Next Article in Journal
Suzuki–Miyaura Coupling Using Monolithic Pd Reactors and Scaling-Up by Series Connection of the Reactors
Previous Article in Journal
Synthesis and Evaluation of Copper-Supported Titanium Oxide Nanotubes as Electrocatalyst for the Electrochemical Reduction of Carbon Oxide to Organics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

One-Step Synthesis of MoS2/TiSi2 via an In Situ Photo-Assisted Reduction Method for Enhanced Photocatalytic H2 Evolution under Simulated Sunlight Illumination

1
College of Chemistry, Chemical Engineering and Materials Science, Soochow University, Suzhou 215123, China
2
Jiangsu Key Laboratory of Precious Metal Chemistry and Technology, Jiangsu University of Technology, Changzhou 213001, China
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(3), 299; https://doi.org/10.3390/catal9030299
Submission received: 11 March 2019 / Revised: 20 March 2019 / Accepted: 20 March 2019 / Published: 25 March 2019

Abstract

:
A new MoS2/TiSi2 complex catalyst was designed and synthesized by a simple one-step in situ photo-assisted reduction procedure. The structural and morphological properties of the composites were characterized by X-ray diffraction (XRD), transmission electron microscopy (TEM), scanning electron microscopy (SEM), and ultraviolet-visible diffused reflectance spectroscopy (UV-vis DRS), which proved the formation of MoS2/TiSi2. MoS2/TiSi2 with optimized composition showed obviously enhanced photocatalytic activity and superior durability for water reduction to produce H2. The H2 generation rate over the MoS2/TiSi2 photocatalyst containing 3 wt % MoS2 reached 214.1 μmol·h−1·g−1 under visible light irradiation, which was ca. 5.6 times that of the pristine TiSi2. The improved photocatalytic activity of MoS2/TiSi2 could be related to the broad response spectrum, large visible light absorption, and synergies among MoS2 and TiSi2 that enhance photoexcited charge transfer and separation.

1. Introduction

In recent decades, semiconductor photocatalysis has received close attention owing to its potential application in the production of renewable hydrogen [1,2]. Since the discovery of photoelectrocatalytic H2 production in the TiO2 electrode by Fujishima and Honda in 1972 [3], the semiconductor photocatalyst (Eg ≈ 3.2 eV) has been investigated extensively [4,5]. However, due to its inherent shortcomings, such as its broad bandgap and low quantum efficiency [6], researchers have been making great efforts to modify TiO2 or seek novel semiconductor photocatalysts to improve the photocatalytic activity and efficiency [7].
A novel compound material, titanium disilicide (TiSi2) has been attracting interest because of its thermal stability, good electrical conductivities, single electron tunneling characteristics, and eminent light absorption ability from near-UV to visible light (e.g., ~3.4–1.5 eV) [8,9]. Ritterskamp et al. first used TiSi2 as a photocatalyst for water splitting [10]. After that, a series of TiSi2-based compounds or composites, including TiSi2–SiC [11], Ti5Si3 [12], TiO2/TiSi2 [7,8], were synthesized. RuO2/TiSi2/RGO hybrid and WS2/TiSi2 composite were also successfully synthesized by our group [1,13]. These novel materials showed improved photoelectrochemical properties and enhanced photocatalytic activity.
In recent years, the indirect-gap semiconductor MoS2 (Eg ≈ 1.8 eV) has found many potential uses in the fields of microelectronics, lithium batteries, H2 storage, and catalysis for hydrodesulphurization. More importantly, composites formed by modifying the primary catalysts with MoS2, such as MoS2/CdS [14], MoS2/grapheme [15], and MoS2/zinc cadmium sulfide [16], demonstrated improved activity for photocatalytic water splitting to produce hydrogen.
The conduction band (CB) of TiSi2 (−0.43 eV) [10] is higher than that of MoS2 (−0.1 eV) [17], and the valence band (VB) of TiSi2 (1.07 eV) [10] is lower than that of MoS2 (1.7 eV) [17]. When TiSi2 and MoS2 are combined to form a composite photocatalyst, the electrons on the conduction band of TiSi2 can be transferred to the conduction band of MoS2, and the holes on the valence band of MoS2 can be transferred to the valence band of TiSi2. The combination of MoS2 used as co-catalysts and TiSi2 used as main catalysts is probably beneficial to the catalytic performance of hydrogen production. According to our own survey of the literature, MoS2-functionalized TiSi2 materials have not been reported. Therefore we cover a novel hybrid consisted of MoS2 and TiSi2 by one-step in situ photo-reduction and its photocatalytic application for the photocatalytic H2 evolution under visible light.

