., onset potential and current density) was increased substantially by engineering the defects [12]. To increase HER activity of MoS2 or MoSe2, various defect engineering techniques have been employed, which include laser irradiation [13], ion irradiation [14,15] and NaClO chemical etching [16]. Li et al. examined various defect sites of MoS2 such as edge sites, S vacancies and grain boundaries and showed that edge sites and S vacancies (with optimal vacancy density of 7–10%) were main HER active sites [17]. Chang et al. employed lithium molten salts to synthesize 2H- and 1T-MoS2 monolayers at calcination temperatures of 400~600 °C and above 1000 °C, respectively [18]. They observed that metallic 1T-MoS2 showed a higher HER activity than that of semiconducting 2H-MoS2. Two step hydrothermal method was employed to synthesize 1T@2H-MoSe2 nanosheets, which showed a higher HER activity [19]. Wang et al. prepared MoS2 nanosheets by mechanical exfoliation, transferred onto a SiO2 surface and made a HER device [20]. Afterwards, they showed that HER activity was increased by applying an extra positive electric field. Guo et al. prepared oxygen-incorporated MoS2 sheets on graphene by a hydrothermal method [21] and showed that the active edge sites and conductivity were increased by oxygen-incorporation and the electrical transfer was increased by hybridization. Consequently, the HER activity was found to be substantially increased. Zhu et al. fabricated h-MoO3/1T-MoS2 heterostructures and tested photoelectrocatalytic HER activity to show better activity compared with those of 1T@2H-MoS2 and α-MoO3/MoS2 [22]. A two-step (MoO3 + H2 → MoO2 and MoO2 + Se vapor → MoSe2) chemical vapor deposition (CVD) process was employed to fabricate vertically aligned core–shell MoO2/MoSe2 nanosheet arrays which showed better HER activity than those of MoO2 and MoSe2 [23]. Electrochemical CO2 reduction is another potential application area for MoS2 and MoSe2 [4,24,25,26,27,28,29]. Francis et al. tested a single crystal MoS2 electrode for CO2 reduction and showed a Faradaic efficiency of ~3.5% at −0.59 V (vs Reversible Hydrogen Electrode) for a reduction product of 1-propanol [4]. Asadi et al. reported that electrochemical CO2 reduction product for vertically aligned MoS2 in an ionic liquid was found to be CO with a CO2 reduction current density of 130 mA cm−2 at −0.764 V [28,29].
Electrodeposition has popularly been employed for the cheap fabrication of thin films on a substrate, where major factors determining the nature of thin film include electrolyte, pH, deposition time and an applied potential [30]. For electrodeposition (under the dark condition) of Mo oxides on a substrate, some studies have been reported [31,32,33,34]. However, no studies have been reported for photoelectrochemical deposition of Mo oxides. Petrova et al. used Al substrates for electrodeposition of Mo oxides in Mo ion electrolyte (Mo(NH4)6Mo7O24·4H2O, 20 g/L) at pH of 8–10 adjusted by a NH3-CH3COONH4 buffer [31]. Pd–MoOx catalyst on glassy carbon electrode was reported to be fabricated by electrodeposition at potential ranges between −0.73 and +0.2 V in a mixed solution of 2 mM PdCl2, 15 mM Na2MoO4 and 0.2 M HCl [32]. The dominant oxidation state of Mo was found to be +6. Uniform Mo oxide (+6, +5 and +4 oxidation states) nanostructure arrays (nanotubes at pH = 2.7 and nanowires at pH = 5.5) were prepared by a template electrodeposition in Mo ion electrolyte ((NH4)6Mo7O24·4H2O, 50 g/L) [33]. Electroless-photochemical deposition (PCD) has also been demonstrated for the preparation of metal sulfide thin films. Soundeswaran et al. prepared CdS films on indium tin oxide (ITO) glass using 1–10 mM Cd(CH3COO)2 and 100 mM Na2S2O3 solution at pH = 3.0–4.5 under irradiation of ultraviolet (UV) light (100 mW/cm2) [35]. For this reaction, S and electrons were initially formed by UV-excitation of S2O32- ions and reacted with Cd metal ions to form CdS. Podder et al. prepared CuxS thin films on ITO glass using a similar method [36].
Herein, a new methodology of combined photoelectrochemical deposition (Mo6+ + 6OH → MoO3 + 3H2O accelerated by an enhanced photocurrent) and rapid-thermal annealing (RTA) sulfurization (or selenization) process (2MoO3 + 4S or Se → 2MoS2 or MoSe2 + 3O2) was introduced to fabricate thin MoS2 and MoSe2 films on Si substrates. HER and CO2 reduction tests were demonstrated to show a potential applicability to energy and environment. A major advantage of the combined method is time-saving and cost effective. Another advantage is morphology and thickness-controlled by tuning applied voltage, deposition time and the electrolyte condition and so forth. Overall, the present developed method could be further improved and widely used for developing better thin film systems for diverse application areas.

2. Results

Surface morphology, crystal phase formation and surface chemical states were examined using scanning electron microscopy (SEM), Raman and X-ray photoelectron spectroscopy (XPS), respectively. Hydrogen evolution reaction (HER) and CO2 reduction were tested using the three electrode system. The experimental results are described below.