2. Results and Discussion

2.1. Morphology and Structure

The XRD patterns of MoS2, TiSi2, and MoS2-modified TiSi2 composite (MoS2/TiSi2-3) are shown in Figure 1. The XRD pattern of the measured MoS2 sample demonstrates peaks located at 13.8°, 32.0°, 39.0°, 57.7°, and 59.6°, which can be indexed to the (002), (100), (103), (110), and (003) hexagonal crystallographic faces of molybdenum disulfide (JCPDS No. 37-1492), respectively. The XRD pattern of TiSi2 displays distinct diffraction peaks at 23.3°, 38.6°, 41.7°, 42.6°, and 49.3°, corresponding to the (111), (311), (040), (022), and (331) planes of face-centered orthorhombic structure of TiSi2 (JCPDS No. 35-0785). From the XRD pattern of MoS2/TiSi2, the characteristic peaks of TiSi2 and MoS2 with similar position can be detected after the reaction process, indicating that MoS2 has been synthesized by photo-reduction.
The morphology of MoS2/TiSi2 was determined by SEM and TEM and demonstrated in Figure 2 and Figure S1. TiSi2 consists of grains of irregular shape with flat surfaces and sharp edges. The size of the grains is in the range of 1–10 μm. MoS2 particles of about 400 nm are well dispersed at the surface of TiSi2. The EDS results of the MoS2/TiSi2 sample (Figure S2) reveal that the composite consists of Mo, S, Ti, and Si, further proving that MoS2 has been successfully deposited at TiSi2.
XPS was carried out and the corresponding results are shown in Figure 3 and Figure S3. Figure 3A shows the high-resolution Ti 2p XPS spectra of the samples. The peaks located at 453.5, 458.9, and 464.8 eV are attributed to Ti0 p3/2, Ti4+ p3/2, and Ti4+ p1/2 [10,18], respectively. The peaks located at 98.6 and 102.4 eV in the pattern of TiSi2 (Figure 3B) were attributed to Si0 p3/2 and Si4+ p3/2 [18,19]. Comparing the above peak positions with those of MoS2/TiSi2, the binding energy of Ti4+ and Si4+ in MoS2/TiSi2-3 is higher than in TiSi2, while the binding energy of Ti0 and Si0 in MoS2/TiSi2-3 is lower than in TiSi2. The peak located at 163.2 eV confirms the presence of S2− in MoS2 (Figure 3C). The blue shift of the S 2p peak position of MoS2/TiSi2-3 can be clearly observed. The changes in binding energy of S were possibly due to the S‒Ti and S‒Si bond formation. The shift of the relative peak position is attributed to the interaction of TiSi2 with MoS2 [1,18,20]. Figure 3D demonstrates the XPS spectra of Mo 3d for MoS2 and MoS2/TiSi2-3. The Mo 3d peak positions of MoS2/TiSi2-3 also shift to higher binding energies, indicating a low load of MoS2 and the remarkable combination effect of MoS2 and TiSi2 [14,21].

2.2. Optical and Photoelectrochemical Properties

Figure 4 exhibits the UV-vis absorption spectra of the obtained samples. The absorption of pure TiSi2 is in the range of ca. 400 up to 850 nm, where its absorption is higher than that of MoS2. The MoS2/TiSi2-3 sample demonstrates higher absorption intensity compared with MoS2 and TiSi2, indicating that the MoS2/TiSi2 composite possesses enhanced ability for harvesting visible light.
The photocurrent responses for TiSi2, MoS2, and MoS2/TiSi2-3 are displayed in Figure 5A. The photocurrent density of TiSi2 or MoS2 is rather low (ca. 0.22 μA·cm−2). The photocurrent density of MoS2/TiSi2-3 (0.59 μA·cm−2) is more than 2.6 times that of TiSi2 in the same irradiation conditions. The results suggest the positive synergetic effect between MoS2 and TiSi2, which leads to enhanced photoinduced chargers transfer and separation. Figure 5B shows the electrochemical impedance spectra (EIS) of the samples are shown in. The diameter of the semicircle of MoS2/TiSi2-3 plot is obviously smaller than that of MoS2 and TiSi2, which indicates effectively enhanced carrier transfer at the interface between the MoS2/TiSi2-3 electrode and the electrolyte [22]. The experimental data for all the electrodes can be expressed as an equivalent circuit, displayed as shown as the inset of Figure 5B, in which CPE1 is constant phase elements connected in parallel with R2 [23,24], R2 is the resistance of solution, and the ohmic series resistance (R1) is the resistance of charge-transfer resistance at interfaces [25,26].