2.1. SEM Morphology

Figure 1 shows the SEM images of MoOx/Si, MoS2/Si and MoSe2/Si samples. For the as-photoelectrodeposited MoOx/Si sample, larger (200~400 nm diameter) and smaller (< 50 nm) nanoparticles were formed on the Si surface. Two different particle size distributions may be due to mixed Volmer-Weber island growth mode and Ostwald ripening process. The oxidation states were confirmed by XPS, discussed below. Under the dark condition, no electrodeposition was observed. For the SEM image of MoS2 formed by RTA process of MoOx/Si sample, two different size distributions were also observed for the sample as expected. For the SEM image of MoSe2 formed by RTA process of MoOx/Si sample, the surface morphology showed more uniform nanostructure.

2.2. Raman Spectroscopy

Raman spectra (Figure 2) were obtained to examine the detailed crystal phase formation for the catalyst systems. For all the samples, the strongest peak was commonly observed at 524 cm−1 (not shown in the Figure), attributed to Si used as a support [37]. A weak peak at ~300 cm−1 for MoOx/Si was due to the phonon mode of Si [37]. No Raman peaks of Mo oxides were observed, indicating that the as-photoelectrodeposited sample was ultrathin and/or amorphous. For MoS2 sample (Figure 2B), two Raman peaks were observed at 385 and 411 cm−1, attributed to in-plane Mo-S (E2g) and out-of-plane Mo-S (A1g) vibration modes of hexagonal MoS2 [38,39]. For MoSe2 sample (Figure 2C), a strong Raman peak was observed at 238 cm−1, attributed to Mo-Se A1g mode. Other two peaks at 169 and 286 cm−1 were assigned to E1g and E2g modes, respectively [40,41]. The A1g mode was found to be stronger than the E2g mode. Overall, based on the Raman data, the RTA process was found to be efficient for the fabrication of MoS2 and MoSe2.

2.3. X-ray Photoelectron Spectroscopy

X-ray photoelectron spectra (XPS) of the three catalyst systems are displayed in Figure 3. For the as-photoelectrodeposited MoOx/Si, Mo and O elements were dominantly observed. Mo 3d3/2 and 3d5/2 XPS peaks were observed at 235.5 and 232.5 eV, respectively with a spin-orbit splitting energy of 3.0 eV. This is attributed to Mo6+ oxidation state of MoO3 [42]. Furthermore, the other Mo 3d3/2 and 3d5/2 XPS peaks were observed at 233.8 and 230.8 eV, respectively with a spin-orbit splitting energy of 3.0 eV. This could be due to Mo5+ oxidation state. Based on the Mo 3d XPS fitting, 42% and 52% of Mo were 6+ and 5+ oxidation states, respectively. The corresponding broad O 1s peak was observed at 530.6 eV, attributed to the lattice oxygen of Mo oxides (MoOx). For the Mo 3d XPS of the MoS2/Si sample, Mo 3d3/2 and 3d5/2 XPS peaks were observed at 232.2 and 229.0 eV, respectively with a spin-orbit splitting energy of 3.2 eV. This is attributed to Mo4+ oxidation state of MoS2 [43,44]. A small peak at 226.3 eV was due to S 2s [45]. The corresponding S 2p1/2 and S 2p3/2 XPS peaks were observed at 163.1 eV and 161.9 eV, respectively with a spin-orbit splitting of 1.2 eV. This is in good agreement with the literature reported by Jian et al. for 1T phase MoS2 [44]. For the Mo 3d XPS of the MoSe2/Si sample, Mo 3d3/2 and 3d5/2 peaks were observed at 231.7 and 228.6 eV, respectively with a spin-orbit splitting energy of 3.1 eV. This is in good agreement with +4 oxidation state of MoSe2 [5]. The corresponding Se 3d5/2 and 3d3/2 peaks were found to be located 55.1 and 54.2 eV, respectively. The O 1s XPS peak was observed at 533.1 eV, attributed adsorbed surface oxygen [42]. For MoOx/Si sample, the Mo:O ratio was estimated to be 1:4.1. The overestimated oxygen was plausibly due to surface oxygen such as H2O and OH. Mo:O ratio was estimated to be 1: 2.3 by only considering the lattice O 1s signal. For MoS2/Si sample, Mo:S ratio was calculated to be 1:1.95, very close to MoS2. For MoSe2/Si sample, Mo:Se ratio was calculated to be 1:2.15, close to MoSe2. For the valence band (VB) XPS spectra, the density of states (DOS) near the Fermi level was more discernible for MoS2 and MoSe2 reflecting metallic/semiconducting states.

2.4. Photoelectrochemcial Hydrogen Evolution

Three different catalyst systems of MoOx/Si, MoS2/Si and MoSe2/Si were tested for hydrogen evolution reaction in 0.1 M H2SO4 electrolyte. Linear sweep voltammetry (LSV) curves (Figure 4) were obtained from +0.2 V to −1.0 V at a scan rate of 10 mV/sec after full nitrogen gas purging under the dark and visible light exposure conditions. Under the dark condition, the current density at -1.0 V (vs Ag/AgCl) showed the order of MoOx/Si (0.08 mA/cm2) < MoS2/Si (0.11 mA/cm2) << MoSe2/Si (0.86 mA/cm2) while the order changed to MoS2/Si (0.51 mA/cm2) < MoOx/Si (0.64 mA/cm2) << MoSe2/Si (4.3 mA/cm2) under the visible light exposure condition. They all commonly showed an increase in HER CD under visible light exposure. The inset in Figure 4 displays LSV curves taken under the light ON-and-OFF condition during the scan. It clearly showed that all three samples have photocatalytic activity.

2.5. Photoelectrochemcial CO2 Reduction

For CO2 reduction in 0.1 M NaHCO3 electrolyte, the LSV measurements (Figure 5) were obtained from +0.2 V to −1.0 V upon full N2 and CO2 gas purging at a scan rate of 10 mV/sec under the dark and visible light exposure conditions, respectively. Upon full N2 gas purging in the electrolyte under dark condition (inset in Figure 5A), the current densities (CDs) at −1.0 V were observed to be 0.0043, 0.0041 and 0.012 mA/cm2 for MoOx/Si, MoS2/Si and MoSe2/Si, respectively. Under the visible light exposure condition (Figure 5A), the CDs were found to be drastically increased to 0.037, 0.011 and 0.018 mA/cm2 for MoOx/Si, MoS2/Si and MoSe2/Si, respectively. Upon full CO2 gas purging in the electrolyte under the dark condition (inset in Figure 5B), the CDs at −1.0 V were found to be 0.007, 0.009 and 0.014 mA/cm2 for MoOx/Si, MoS2/Si and MoSe2/Si, respectively. Under the visible light exposure condition (Figure 5B), the CDs were found to be increased to 0.043, 0.023 and 0.13 mA/cm2 for MoOx/Si, MoS2/Si and MoSe2/Si, respectively. The CDs were increased by 6.0, 2.6 and 9.4×, respectively upon visible light exposure. The inset in Figure 5B displays LSV curves taken under the light ON-and-OFF condition during the scan. It clearly showed that all three samples also have photocatalytic activity.

3. Discussion

The photo-electrochemical deposition method was first successfully employed to fabricate MoOx on a Si support. Upadhyay et al. reported MoOx nanoparticles on a stainless steel support prepared by electrodeposition at -1.0 V in a mixed solution of 0.1 M Na2MoO4 and 0.1 M NH4NO3 [46]. They reported oxidation states of Mo6+ and Mo5+ for MoOx and the corresponding O 1s peak at 530.6 eV. This is in good agreement with our present results. Liu et al. performed electrodeposition of MoOx films on a Ti (130 nm)/Si substrate in a mixed solution of 0.1 M Na2MoO4, 0.1 M Na2EDTA and 0.1 M CH3COONH4 [47]. They concluded that the as-electrodeposited MoOx film (with oxidation states of Mo4+ and Mo5+) was amorphous. This is also in good agreement with the present result, as discussed above (Raman spectra in Figure 2). Based on the SEM image, the photoelectrochemical deposition of MoOx on a Si support appeared to be occurred through Volmer-Weber island growth process [48]. Small islands were initially formed on the entire surface and then larger islands subsequently were grown. For the SEM images of MoS2 and MoSe2 by the RTA process, the morphology of MoS2 was more similar to that of MoOx, compared with that of MoSe2. This indicates that less energy was required for the formation of MoS2, compared with that for MoSe2. Overall, the RTA process was found to be efficient for the formation of MoS2 and MoSe2 without much impacting the original morphology of electrodeposited MoOx.
For HER in 0.1 M H2SO4, MoSe2/Si showed a much higher electrochemical activity (or current density) than MoS2/Si. For HER mechanism in the acidic condition, adsorbed hydrogen is known to be formed via H3O+ + e→ Had + H2O. Then, hydrogen is generated via Had + H3O+ + e → H2 + H2O or Had + Had → H2 [23]. In the mechanism, hydrogen adsorption Gibbs free energy, ΔGHX (X = S or Se) is known to play a major role in determining the activity [10]. The optimal condition is ΔGHX = 0 eV and ΔGHX of MoSe2 is closer to the optimal condition than that of MoS2. The HER activity of MoSe2 has commonly been reported to be higher than that of MoS2 [49]. This is in good consistent with the present result. For HER of nanoflowers-like MoS2 and MoSe2 materials on GC electrodes in 0.5 M H2SO4, Ravikumar et al. reported that the activity (11 mA/cm2 at 0.3 V vs RHE) of MoSe2 showed a higher than that (7 mA/cm2 at 0.3 V vs RHE) of MoS2, attributed to higher electrical properties and defects [50], in good consistent with the present result. Evidently, based on the DOS near the Fermi level as discussed in Figure 3, the enhanced electronic conductivity could play an important role in HER and CO2 reduction performances [22]. Because the Gibbs free energy is also known to be dependent on morphology (e.g., defects and edge sites) the catalyst fabrication methods is important for improving a catalyst activity. Upon visible light exposure on the catalyst surface during the LSV, an increased CD was commonly been observed. The enhancement factors (light ON/OFF CD ratio) for HER at −1.0 V upon visible light irradiation were observed to be 8.0, 4.7 and 5.0 for MoOx/Si, MoS2/Si and MoSe2/Si, respectively (Figure 6). The photocatalytic HER activity under visible light was due to photo-generated electron by absorption of light in the visible region [22,42]. As mentioned above, in HER mechanism electron plays a crucial role in generation of hydrogen. Overall, the photocatalytic activity is all dependent on material nature (e.g., electrical conductivity), morphology, surface natures (e.g., defects) and light absorption efficiency. For the splitting of water, the molar stoichiometric ratio of H2/O2 is ideally 2.0 [51] assuming that no other side electrochemical reactions are involved for MoS2 and MoSe2 [1]. Before further discussion, it should be here mentioned that our conclusion was based only on the CD. The H2/O2 production ratios and CO2 reduction products are needed to be further examined by gas chromatography [51].
For electrochemical CO2 reduction, the current density was commonly been increased in CO2-purged 0.1 M NaHCO3 electrolyte, compared with that in N2-purged 0.1 M NaHCO3 electrolyte. This indicates that CO2 reduction was occurred for all the samples. For CO2 reduction (Figure 6), under the dark condition, the enhancement factors before (only N2 bubbling) and after CO2 bubbling were observed to be 1.6, 2.2, 1.2 for MoOx/Si, MoS2/Si and MoSe2/Si, respectively. Under the visible light-on condition, the enhancement factors before (only N2 bubbling) and after CO2 bubbling were observed to be 1.1, 2.1 and 7.2 for MoOx/Si, MoS2/Si and MoSe2/Si, respectively. The CDs upon light exposure were increased by 6.0, 2.6 and 9.4× for MoOx/Si, MoS2/Si and MoSe2/Si, respectively. Overall, MoSe2 showed the most dramatic photo-electrochemical CO2 reduction efficiency. For MoSe2, photogenerated electrons are created by light absorption in the visible region and electron-hole recombination is suppressed by good electron transport [1]. Consequently, the CD by photoelectrochemical CO2 reduction is enhanced.