2.3. Photocatalytic Activity

The photocatalytic activity for H2 production of the samples are shown in Figure 6 and Figure S4. MoS2 alone shows little photocatalytic activity. The hydrogen production rate for TiSi2 is 38.4 μmol·h−1·g−1. The hydrogen production rate for the MoS2/TiSi2-1 sample increases to 119.6 μmol·h−1·g−1. The photocatalytic activity of the MoS2/TiSi2 augments with the increasing of the quantity of MoS2 in MoS2/TiSi2. The hydrogen production rate of the MoS2/TiSi2-3 sample reaches the maximum value 214.1 μmol·h−1·g−1, which is approximately 5.6 times that of TiSi2. The fact that composite catalysts exhibit much enhanced activity for H2 evolution can be attributed to the synergistic effect between TiSi2 andMoS2. Through the heterojunction formed between MoS2 and TiSi2, the photogenerated electrons in TiSi2 easily transfer to the conduction band of MoS2, which inhibits the recombination of e-h+, thus improving the photocatalytic activity.
The results of the stability are shown in Figure 6. In the total 25 h recycle tests, the photocatalytic activity for H2 production of MoS2/TiSi2-3 was almost invariable and the average value was 214 μmol·h−1·g−1, indicating that the composite prepared by the in situ photo-assisted reduction of MoS2 on TiSi2 possesses good photocatalytic stability.
Because of the wide-bandgap energy (3.4~1.5 eV) for TiSi2, TiSi2 can easily absorb the photons under visible light irradiation to generate plenty of electrons and holes in its conduction and valence band, respectively [8]. Well-combined MoS2/TiSi2 heterojunctions can prompt the transfer of photogenerated chargers between MoS2 and TiSi2, and provide a pathway for charges transfer simultaneously (Scheme S1). Since the CB of TiSi2 (−0.43 eV) is more negative than that of MoS2 (−0.1 eV), the photogenerated electrons in the CB of TiSi2 were easy to transfer through the heterojunction interface between MoS2 and TiSi2 to the CB of MoS2 particles deposited on the surface of TiSi2. Simultaneously, the holes transfer from the higher VB of MoS2 (1.7 eV) to the VB of TiSi2 (1.07 eV). The shorter charge transfer route effectively restrains the recombination process of the electron‒hole pairs. H+ is reduced to hydrogen atom, to form hydrogen by the electrons in the CB of MoS2, while the holes on the TiSi2 surface are rapidly scavenged by H2O and OH-, generating ·OH to oxidize sacrificial agents.

2.4. Formation Mechanism of the MoS2/TiSi2 Photocatalyst

Scheme 1 shows the proposed formation mechanism of the MoS2/TiSi2 photocatalyst. The electrons are generated at the VB of TiSi2 are excited by visible-light, which in turn reduce [MoS4]2− to MoS2 [16], as shown in Equation (1):
[MoS4]2− + 2e → MoS2 + 2S2−.
The holes (h) of TiSi2 are scavenged by lactic acid. MoS2 particles supported on the surface of TiSi2 not only form a heterojunction with TiSi2 but also provide effective active sites to improve hydrogen production.

3. Experimental

3.1. Materials

Titanium disilicide was obtained from Alfa Aesar and all other chemicals (analytical purity) were obtained from J&K Scientific Limited (Beijing, China). All of the reagents and chemicals were utilized as received without further purification.

3.2. Synthesis

Fifty milliliters of 10 vol % lactic acid (LA) solution was added into to a 60 mL three-necked round-bottom flask with a quartz window, 50 mg of TiSi2 and a certain amount of (NH4)2MoS4 were dispersed in the lactic acid solution. The mixed solution was ventilated via bubbling argon with 30 min, then the mixture was irradiated by 150 W Xe lamp through the quartz window with a cutoff filter at 420 nm. After 60 min irradiation, the suspension was centrifuged and the solid was washed by ethanol first and then through deionized water. The obtained solid was dried under vacuum under 50 °C overnight, obtaining MoS2 modified TiSi2(MoS2/TiSi2-x), where x represents the mass percentage of MoS2 in the MoS2/TiSi2 composite. For comparison, MoS2 was obtained in the same process without adding TiSi2.

3.3. Characterization

X-Ray diffraction (XRD) measurement was performed under a Philips diffractometer (X’Pert-Pro MRD, Amsterdam, The Netherland) with a Ni-filtered Cu Kα source (λ = 0.15418 nm) in the 2θ scanning range from 10° to 90°. Scanning electron microscopy (SEM) was taken on SEM Hitachi S-4800 (Hitachi High-Tech, Tokyo, Japan). The energy-dispersive X-ray (EDX) was carried out in a KEVEX X-ray energy detector (KEVEX, Newark, NJ, USA). The transmission electron microscopy (TEM) was carried out by TECNAI-G2 electron microscope (FEI, Hillsboro, OR, USA) using 200 kV accelerating voltage. UV-vis absorption spectra (DRS) were performed on a Hitachi UV-3010 spectrophotometer (Hitachi High-Tech, Tokyo, Japan) with BaSO4 as white standard. X-ray photoelectron spectroscopy (XPS) was determined on an AXIS Ultra DLD system (Kratos Analytical Inc., Manchester, UK) used a monochromatic Al Kα radiation and C 1s peak (285.5 eV) was as a reference to calibrate all of the XPS spectra using XPS Peak software (Version 4.1, Raymund W.M. Kwok, Hongkong, China).

3.4. Photoelectrochemical Measurements

All photoelectrochemical measurements were carried out in a three-electrode system connected with a CHI660D (CH Instruments Inc., Shanghai, China) electrochemical workstation. An indium tin oxide (ITO) glass covered by the sample was employed as the working electrode, a saturated calomel electrode (SCE) as the reference electrode and a platinum wire as the counter electrode. The working electrode was obtained first through dispersing 0.2 mg of the sample in 1 mL of a solution composed of 0.4 mL of ethanol, 0.4 mL of ethylene glycol and 0.2 mL of chloroform, after grinding and sonication, the slurry was then dropped onto a clean ITO glass and dried in vacuum at 45 °C. The sample area on the ITO glass was ca. 1.0 cm2. During the photoelectrochemical measurements, the electrodes were immersed in 50 mL of 0.5 M Na2SO4 solution containing 5 mL lactic acid and the ITO glass with catalyst was irradiated by a GY-10 xenon lamp (150 W) (TIAN JIN TUO PU, Tianjin, China). Electrochemical impedance spectroscopy (EIS) was recorded under the frequency range 1–105 Hz with an AC perturbation signal of 5 mV.

3.5. Photocatalytic Reaction for Hydrogen Evolution

20 mg of as-prepared photocatalyst was added in 50 mL of 10 vol % lactic acid (LA) solution. The suspension was added into a 60 mL three-necked flask with quartz window. The area of the effective optical channel is ca. 3 cm2. The reaction mixture was vigorously stirred and degassed via Ar with 30 min. Then the mixture was irradiated by 150 W Xe lamp through the quartz window with a cut-off filter at 420 nm. The lamp was positioned ca. 10 cm away from the optical entry window of the reactor. The distance between the lamp and the quartz window was ca. 10 cm. The gas chromatograph GC1650 (Ke Xiao Instruments Co., Ltd., Hangzhou, China) with a thermal conductivity detector (molecular sieve 5 A column, Ar carrier) was used to detected the amount of hydrogen production.

4. Conclusions

A robust and effective MoS2/TiSi2 photocatalyst has been successfully fabricated through the convenient in situ photo-reduction method. The characterization results of prepared catalysts reveal that the MoS2 was evenly distributed at TiSi2 forming heterojunctions benefiting photoexcited charges transfer and separation. The hydrogen production rate by the optimized MoS2/TiSi2 catalyst reaches 214.1 μmol·h−1·g−1 by visible light illumination, which is significantly superior to that of TiSi2 and MoS2. The characteristics possessed by MoS2/TiSi2, such as broad spectral response, enhanced absorption capability and the synergistic effect between MoS2 and TiSi2, are owed to high photocatalytic performance of the catalyst. This work demonstrates the feasibility of increasing H2 evolution activity of TiSi2-based catalysts by combining TiSi2 with MoS2.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/9/3/299/s1, Figure S1: Morphology characterization of (A) TiSi2 and (B) MoS2 using SEM. Figure S2: EDX results of MoS2/TiSi2-3. Figure S3: A full-scan XPS of TiSi2, MoS2 and MoS2/TiSi2-3. Figure S4: The amount of H2 evolved over the samples. Scheme S1: Schematic illustration of the catalytic mechanism.