4. Materials and Methods

For photoelectrochemical Mo deposition (MoOx/Si), a three-electrode (Ag/AgCl reference, Pt wire counter and Si working electrode) electrochemical cell was used using a VersaSTAT3 (Princeton Applied Research, Oak Ridge, TN, USA) potentiostat galvanostat. For the preparation of a Si working support electrode before electrodeposition, a single-side polished Si (100) wafer (B-doped p-type, thickness of 525 ± 20 μm, resistivity of 1–10 Ω·cm, 2 cm × 0.5 cm) was used as the support, cleaned in 2% HF solution to remove oxide layer and washed with deionized water by sonication. The electrolyte was a mix of 15 mM Na2MoO4 (99.0% extra pure, Samchun Chem. Co., Seoul, Republic of Korea), 1.0 M NaCl and 1.0 M NH3Cl, where pH was adjusted to 9.2 using NH3OH (28~30%, Samchun Chem. Co., Seoul, Republic of Korea). The photoelectrochemical deposition was performed at an applied potential of −1.5V (versus Ag/AgCl electrode) for 10 sec under visible light exposure onto the working electrode. In the present study, we only showed the samples prepared at fixed parameters among different applied potentials and deposition times. No efficient (or less uniform) Mo deposition occurred under the dark condition although the applied voltages and deposition times were varied. CDs of −0.10 and −0.92 mA/cm2 were measured at −1.5 V under dark and visible light, respectively. The CD was enhanced by 9–10× in the potential ranges from −1.0 to −2.0 V. For the preparation of MoS2 and MoSe2 on Si support, a rapid-thermal annealing (RTA) method was employed using a LABSYS RTP-1200 (Nextron Co., Ltd., Busan, Republic of Korea). For this, a MoOx/Si substrate was placed on a quartz plate (15 mm × 20 mm) in the RTA chamber. Sulfur (S.P.C. GR reagent, Shinyo Pure Chem. Co., Ltd., Hyogo, Japan) or Selenium (99.5+%, 100mesh, Sigma-Aldrich, St. Louis, MO, USA) powder (~0.02 g) was placed below the substrate. The chamber was maintained in 5% H2 in He balance. The temperature heating rate was 12 °C/sec and the time was 6 sec at the maximum temperature of 700 °C. The surface morphology of the prepared samples was examined using a Hitachi S-4800 (Tokyo, Japan) scanning electron microscope at an electron acceleration voltage of 10.0 kV. Raman spectra with 514 nm laser line were obtained using a LabRAM HR-800 microRaman spectrometer (HORIBA Jobin Yvon, Kyoto, Japan). X-ray photoelectron spectra were taken using a Thermo Scientific K-Alpha+ X-ray photoelectron spectrometer with micro-focused monochromatic Al Kα X-ray source and a hemispherical energy analyzer. XPS spectra curve fitting was performed using a XPSPEAK ver. 4.1 software. For XPS element quantification, XPS sensitivity factors of 2.75, 0.66, 0.54, and 0.67 were used for Mo 3d, O 1s, S 2p and Se 3d, respectively [52]. For photoelectrochemical HER and CO2 reduction, a three-electrode electrochemical cell was also used using a VersaSTAT3 potentiostat/galvanostat. For HER, nitrogen gas was fully purged into the electrolyte (0.1 M H2SO4 solution) to minimize an effect of dissolved oxygen. Linear sweep voltammetry (LSV) was carried out at a scan rate of 10 mV s−1 from +0.2 V to −1.0 V under dark and visible light exposure conditions. A white LED USB Flashlight (A-10, Teckmedia) was used for visible light (400~700 nm) [53]. For CO2 reduction experiment in 0.1 M NaHCO3 solution, LSV was conducted after N2 gas purging at a scan rate of 10 mV s−1 from +0.2 V to −1.0 V under dark and visible light irradiation conditions. The same LSV experiment was also conducted after CO2 gas purging into the electrolyte to examine the CO2 effect.

5. Conclusions

In this work, a combined photoelectrochemical deposition and rapid-thermal annealing method was first been employed to fabricate MoS2 and MoSe2 thin films on Si substrates. Photoelectrochemical HER and CO2 reduction were demonstrated for the newly developed catalytic systems. The main results are as follows:
  • Mo oxides were successfully electrodeposited on a Si support under visible light exposure. Under the dark condition, the electrochemical deposition was less efficient.
  • A rapid-thermal annealing method was successfully introduced for the electrodeposited MoOx/Si to fabricate MoS2/Si and MoSe2/Si catalyst systems. Other impurity phases were not detected mainly based on the Raman and XPS results. The maximum temperature was achieved by rapid heating to 700 °C of S or Se powers on the MoOx/Si and maintained for 6 sec.
  • HER tests in 0.1 M H2SO4 electrolyte showed a catalytic activity order of MoOx/Si < MoS2/Si << MoSe2/Si under dark and visible light-on conditions. The HER activity (4.5 mA/cm2 at −1.0 V vs Ag/AgCl) of MoSe2/Si was increased by 4.8× compared with that under the dark condition.
  • CO2 reduction tests in 0.1 M NaHCO3 electrolyte showed a catalytic activity order of MoOx/Si, MoS2/Si << MoSe2/Si and MoS2/Si < MoOx/Si << MoSe2/Si under dark and visible light-on conditions, respectively. The reduction activity (0.127 mA/cm2) of MoSe2/Si was increased by 9.3× compared with that under the dark condition.
The newly developed catalyst preparation method could be very useful for developing thin film catalyst systems for diverse application areas.

Author Contributions

Y.S. designed the experiments and wrote the paper; S.H. performed the experiments; C.K.R. provided valuable idea for obtaining the data.

Funding

This research was funded by the National Research Foundation of Korea (NRF) grant funded by the Korean government (MEST), grant number NRF-2016R1D1A3B04930123. The APC was funded by the National Research Foundation of Korea (NRF).

Acknowledgments

This work was financially supported by the National Research Foundation of Korea (NRF) grant funded by the Korean government (MEST) (NRF-2016R1D1A3B04930123).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, L.; Liu, P.; Li, J.; Xiang, B. Two-dimensional material molybdenum disulfides as electrocatalysts for hydrogen evolution. Catalysts 2017, 7, 285. [Google Scholar] [CrossRef]
  2. Theerthagiri, J.; Senthil, R.A.; Senthilkumar, B.; Polu, A.R.; Madhavan, J.; Ashokkumar, M. Recent advances in MoS2 nanostructured materials for energy and environmental applications—A review. J. Sol. Stat. Chem. 2017, 252, 43–71. [Google Scholar] [CrossRef]
  3. Eftekhari, A. Molybdenum diselenide (MoSe2) for energy storage, catalysis and optoelectronics. Appl. Mater. Today 2017, 8, 1–17. [Google Scholar] [CrossRef]
  4. Francis, S.A.; Velazquez, J.M.; Ferrer, I.M.; Torelli, D.A.; Guevarra, D.; McDowell, M.T.; Sun, K.; Zhou, X.; Saadi, F.H.; John, J.; et al. Reduction of aqueous CO2 to 1-propanol at MoS2 Electrodes. Chem. Mater. 2018, 30, 4902–4908. [Google Scholar] [CrossRef]
  5. Guo, W.; Chen, Y.; Wang, L.; Xu, J.; Zeng, D.; Peng, D.L. Colloidal synthesis of MoSe2 nanonetworks and nanoflowers with efficient electrocatalytic hydrogen-evolution activity. Electrochim. Acta 2017, 231, 69–76. [Google Scholar]
  6. Guo, H.; Jin, C.; Liu, X.; Liu, J.; Vasileff, A.; Jiao, Y.; Zheng, Y.; Qiao, S. Emerging Two-dimensional nanomaterials for electrocatalysis. Chem. Rev. 2018, 118, 6337–6408. [Google Scholar]
  7. Mao, J.; Wang, Y.; Zheng, Z.; Deng, D. The rise of two-dimensional MoS2 for catalysis. Front. Phys. 2018, 13, 138118. [Google Scholar] [CrossRef]
  8. Sun, Z.; Ma, T.; Tao, H.; Fan, Q.; Han, B. Fundamentals and challenges of electrochemical CO2 reduction using two-dimensional materials. Chem 2017, 3, 560–587. [Google Scholar] [CrossRef]
  9. Gupta, U.; Rao, C.N.R. Hydrogen generation by water splitting using MoS2 and other transition metal dichalcogenides. Nano Energy 2017, 41, 49–65. [Google Scholar] [CrossRef]
  10. Eftekhari, A. Electrocatalysts for hydrogen evolution reaction. Int. J. Hydrog. Eng. 2017, 42, 11053–11077. [Google Scholar] [CrossRef]
  11. Cai, Z.; Liu, B.; Zou, X.; Cheng, H. Chemical vapor deposition growth and applications of two-dimensional materials and their heterostructures. Chem. Rev. 2018, 118, 6091–6133. [Google Scholar] [CrossRef]
  12. Ye, G.; Gong, Y.; Lin, J.; Li, B.; He, Y.; Pantelides, S.T.; Zhou, W.; Vajtai, R.; Ajayan, P.M. Defects engineered monolayer MoS2 for improved hydrogen evolution reaction. Nano Lett. 2016, 16, 1097–1103. [Google Scholar] [CrossRef]
  13. Meng, C.; Lin, M.C.; Du, X.; Zhou, Y. Molybdenum disulfide modified by laser irradiation for catalyzing hydrogen evolution. ACS Sustain. Chem. Eng. 2019, 7, 6999–7003. [Google Scholar] [CrossRef]
  14. He, Z.; Zhao, R.; Chen, X.; Chen, H.; Zhu, Y.; Su, H.; Huang, S.; Xue, J.; Dai, J.; Cheng, S.; et al. Defect engineering in single-layer MoS2 using heavy ion irradiation. ACS Appl. Mater. Interfaces 2018, 10, 42524–42533. [Google Scholar] [CrossRef]
  15. Xia, B.; Wang, T.; Jiang, X.; Zhang, T.; Li, J.; Xiao, W.; Xi, P.; Gao, D.; Xue, D.; Ding, J. Ar2+ beam irradiation-induced multivancancies in MoSe2 nanosheet for enhanced electrochemical hydrogen evolution. ACS Energy Lett. 2018, 3, 2167–2172. [Google Scholar] [CrossRef]
  16. Zhang, P.; Xiang, H.; Tao, L.; Dong, H.; Zhou, Y.; Hu, T.S.; Chen, X.; Liu, S.; Wang, S.; Garaj, S. Chemically activated MoS2 for efficient hydrogen production. Nano Energy 2019, 57, 535–541. [Google Scholar] [CrossRef]