Author Contributions

C.Z., Y.D. and P.Y. conceived and designed the experiments; C.Z., A.L. and K.L. performed the experiments; C.Z. and P.Y. analyzed the data and wrote the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (21373143).

Conflicts of Interest

None of the authors has any competing interests.

References

  1. Chu, D.; Zhang, C.; Yang, P.; Du, Y.; Lu, C. WS2 as an Effective Noble-Metal Free Cocatalyst Modified TiSi2 for Enhanced Photocatalytic Hydrogen Evolution under Visible Light Irradiation. Catalysts 2016, 6, 136. [Google Scholar] [CrossRef]
  2. Boyjoo, Y.; Sun, H.; Liu, J.; Pareek, V.K.; Wang, S. A review on photocatalysis for air treatment: From catalyst development to reactor design. Chem. Eng. J. 2017, 310, 537–559. [Google Scholar] [CrossRef]
  3. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  4. Zhao, H.; Wu, M.; Liu, J.; Deng, Z.; Li, Y.; Su, B.-L. Synergistic promotion of solar-driven H2 generation by three-dimensionally ordered macroporous structured TiO2-Au-CdS ternary photocatalyst. Appl. Catal. B Environ. 2016, 184, 182–190. [Google Scholar] [CrossRef]
  5. Wen, J.; Li, X.; Li, H.; Ma, S.; He, K.; Xu, Y.; Fang, Y.; Liu, W.; Gao, Q. Enhanced visible-light H2 evolution of g-C3N4 photocatalysts via the synergetic effect of amorphous NiS and cheap metal-free carbon black nanoparticles as co-catalysts. Appl. Surf. Sci. 2015, 358 Pt A, 204–212. [Google Scholar] [CrossRef]
  6. Rajamanickam, D.; Dhatshanamurthi, P.; Shanthi, M. Enhanced photocatalytic efficiency of NiS/TiO2 composite catalysts using sunset yellow, an azo dye under day light illumination. Mater. Res. Bull. 2015, 61, 439–447. [Google Scholar] [CrossRef]
  7. Lin, Y.; Zhou, S.; Liu, X.; Sheehan, S.; Wang, D. TiO2/TiSi2 Heterostructures for High-Efficiency Photoelectrochemical H2O Splitting. J. Am. Chem. Soc. 2009, 131, 2772–2773. [Google Scholar] [CrossRef] [PubMed]
  8. Banerjee, S.; Mohapatra, S.K.; Misra, M. Water Photooxidation by TiSi2–TiO2 Nanotubes. J. Phys. Chem. C 2011, 115, 12643–12649. [Google Scholar] [CrossRef]
  9. Pelleg, J.; Sade, G. TiB2/TiSi2 bilayer fabrication by various techniques: Structure and contact properties. Phys. B Condens. Matter 2006, 371, 20–34. [Google Scholar] [CrossRef]
  10. Ritterskamp, P.; Kuklya, A.; Stkamp, M.-A.; Kerpen, K.; Weidenthaler, C.; Demuth, M. A Titanium Disilicide Derived Semiconducting Catalyst for Water Splitting under Solar Radiation—Reversible Storage of Oxygen and Hydrogen. Angew. Chem. Int. Ed. 2007, 46, 7770–7774. [Google Scholar] [CrossRef]
  11. Shon, I.-J.; Park, H.-K.; Kim, H.-C.; Yoon, J.-K.; Hong, K.-T.; Ko, I.-Y. One-step synthesis and densification of nanostructured TiSi2–SiC composite from mechanically activated (TiC + 3Si) powders by high-frequency-induced heated combustion. Scr. Mater. 2007, 56, 665–668. [Google Scholar] [CrossRef]
  12. Liu, J.; Bai, Y.; Chen, P.; Cui, N.; Yin, H. Reaction synthesis of TiSi2 and Ti5Si3 by ball-milling and shock loading and their photocatalytic activities. J. Alloys Compd. 2013, 555, 375–380. [Google Scholar] [CrossRef]