  17. Li, G.; Zhang, D.; Qiao, Q.; Yu, Y.; Peterson, D.; Zafar, A.; Kumar, R.; Curtarolo, S.; Hunte, F.; Shannon, S.; et al. All the catalytic active sites of MoS2 for hydrogen evolution. J. Am. Chem. Soc. 2016, 138, 16632–16638. [Google Scholar] [CrossRef]
  18. Chang, K.; Hai, X.; Pang, H.; Zhang, H.; Shi, L.; Liu, G.; Liu, H.; Zhao, G.; Li, M.; Ye, J. Targeted synthesis of 2H- and 1T-phase MoS2 monolayers for catalytic hydrogen evolution. Adv. Mater. 2016, 28, 10033–10041. [Google Scholar] [CrossRef]
  19. Zhang, J.; Wang, T.; Liu, P.; Liu, Y.; Ma, J.; Gao, D. Enhanced catalytic activities of metal-phase-assisted 1T@2H-MoSe2 nanosheets for hydrogen evolution. Electrochim. Acta 2016, 217, 181–186. [Google Scholar] [CrossRef]
  20. Wang, J.; Yan, M.; Zhao, K.; Liao, X.; Wang, P.; Pan, X.; Yang, W.; Mai, L. Field effect enhanced hydrogen evolution reaction of MoS2 nanosheets. Adv. Mater. 2016, 29, 1604464. [Google Scholar] [CrossRef]
  21. Guo, J.; Li, F.; Sun, Y.; Zhang, X.; Tang, L. Oxygen-incorporated MoS2 ultrathin nanosheets grown on graphene for efficient electrochemical hydrogen evolution. J. Power Sources 2015, 291, 195–200. [Google Scholar] [CrossRef]
  22. Zhu, C.; Xu, Q.; Liu, W.; Ren, Y. CO2-assisted fabrication of novel heterostructures of h-MoO3/1T-MoS2 for enhanced photoelectrocatalytic performance. Appl. Surf. Sci. 2017, 425, 56–62. [Google Scholar] [CrossRef]
  23. Chen, X.; Liu, G.; Zheng, W.; Feng, W.; Cao, W.; Hu, W.; Hu, P. Vertical 2D MoO2/MoSe2 core–shell nanosheet arrays as high performance electrocatalysts for hydrogen evolution reaction. Adv. Funct. Mater. 2016, 26, 8537–8544. [Google Scholar] [CrossRef]
  24. Hong, X.; Chan, K.; Tsai, C.; Nørskov, J.K. How doped MoS2 breaks transition-metal scaling relations for CO2 electrochemical reduction. ACS Catal. 2016, 6, 4428–4437. [Google Scholar] [CrossRef]
  25. Chan, K.; Tsai, C.; Hansen, H.A.; Nørskov, J.K. Molybdenum sulfides and selenides as possible electrocatalysts for CO2 reduction. ChemCatChem 2014, 6, 1899–1905. [Google Scholar] [CrossRef]
  26. Ji, Y.; Norskov, J.K.; Chan, K. Scaling relations on basal plane vacancies of transition metal dichalcogenides for CO2 reduction. J. Phys. Chem. C 2019, 123, 4256–4261. [Google Scholar] [CrossRef]
  27. Jia, P.; Guo, R.; Pan, W.; Huang, C.; Tang, J.; Liu, X.; Qin, H.; Xu, Q. The MoS2/TiO2 heterojunction composites with enhanced activity for CO2 photocatalytic reduction under visible light irradiation. Colloids Surf. A 2019, 570, 306–316. [Google Scholar] [CrossRef]
  28. Asadi, M.; Kumar, B.; Behranginia, A.; Rosen, B.A.; Baskin, A.; Repnin, N.; Pisasale, D.; Phillips, P.; Zhu, W.; Haasch, R.; et al. Robust carbon dioxide reduction on molybdenum disulphide edges. Nat. Commun. 2014, 5, 4470. [Google Scholar] [CrossRef] [Green Version]
  29. Asadi, M.; Kim, K.; Liu, C.; Addepalli, A.V.; Abbasi, P.; Yasaei, P.; Phillips, P.; Behranginia, A.; Cerrato, J.M.; Haasch, R.; et al. Nanostructured transition metal dichalcogenide electrocatalysts for CO2 reduction in ionic liquid. Science 2016, 353, 467–470. [Google Scholar] [CrossRef]
  30. Bicelli, L.P.; Bozzini, B.; Mele, C.; D’Urzo, L. A review of nanostructural aspects of metal electrodeposition. Int. J. Electrochem. Sci. 2008, 3, 356–408. [Google Scholar]
  31. Petrova, M.L.; Bojinov, M.; Gadjov, I.H. Electrodeposition of molybdenum oxides from weakly alkaline ammonia-molybdate electrolytes. Bulg. Chem. Commun. 2011, 43, 60–63. [Google Scholar]
  32. Li, R.; Hao, H.; Huang, T.; Yu, A. Electrodeposited Pd–MoOx catalysts with enhanced catalytic activity for formic acid electrooxidation. Electrochim. Acta 2012, 76, 292–299. [Google Scholar] [CrossRef]
  33. Inguanta, R.; Spanò, T.; Piazza, S.; Sunseri, C.; Barreca, F.; Fazio, E.; Neri, F.; Silipigni, L. Electrodeposition and characterization of Mo oxide nanostructures. Chem. Eng. Trans. 2015, 43, 685–690. [Google Scholar]