  13. Mou, Z.; Yin, S.; Zhu, M.; Du, Y.; Wang, X.; Yang, P.; Zheng, J.; Lu, C. RuO2/TiSi2/graphene composite for enhanced photocatalytic hydrogen generation under visible light irradiation. Phys. Chem. Chem. Phys. 2013, 15, 2793–2799. [Google Scholar] [CrossRef]
  14. Liu, Y.; Yu, Y.-X.; Zhang, W.-D. MoS2/CdS Heterojunction with High Photoelectrochemical Activity for H2 Evolution under Visible Light: The Role of MoS2. J. Phys. Chem. C 2013, 117, 12949–12957. [Google Scholar] [CrossRef]
  15. Chang, K.; Chen, W. In situ synthesis of MoS2/graphene nanosheet composites with extraordinarily high electrochemical performance for lithium ion batteries. Chem. Commun. 2011, 47, 4252–4254. [Google Scholar] [CrossRef]
  16. Nguyen, M.; Tran, P.D.; Pramana, S.S.; Lee, R.L.; Batabyal, S.K.; Mathews, N.; Wong, L.H.; Graetzel, M. In situ photo-assisted deposition of MoS2 electrocatalyst onto zinc cadmium sulphide nanoparticle surfaces to construct an efficient photocatalyst for hydrogen generation. Nanoscale 2013, 5, 1479–1482. [Google Scholar] [CrossRef]
  17. Lu, X.; Jin, Y.; Zhang, X.; Xu, G.; Wang, D.; Lv, J.; Zheng, Z.; Wu, Y. Controllable synthesis of graphitic C3N4/ultrathin MoS2 nanosheet hybrid nanostructures with enhanced photocatalytic performance. Dalton Trans. 2016, 45, 15406–15414. [Google Scholar] [CrossRef]
  18. Larciprete, R.; Danailov, M.; Barinov, A.; Gregoratti, L.; Kiskinova, M. Thermal and pulsed laser induced surface reactions in Ti/Si(001) interfaces studied by spectromicroscopy with synchrotron radiation. J. Appl. Phys. 2001, 90, 4361–4369. [Google Scholar] [CrossRef]
  19. Fouad, O.A.; Yamazato, M.; Era, M.; Nagano, M.; Hirai, T.; Usui, I. Preparation and properties of TiSi2 thin films from TiCl4/H2 by plasma enhanced chemical vapor deposition. J. Cryst. Growth 2002, 234, 440–446. [Google Scholar] [CrossRef]
  20. Wee, A.T.S.; Huan, A.C.H.; Osipowicz, T.; Lee, K.K.; Thian, W.H.; Tan, K.L.; Hogan, R. Surface and interface studies of titanium silicide formation. Thin Solid Films 1996, 283, 130–134. [Google Scholar] [CrossRef]
  21. Zong, X.; Yan, H.; Wu, G.; Ma, G.; Wen, F.; Wang, L.; Li, C. Enhancement of Photocatalytic H2 Evolution on CdS by Loading MoS2 as Cocatalyst under Visible Light Irradiation. J. Am. Chem. Soc. 2008, 130, 7176–7177. [Google Scholar] [CrossRef]
  22. Yue, Z.; Chu, D.; Huang, H.; Huang, J.; Yang, P.; Du, Y.; Zhu, M.; Lu, C. A novel heterogeneous hybrid by incorporation of Nb2O5 microspheres and reduced graphene oxide for photocatalytic H2 evolution under visible light irradiation. RSC Adv. 2015, 5, 47117–47124. [Google Scholar] [CrossRef]
  23. Franceschini, E.A.; Gomez, M.J.; Lacconi, G.I. One step synthesis of high efficiency nickel/mesoporous TiO2 hybrid catalyst for hydrogen evolution reaction. J. Energy Chem. 2019, 29, 79–87. [Google Scholar] [CrossRef]
  24. Bahrami, A.; Kazeminezhad, I.; Abdi, Y. Pt-Ni/rGO counter electrode: Electrocatalytic activity for dye-sensitized solar cell. Superlattices Microstruct. 2019, 125, 125–137. [Google Scholar] [CrossRef]
  25. Li, Y.; Zhang, H.-X.; Liu, F.-T.; Dong, X.-F.; Li, X.; Wang, C.-W. New design of oriented NiS nanoflower arrays as platinum-free counter electrode for high-efficient dye-sensitized solar cells. Superlattices Microstruct. 2019, 125, 66–71. [Google Scholar] [CrossRef]
  26. Vishnu, N.; Badhulika, S. Single step grown MoS2 on pencil graphite as an electrochemical sensor for guanine and adenine: A novel and low cost electrode for DNA studies. Biosens. Bioelectron. 2019, 124, 122–128. [Google Scholar] [CrossRef] [PubMed]
Figure 1. XRD patterns of the MoS2 (a), TiSi2 (b), and MoS2/TiSi2-3 (c).
Figure 1. XRD patterns of the MoS2 (a), TiSi2 (b), and MoS2/TiSi2-3 (c).
Catalysts 09 00299 g001
Figure 2. SEM (A) and TEM (B) images of MoS2/TiSi2-3.
Figure 2. SEM (A) and TEM (B) images of MoS2/TiSi2-3.
Catalysts 09 00299 g002
Figure 3. XPS of (A) Ti 2p for TiSi2 and MoS2/TiSi2-3, (B) Si 2p for TiSi2 and MoS2/TiSi2-3, (C) S 2p for MoS2 and MoS2/TiSi2-3, (D) Mo 3d for MoS2 and MoS2/TiSi2-3.
Figure 3. XPS of (A) Ti 2p for TiSi2 and MoS2/TiSi2-3, (B) Si 2p for TiSi2 and MoS2/TiSi2-3, (C) S 2p for MoS2 and MoS2/TiSi2-3, (D) Mo 3d for MoS2 and MoS2/TiSi2-3.
Catalysts 09 00299 g003
Figure 4. The UV–vis absorption spectra of the obtained samples.
Figure 4. The UV–vis absorption spectra of the obtained samples.
Catalysts 09 00299 g004
Figure 5. (A) The photocurrent responses of samples under 150 W xenon lamp. The electrolyte was 60 mL of 0.5 M Na2SO4 solution containing 6 mL lactic acid; (B) The electrochemical impedance spectra (EIS) for the samples in 60 mL of 0.5 M Na2SO4 solution containing 6 mL lactic acid.
Figure 5. (A) The photocurrent responses of samples under 150 W xenon lamp. The electrolyte was 60 mL of 0.5 M Na2SO4 solution containing 6 mL lactic acid; (B) The electrochemical impedance spectra (EIS) for the samples in 60 mL of 0.5 M Na2SO4 solution containing 6 mL lactic acid.
Catalysts 09 00299 g005
Figure 6. (A) The amount of H2 evolved over the samples and (B) cycling measurement.
Figure 6. (A) The amount of H2 evolved over the samples and (B) cycling measurement.
Catalysts 09 00299 g006
Scheme 1. Illustration of the preparation of MoS2/TiSi2 via an in situ photo-assisted reduction method.
Scheme 1. Illustration of the preparation of MoS2/TiSi2 via an in situ photo-assisted reduction method.
Catalysts 09 00299 sch001

Share and Cite

MDPI and ACS Style

Zhang, C.; Liu, A.; Li, K.; Du, Y.; Yang, P. One-Step Synthesis of MoS2/TiSi2 via an In Situ Photo-Assisted Reduction Method for Enhanced Photocatalytic H2 Evolution under Simulated Sunlight Illumination. Catalysts 2019, 9, 299. https://doi.org/10.3390/catal9030299

AMA Style

Zhang C, Liu A, Li K, Du Y, Yang P. One-Step Synthesis of MoS2/TiSi2 via an In Situ Photo-Assisted Reduction Method for Enhanced Photocatalytic H2 Evolution under Simulated Sunlight Illumination. Catalysts. 2019; 9(3):299. https://doi.org/10.3390/catal9030299

Chicago/Turabian Style

Zhang, Chunyong, Aijuan Liu, Kezhen Li, Yukou Du, and Ping Yang. 2019. "One-Step Synthesis of MoS2/TiSi2 via an In Situ Photo-Assisted Reduction Method for Enhanced Photocatalytic H2 Evolution under Simulated Sunlight Illumination" Catalysts 9, no. 3: 299. https://doi.org/10.3390/catal9030299

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