  34. Banica, R.; Barvinschi, P.; Vaszilcsin, N.; Nyari, T. A comparative study of the electrochemical deposition of molybdenum oxides thin films on copper and platinum. J. Alloys Compd. 2009, 483, 402–405. [Google Scholar] [CrossRef]
  35. Soundeswaran, S.; Kumar, O.S.; Babu, S.M.; Ramasamy, P.; Dhanasekaran, R. Influence of ultrasonification in CdS thin film deposition in PCD technique. Mater. Lett. 2005, 59, 1795–1800. [Google Scholar] [CrossRef]
  36. Podder, J.; Kobayashi, R.; Ichimura, M. Photochemical deposition of CuxS thin films from aqueous solutions. Thin Solid Films 2005, 472, 71–75. [Google Scholar] [CrossRef]
  37. Zhang, W.L.; Zhang, S.; Yang, M.; Chen, T.P. Microstructure of magnetron sputtered amorphous SiOx films: Formation of amorphous Si core-shell nanoclusters. J. Phys. Chem. C 2010, 114, 2414–2420. [Google Scholar] [CrossRef]
  38. Wu, S.; Huang, H.; Shang, M.; Du, C.; Wu, Y.; Song, W. High visible light sensitive MoS2 ultrathin nanosheets for photoelectrochemical biosensing. Biosens. Bioelectron. 2017, 92, 646–653. [Google Scholar] [CrossRef]
  39. Deng, S.; Zhong, Y.; Zeng, Y.; Wang, Y.; Yao, Z.; Yang, F.; Lin, S.; Wang, X.; Lu, X.; Xia, X.; et al. Directional construction of vertical nitrogen-doped 1T-2H MoSe2/graphene shell/core nanoflake arrays for efficient hydrogen evolution reaction. Adv. Mater. 2017, 29, 1700748. [Google Scholar] [CrossRef]
  40. Jiang, Q.; Lu, Y.; Huang, Z.; Hu, J. Facile solvent-thermal synthesis of ultrathin MoSe2 nanosheets for hydrogen evolution and organic dyes adsorption. Appl. Surf. Sci. 2017, 402, 277–285. [Google Scholar] [CrossRef]
  41. Nam, D.; Lee, J.U.; Cheong, H. Excitation energy dependent Raman spectrum of MoSe2. Sci. Rep. 2015, 5, 17113. [Google Scholar] [CrossRef]
  42. Gu, J.I.; Lee, J.; Rhee, C.K.; Sohn, Y. Enhanced electrochemical hydrogen evolution over defect-induced hybrid MoO3/Mo3O9·H2O microrods. Appl. Surf. Sci. 2019, 469, 348–356. [Google Scholar] [CrossRef]
  43. Xu, X.B.; Sun, Y.; Qiao, W.; Zhang, X.; Chen, X.; Song, X.Y.; Du, Y.W. 3D MoS2-graphene hybrid aerogels as catalyst for enhanced efficient hydrogen evolution. Appl. Surf. Sci. 2017, 396, 1520–1527. [Google Scholar] [CrossRef]
  44. Jian, W.; Cheng, X.; Huang, Y.; You, Y.; Zhou, R.; Sun, T.; Xu, J. Arrays of ZnO/MoS2 nanocables and MoS2 nanotubes with phase engineering for bifunctional photoelectrochemical and electrochemical water splitting. Chem. Eng. J. 2017, 328, 474–483. [Google Scholar] [CrossRef]
  45. Zhang, J.; Wu, M.H.; Shi, Z.T.; Jiang, M.; Jian, W.J.; Xiao, Z.; Li, J.; Lee, C.S.; Xu, J. Composition and interface engineering of alloyed MoS2xSe2(1–x) nanotubes for enhanced hydrogen evolution reaction activity. Small 2016, 12, 4379–4385. [Google Scholar] [CrossRef]
  46. Upadhyay, K.K.; Nguyen, T.; Silva, T.M.; Carmezim, M.J.; Montemor, M.F. Electrodeposited MoOx films as negative electrode materials for redox supercapacitors. Electrochim. Acta 2017, 225, 19–28. [Google Scholar] [CrossRef]
  47. Liu, C.; Xie, Z.; Wang, W.; Li, Z.; Zhang, Z. Fabrication of MoOx film as a conductive anode material for micro-supercapacitors by electrodeposition and annealing. J. Electrochem. Soc. 2014, 161, A1051–A1057. [Google Scholar] [CrossRef]
  48. Guo, L.; Oskam, G.; Radisic, A.; Hoffmann, P.M.; Searson, P.C. Island growth in electrodeposition. J. Phys. D 2011, 44, 443001. [Google Scholar] [CrossRef]
  49. Shu, H.; Zhou, D.; Li, F.; Cao, D.; Chen, X. Defect engineering in MoSe2 for the hydrogen evolution reaction: From point defects to edges. ACS Appl. Mater. Interfaces 2017, 9, 42688–42698. [Google Scholar] [CrossRef]
  50. Ravikumar, C.H.; Nair, G.V.; Muralikrishna, S.; Nagaraju, D.H.; Balakrishna, R.G. Nanoflower like structures of MoSe2 and MoS2 as efficient catalysts for hydrogen evolution. Mater. Lett. 2018, 220, 133–135. [Google Scholar] [CrossRef]
  51. Sekizawa, K.; Oh-Ishi, K.; Kataoka, K.; Arai, T.; Suzuki, T.M.; Morikawa, T. Stoichiometric water splitting using p-type Fe2O3 based photocathode with the aid of multi-heterojunction. J. Mater. Chem. A 2017, 5, 6483–6493. [Google Scholar] [CrossRef]
  52. Briggs, D.; Seah, M.P. Practical Surface Analysis, 2nd ed.; Wiley and Sons: Hoboken, NJ, USA, 1990; Volume 1. [Google Scholar]
  53. Tosini, G.; Ferguson, I.; Tsubota, K. Effects of blue light on the circadian system and eye physiology. Mol. Vis. 2016, 22, 61–72. [Google Scholar] [PubMed]
Figure 1. Scanning electron microscope (SEM) images of MoOx/Si (A and A1), MoS2/Si (B and B1) and MoSe2/Si (C and C1) samples.
Figure 1. Scanning electron microscope (SEM) images of MoOx/Si (A and A1), MoS2/Si (B and B1) and MoSe2/Si (C and C1) samples.
Catalysts 09 00494 g001
Figure 2. Raman spectra of (A) MoOx/Si, (B) MoS2/Si and (C) MoSe2/Si samples.
Figure 2. Raman spectra of (A) MoOx/Si, (B) MoS2/Si and (C) MoSe2/Si samples.
Catalysts 09 00494 g002
Figure 3. Mo 3d, O 1s, Se 3d, S 2p and valence band (VB) X-ray photoelectron spectra (XPS) for MoOx/Si, MoS2/Si and MoSe2/Si samples.
Figure 3. Mo 3d, O 1s, Se 3d, S 2p and valence band (VB) X-ray photoelectron spectra (XPS) for MoOx/Si, MoS2/Si and MoSe2/Si samples.
Catalysts 09 00494 g003
Figure 4. Linear sweep voltammetry curves (voltage range: +0.2~1.0 V) at a scan rate of 10 mV/sec) under visible light exposure for (a) MoOx/Si, (b) MoS2/Si and (c) MoSe2/Si samples. Inset LSV curves were taken under the dark condition. The corresponding light ON-and-OFF LSV curves are displayed in the inset. Inset photo shows the bubble formed on the catalyst surface during LSV.
Figure 4. Linear sweep voltammetry curves (voltage range: +0.2~1.0 V) at a scan rate of 10 mV/sec) under visible light exposure for (a) MoOx/Si, (b) MoS2/Si and (c) MoSe2/Si samples. Inset LSV curves were taken under the dark condition. The corresponding light ON-and-OFF LSV curves are displayed in the inset. Inset photo shows the bubble formed on the catalyst surface during LSV.
Catalysts 09 00494 g004
Figure 5. Linear sweep voltammetry curves (voltage range: +0.2~1.0 V) in N2- (A) and CO2-purged (B) 0.1 M NaHCO3 electrolyte at a scan rate of 10 mV/sec under dark (in the corresponding inset Figure) and the visible light exposure condition for (a) MoOx/Si, (b) MoS2/Si and (c) MoSe2/Si samples. The inset (B) shows the light ON-and-OFF LSV curves.
Figure 5. Linear sweep voltammetry curves (voltage range: +0.2~1.0 V) in N2- (A) and CO2-purged (B) 0.1 M NaHCO3 electrolyte at a scan rate of 10 mV/sec under dark (in the corresponding inset Figure) and the visible light exposure condition for (a) MoOx/Si, (b) MoS2/Si and (c) MoSe2/Si samples. The inset (B) shows the light ON-and-OFF LSV curves.
Catalysts 09 00494 g005
Figure 6. Enhancement factors for HER CDlight ON/ HER CDlight OFF, CDCO2 bubbling/CDN2 bubbling under light ON condition and CDCO2 bubbling/CDN2 bubbling under the light OFF condition for (a) MoOx/Si, (b) MoS2/Si and (c) MoSe2/Si samples.
Figure 6. Enhancement factors for HER CDlight ON/ HER CDlight OFF, CDCO2 bubbling/CDN2 bubbling under light ON condition and CDCO2 bubbling/CDN2 bubbling under the light OFF condition for (a) MoOx/Si, (b) MoS2/Si and (c) MoSe2/Si samples.
Catalysts 09 00494 g006

Share and Cite

MDPI and ACS Style

Hong, S.; Rhee, C.K.; Sohn, Y. Photoelectrochemical Hydrogen Evolution and CO2 Reduction over MoS2/Si and MoSe2/Si Nanostructures by Combined Photoelectrochemical Deposition and Rapid-Thermal Annealing Process. Catalysts 2019, 9, 494. https://doi.org/10.3390/catal9060494

AMA Style

Hong S, Rhee CK, Sohn Y. Photoelectrochemical Hydrogen Evolution and CO2 Reduction over MoS2/Si and MoSe2/Si Nanostructures by Combined Photoelectrochemical Deposition and Rapid-Thermal Annealing Process. Catalysts. 2019; 9(6):494. https://doi.org/10.3390/catal9060494

Chicago/Turabian Style

Hong, Sungmin, Choong Kyun Rhee, and Youngku Sohn. 2019. "Photoelectrochemical Hydrogen Evolution and CO2 Reduction over MoS2/Si and MoSe2/Si Nanostructures by Combined Photoelectrochemical Deposition and Rapid-Thermal Annealing Process" Catalysts 9, no. 6: 494. https://doi.org/10.3390/catal9060494

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop