Next Article in Journal
Insights into the Pyrolysis Processes of Ce-MOFs for Preparing Highly Active Catalysts of Toluene Combustion
Next Article in Special Issue
Recent Advances in Selective Photo-Epoxidation of Propylene: A Review
Previous Article in Journal
Insights into the TiO2-Based Photocatalytic Systems and Their Mechanisms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improvement of Photocatalytic H2-Generation under Visible Light Irradiation by Controlling the Band Gap of ZnIn2S4 with Cu and In

1
Global Environment Center for Education & Research, Mie University, Mie 514-8507, Japan
2
Department of Chemistry for Materials, Graduate School of Engineering, Mie University, Mie 514-8507, Japan
*
Author to whom correspondence should be addressed.
Submission received: 12 July 2019 / Revised: 29 July 2019 / Accepted: 6 August 2019 / Published: 10 August 2019
(This article belongs to the Special Issue State-of-the-Art Photocatalytical Technology in Asia)

Abstract

:
The band gap controlled photocatalyst (Zn0.74Cu0.13In2S3.805) was prepared via a simple one-step solvothermal method. The effects of doping of Cu+ and excess In on the photocatalytic activity of ZnIn2S4 photocatalyst were investigated. In addition, optical properties, surface morphology and crystal structure were evaluated. The maximum H2 evolution rate (2370 µmol h−1 g−1) was achieved with Zn0.74Cu0.13In2S3.805, which was about five times higher than that of untreated ZnIn2S4 under visible light (λ ≥ 420 nm). The band gap of Zn0.74Cu0.13In2S3.805 decreased to 1.98 eV by raising the maximum position of the valence band, compared to ZnIn2S4. Furthermore, the recombination of electron hole pairs was effectively reduced. This research contributes to the development of highly active photocatalysts under visible light.

1. Introduction

Hydrogen is one of the more demanded synthetic energy carriers. Until now, hydrogen can be produced chemically, thermochemically, biologically, biochemically, biophotochemically, etc. [1,2,3,4]. Among these techniques, hydrogen generation by water splitting using a photocatalyst has been expected as a clean and sustainable energy technology, because it can directly convert solar energy into chemical energy by using only water as a raw material [5,6,7]. It is important to develop highly efficient photocatalysts to replace the current hydrogen generation technology with the photocatalytic water splitting process. Sunlight contains ultraviolet light, visible light and infrared light, and it is ideal to use all the wavelengths when using sunlight. It is known that the wavelength ranges of light, in which the photocatalyst reacts, is dependent on the size of the band gap of the photocatalyst [8,9,10]. Therefore, photocatalytic activity can be exhibited even under longer wavelength light [11,12,13]. The addition of foreign elements, which is one of the ways to change photocatalytic ability, greatly affects the characteristics of the catalyst. Specific requirements to improve the activity of the photocatalytic material generally include efficient light absorption, effective separation of photogenerated charge carriers, and better efficiency of the interface for direct release of hydrogen and/or oxygen from water [14,15,16]. In efficient light absorption, the size of the band gap formed by the conduction band (CB) and the valence band (VB) of the semiconductor photocatalyst is the most important issue [17,18,19]. With regard to the effective separation of photogenerated charge carriers, many elements such as defects in the crystal structure and band structure are involved in a complex manner [20,21,22].
Sulfide photocatalysts are advantageous for visible light driven photocatalysts, because they have narrow band gaps and negative valence bands due to the electron orbit of S, compared to oxide based photocatalysts [23,24,25]. ZnIn2S4 has been tested to have a suitable band gap corresponding to the visible-light absorption region, high photocatalytic activity and considerable chemical stability for photocatalytic H2 evolution [26,27,28]. Yao et al. described the fabrication of Z-scheme PtS–ZnIn2S4/WO3-MnO2 for overall photo-catalytic water splitting [29]. It was reported by Zhao et al. that the combined effects of octahedron NH2-UiO-66 and flowerlike ZnIn2S4 microspheres for photocatalytic hydrogen evolution [30]. However, the absorption wavelength of pure ZnIn2S4 is limited to about 500 nm. Cu species such as Cu+ and Cu2+ affected the valence band of ZnIn2S4 and formed an advantageous band structure for photocatalysts [31,32,33]. In In-Zn-S compounds, the [Zn2+]/[In3+] ratio changes the structural and optoelectronic properties, and greatly affects the composition of In–Zn–S [34]. Thus far, there are few reports that ZnIn2S4 is co-doped with Cu+ and excess indium. In this study, we investigated photocatalytic activity, optical properties and surface morphology of ZnIn2S4 simultaneously doped with Cu+ and excess indium.

2. Results and Discussion

2.1. Structural Characterization

The XRD patterns of ZnIn2S4, Zn0.87Cu0.13In2S3.935, Zn0.87In2S3.87 and Zn0.74Cu0.13In2S3.805 are shown in Figure 1. The XRD pattern of ZnIn2S4 could be indexed as the hexagonal structure (JCPDS No. 65-2023). The three diffraction peaks at around 20.8°, 27.5°, 47.2° and 56.4° could be assigned to the (006), (101), (110) and (202) planes, respectively [35,36]. The XRD patterns of other Zn0.87Cu0.13In2S3.935, Zn0.87In2S3.87 and Zn0.74Cu0.13In2S3.805 photocatalysts also showed similar results. These XRD patterns were in agreement with those reported in previous studies [35], which revealed that Zn0.74Cu0.13In2S3.805 is a hexagonal structure and contains almost no impurities of ZnS. The reason that the Cu and In derived peaks were not observed could be due to the very low doping amount. In addition, with respect to the peak of the (006) plane, a slight shift toward the high angle was observed when doping Cu and increasing In. This means that the interplanar spacing was reduced by doping, suggesting that Cu and excess In may be incorporated into the crystal structure of ZnIn2S4 and exist as a solid solution.
The results of the X-ray photoelectron spectroscopy (XPS) measurement for further structural analysis of photocatalysts are shown in Figure 2. The elemental ratios determined from the XPS spectrum are shown in Table S2. XPS survey spectra of Zn0.87In2S3.87, Zn0.87Cu0.13In2S3.935 and Zn0.74Cu0.13In2S3.805 matched the material ZnIn2S4. This indicates that the impurities were not contained regardless of the doping Cu+ and excess In. Furthermore, from the result of elemental ratio analysis of XPS, it could be confirmed that the elemental ratio of the prepared catalysts were substantially in agreement with the theoretical ratio. Typical narrow spectra of Zn0.74Cu0.13In2S3.805 are shown in Figure 2a. In the XPS spectrum of Zn 2p, the peak of Zn 2p3/2 (1020.4 eV) was observed. This peak was derived from the ZnIn2S4 component in Zn0.74Cu0.13In2S3.805. In the XPS spectrum of In 3d from ZnIn2S4 and In2S3, binding energies of 443.7 (In 3d5/2) and 451.0 eV (In 3d3/2) were observed. The peaks of S 2p were at 161.3 eV (S 2p3/2) and 162.6 eV (S 2p1/2) [37,38]. In addition, since the peak position of Cu 2p3/2 is observed only at 932 eV and Cu 2p3/2 satellite peak derived from Cu2+ was not present at 942 eV, it can be seen that Cu was doped in a monovalent state [39,40]. Both results of XRD and XPS show that the basic structure of Zn0.74Cu0.13In2S3.805 is hexagonal ZnIn2S4, doped with Cu+ and excess In.

2.2. Morphological Analysis

In order to investigate the influence of Cu and excess In doping on the characteristic surface morphology, SEM images of the prepared photocatalysts were observed. The results are shown in Figure 3. In pure ZnIn2S4, a microsphere structure formed by the overlapping of nanosheets was observed [41]. When Cu and excess In was doped, the spherical structure was destroyed. The nanosheet structure also collapsed and aggregated. The shapes of pure ZnIn2S4 and Zn0.74Cu0.13In2S3.805 were very different. Maybe, Cu+ and excess In formed a solid solution with ZnIn2S4.

2.3. Optical Analysis

Figure 4 shows the absorption wavelength region of light using the UV-visible diffuse reflectance spectrum. Furthermore, a Tauc plot calculated from the UV-vis diffuse reflectance spectra (DRS) spectrum is shown in Figure S1, in order to obtain a band gap. As shown in the results of DRS and Tauc plot, the absorption edge of pure ZnIn2S4 was 480 nm, and the size of the band gap was 2.67 eV. When [Zn2+]/[In3+] mole ratio was changed (Zn0.87In2S3.87), the absorption tendencies were almost similar to ZnIn2S4. However, when Cu+ doping (Zn0.87Cu0.13In2S3.935) and excess In doping (Zn0.74Cu0.13In2S3.805) were performed, the absorption edges were extended to about 700 nm. Consequently, the band gap of ZnIn2S4 similarly decreased in the case of the doping Cu and excess In. So as to analyze the band structure, valence band edge measurement for the prepared photocatalysts was performed by XPS. It is clear from the results in Figure S2 that the doping of Cu+ shifted the valence band edge to the negative side. It has been reported that the doping of In3+ forms a sub-band on the positive side of the conduction band of ZnIn2S4 [36]. Therefore, it is reasonable that the doping of Cu+ and In into the photocatalyst reduce the band gap energy of ZnIn2S4.
We investigated the photoluminescence spectra in order to investigate the electron–hole pair separation efficiency, and the results are shown in Figure 5. Luminescence in ultraviolet and visible regions was observed in the photoluminescence spectrum. In general, ultraviolet emission is associated with exciton transition and recombination from the conduction band level to the valence band, and visible light emission is mainly associated with intrinsic or extrinsic defects in the catalyst. The photoluminescence spectra of ZnIn2S4 was approximately close to Zn0.87In2S3.87. On the other hand, the spectra from Zn0.87Cu0.13In2S3.935 and Zn0.74Cu0.13In2S3.805 were lowered by doping of Cu+ and excess In. This may have been due to the decrease of the recombination rate between the photogenerated holes and the electrons photogenerated in the valence band. The photogenerated electrons may have been trapped in the oxygen vacancies generated in the photocatalyst by doping.

2.4. Photocatalytic Activity

Hydrogen generation was conducted using the photocatalyst Zn0.74Cu0.13In2S3.805. ZnIn2S4, Zn0.87In2S3.87 and Zn0.87Cu0.13In2S3.935 were used as comparative objects. All catalysts were loaded with 1wt% Pt as a cocatalyst. The results are shown in Figure 6. The photocatalyst ZnIn2S4 mono-doped with excess In showed slightly higher hydrogen generation activity than that of pure ZnIn2S4. On the other hand, ZnIn2S4 doped only with Cu+ greatly improved the photocatalytic activity. Furthermore, the photocatalyst doped with Cu+ and excess In showed the highest hydrogen generation activity. The maximum H2 evolution rate was 2370 μmol h−1 g−1, which showed about five times better results than that of untreated ZnIn2S4. The reproducibility of the H2 production (relative standard deviation (RSD), for hydrogen amounts) was better than RSD 4% for three repeated measurements.
The TEM images before and after the hydrogen generation of Zn0.74Cu0.13In2S3.805 photocatalyst are shown in Figure S3. Only after hydrogen generation, a 3–4 nm spot was observed on the sample surface. The spot deposition was Pt, which was reduced during the hydrogen production.

2.5. Proposed Hydrogenation Mechanism

The reaction mechanism is shown in Scheme 1. Irradiation of light having a wavelength corresponding to the band gap energy of the Zn0.74Cu0.13In2S3.805 photocatalyst excites electrons in the valence band to the conduction band to produce an electron–hole pair (Formula (1)). From the results of DRS and photoluminescence (PL), it could be considered that Cu+ and excess In doping enhanced charge separation by forming an impurity level on the negative side of the valence band of ZnIn2S4 and the positive side of the conduction band, narrowing the band gap. In Formula (2), a part of electrons excited in the conduction band is consumed for photodeposition of Pt. The deposited Pt reduces H+ by using electrons transferred from the photocatalyst to generate H2 (Formula (3)). On the other hand, sulfite ions and sulfide ions consume the holes for promoting a hydrogen generation reaction. The presence of Na2S is very important in enhancing the photocatalytic activity, as Na2S stabilizes the surface of the metal sulfide by suppressing the formation of surface defects as a scavenger of holes. However, when the concentration of Na2S is high, the pH becomes high. High pH values are thermodynamically disadvantageous in the reaction represented by Formula (4). As described in the Formulas (5)–(7), SO32− and S2− consume holes. According to the Formula (6), S22− ions are generated and act as an optical filter. If S22− is not consumed, it interferes with light absorption. As shown in Formula (8), the reaction between S22− and SO32− forms S2O32−, which is colorless and can hardly affect to light absorption.
Photocatalyst + hν → e + h+
Pt2+ + 2eCB → Pt
2H+ + 2eCB → H2
2H2O + 2eCB → H2 + 2OH
SO32− + H2O + 2h+VB → SO42− + 2H+
2S2− + 2h+VB → S22−
SO32− + S2− + 2h+VB → S2O32−
S22− + SO32− → S2O32− + S2−

3. Materials and Methods

3.1. Preparation of Photocatalysts

All chemicals were analytical grade and used as received without further purification. Zn0.74Cu0.13In2S3.805 was prepared by a simple hydrothermal method. Cetyltrimethylammonium bromide (CTAB, 3.76 mmol) (Wako Pure Chemical Industries, Ltd., Osaka, Japan), stoichiometric mole of ZnSO4·7H2O (Nacalai Tesque, Inc., Kyoto, Japan), InCl3·4H2O and CuCl (I) and an excess of thioacetamide (TAA) (Wako Pure Chemical Industries, Ltd., Osaka, Japan) were dissolved in 50 mL of distilled water. At the same time, in order to keep Cu monovalent, nitrogen was purged into the solution for 10 min to remove dissolved oxygen. The mixed solution was then transferred into a 100 mL Teflon autoclave. The autoclave was sealed, kept at 160 °C for 1 h and cooled to room temperature naturally. After cooling, the product was dried in a vacuum at 40 °C for 4 h and was ground for 30 min. ZnIn2S4 (not doped), Zn0.87Cu0.13In2S3.935 (Cu+ doped) and Zn0.87In2S3.87 (change of Zn2+/In3+) were also prepared by the same method as reference materials. The prepared photocatalysts are shown in F1.

3.2. Characterization of Samples

X-ray powder diffraction (XRD) measurements were performed using a Rigaku RINT Ultima-IV diffractometer. It was carried out by using Cu radiation at a scan rate of 0.04°/s in a scan range of 10°–80°. X-ray photoelectron spectroscopy (XPS) measurements were carried out with a PHI Quantera SXM photoelectron spectrometer using Al Kα radiation. To compensate for surface charge effects, binding energies were calibrated using the C1s peak at 284 eV as the reference. Scanning electron microscope (SEM) observations were performed using a Hitachi S-4000 SEM. The transmission electron microscopy (TEM) images were taken on JEOL product JEM1011. The UV–vis diffuse reflectance spectra (DRS) of the photocatalysts were recorded using a Shimadzu UV-2450 spectrophotometer equipped with an integral sphere assembly, using BaSO4 as a reflectance standard. Photoluminescence (PL) spectra were obtained at an excitation wavelength of 350 nm by using a Shimadzu RF-5300PC spectrofluorophotometer.

3.3. Photocatalytic Hydrogen Generation

A Pyrex column vessel reactor (inner volume: 123 mL) was used for the photocatalytic production of hydrogen from aqueous sulfide solution. In all experiments, 40 mL of solution containing 40 mg of catalyst, 10 mL of 0.04 ppm H2PtCl6 solution and 0.25 M Na2SO3/0.35 M Na2S mixed sacrificial agent was added into the reaction cell. The light source was a 4000–4500 µW/cm2 Xe-lamp (300 W), with a cut-off filter (λ ≥ 420 nm). Nitrogen was purged into the system for 30 min before the reaction to remove oxygen. The concentrations of H2 were measured with an online gas chromatograph (GC). Injection, column and detector in GC were 50 °C. A thermal conductivity detector (TCD) was used as detector. The hydrogen generation experimental conditions are shown in Table 1.

4. Conclusions

The Zn0.74Cu0.13In2S3.805 photocatalyst, in which ZnIn2S4 was doped with Cu+ and excess In, was prepared by a simple one-pot solvothermal method. From the SEM, XRD and XPS results, it is highly possible that Zn0.74Cu0.13In2S3.805 is a solid solution with a hexagonal ZnIn2S4 as a basic structure. Control of the band gap and suppression of electron–hole recombination were confirmed by doping ZnIn2S4 with Cu+ and excess In. In addition, an increase in the absorption wavelength range and improved catalytic activity were observed. The hydrogen generation rate by Zn0.74Cu0.13In2S3.805 was 2370 μmol g-1 h-1, which was almost five times larger compared with that obtained with ZnIn2S4. The present work provides a strategy for water splitting systems consisting of sulfide materials with narrow band gaps for efficient hydrogen production.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2073-4344/9/8/681/s1. Table S1: Expected composite components of photocatalyst (molar ratio), Table S2: Elemental ratios of ZnIn2S4, Zn0.87In2S3.87, Zn0.87Cu0.13In2S3.935 and Zn0.74Cu0.13In2S3.805 from XPS results, Figure S1: Tauc plots of (a) ZnIn2S4, (b) Zn0.87In2S3.87, (c) Zn0.87Cu0.13In2S3.935 and (d) Zn0.74Cu0.13In2S3.805, Figure S2: Valence-band XPS spectra of (a) ZnIn2S4, (b) Zn0.87In2S3.87, (c) Zn0.87Cu0.13In2S3.935 and (d) Zn0.74Cu0.13In2S3.805, Figure S3: TEM images of Zn0.74Cu0.13In2S3.805 (a) before and (b) after irradiation.

Author Contributions

I.T. and H.K. conceived and designed the experiments. I.T. performed the experiments and wrote the paper. I.T., M.F., S.K. and H.K. analyzed the results and advised the project.

Funding

This research received no external funding.

Note

All experiments were conducted at Mie University. Any opinions, findings, conclusions or recommendations expressed in this paper are those of the authors and do not necessarily reflect the view of the supporting organizations.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Skoplyak, O.; Barteau, M.A.; Chen, J.G.G. Comparison of H2 production from ethanol and ethylene glycol on M/Pt(1 1 1) (M = Ni, Fe, Ti) bimetallic surfaces. Catal. Today 2009, 147, 150–157. [Google Scholar] [CrossRef]
  2. Barthos, R.; Sze’chenyi, A.; Solymos, F. Efficient H2 Production from Ethanol over Mo2 C/C Nanotube Catalyst. Catal. Lett. 2008, 120, 161–165. [Google Scholar] [CrossRef]
  3. Baamran, K.S.; Tahir, M. Thermodynamic investigation and experimental analysis on phenol steam reforming towards enhanced H2 production over structured Ni/ZnTiO3 nanocatalyst. Energy Convers. Manag. 2019, 180, 796–810. [Google Scholar] [CrossRef]
  4. Politano, A.; Cattelan, M.; Boukhvalov, D.W.; Campi, D.; Cupolillo, A.; Agnoli, S.; Apostol, N.G.; Lacovig, P.; Lizzit, S.; Farías, D. Unveiling the Mechanisms Leading to H2 Production Promoted by Water Decomposition on Epitaxial Graphene at Room Temperature. ACS Nano 2016, 10, 4543–4549. [Google Scholar] [CrossRef] [PubMed]
  5. Kudo, A. Development of photocatalyst materials for water splitting. Int. J. Hydrogen Energy 2006, 31, 197–202. [Google Scholar] [CrossRef]
  6. Kato, H.; Kudo, A. Photocatalytic water splitting into H2 and O2 over various tantalate photocatalysts. Catal. Today 2003, 78, 561–569. [Google Scholar] [CrossRef]
  7. Chen, X.; Shen, S.; Guo, L.; Mao, S.S. Semiconductor-based Photocatalytic Hydrogen Generation. Chem. Rev. 2010, 110, 6503–6570. [Google Scholar] [CrossRef]
  8. Tsuji, I.; Kato, H.; Kobayashi, H.; Kudo, A. Photocatalytic H2 evolution reaction from aqueous solutions over band structure-controlled (Agln)xZn2(1−x)S2 solid solution photocatalysts with visible-light response and their surface nanostructures. J. Am. Chem. Soc. 2004, 126, 13406–13413. [Google Scholar] [CrossRef]
  9. Peng, S.; Zhu, P.; Thavasi, V.; Mhaisalkar, S.G.; Ramakrishna, S. Facile solution deposition of ZnIn2S4 nanosheet films on FTO substrates for photoelectric application. Nanoscale 2011, 3, 2602–2608. [Google Scholar] [CrossRef]
  10. Zhang, X.; Yu, L.; Zhuang, C.; Peng, T.; Li, R.; Li, X. Highly Asymmetric Phthalocyanine as a Sensitizer of Graphitic Carbon Nitride for Extremely Efficient Photocatalytic H2 Production under Near-Infrared Light. ACS Catal. 2014, 4, 162–170. [Google Scholar] [CrossRef]
  11. Dai, W.W.; Zhao, Z.Y. DFT study on the interfacial properties of vertical and in-plane BiOI/BiOIO3 hetero-structures, Phys. Chem. Chem. Phys. 2017, 19, 9900–9911. [Google Scholar]
  12. Vaiano, V.; Iervolino, G.; Rizzo, L. Cu-doped ZnO as efficient photocatalyst for the oxidation of arsenite to arsenate under visible light. Appl. Catal. B 2018, 238, 471–479. [Google Scholar] [CrossRef]
  13. Shafaee, M.; Goharshadi, E.K.; Mashreghi, M.; Sadeghinia, M. TiO2 nanoparticles and TiO2@graphene quantum dots nancomposites as effective visible/solar light photocatalysts. J. Photochem. Photobiol. A Chem. 2018, 357, 90–102. [Google Scholar] [CrossRef]
  14. Chang, C.J.; Wang, C.W.; Wei, Y.H.; Chen, C.Y. Enhanced photocatalytic H2 production activity of Ag-doped Bi2WO6-graphene based photocatalysts. Int. J. Hydrogen Energy 2018, 43, 11345–11354. [Google Scholar] [CrossRef]
  15. Jia, Y.; Zhao, D.; Li, M.; Han, H.; Li, C. La and Cr Co-doped SrTiO3 as an H2 evolution photocatalyst for construction of a Z-scheme overall water splitting system. Chin. J Catal. 2018, 39, 421–430. [Google Scholar] [CrossRef]
  16. Yan, X.; Xue, C.; Yang, B.; Yang, G. Novel three-dimensionally ordered macroporous Fe3+-doped TiO2 photocatalysts for H2 production and degradation applications. Appl. Surf. Sci. 2017, 394, 248–257. [Google Scholar] [CrossRef]
  17. Singh, J.; Sharma, S.; Sharma, S.; Chand Singh, R. Effect of tungsten doping on structural and optical properties of rutile TiO2 and band gap narrowing. Optik 2019, 182, 538–547. [Google Scholar] [CrossRef]
  18. Sharma, P.K.; Cortes, M.A.L.R.M.; Hamilton, J.W.J.; Han, Y.; Byrne, J.A.; Nolan, M. Surface modification of TiO2 with copper clusters for band gap narrowing. Catal. Today 2018, 321–322, 9–17. [Google Scholar] [CrossRef]
  19. Liu, C.; Chai, B.; Wang, C.; Yan, J.; Ren, Z. Solvothermal fabrication of MoS2 anchored on ZnIn2S4 microspheres with boosted photocatalytic hydrogen evolution activity. Int. J. Hydrogen Energy 2018, 43, 6977–6986. [Google Scholar] [CrossRef]
  20. Gao, P.; Li, A.; Sun, D.D.; Ng, W.J. Effects of various TiO2 nanostructures and graphene oxide on photocatalytic activity of TiO2. J. Hazard. Mater. 2014, 279, 96–104. [Google Scholar] [CrossRef]
  21. Guo, J.; Liao, X.; Lee, M.-H.; Hyett, G.; Huang, C.C.; Hewak, D.W.; Mails, S.; Zhou, W.; Jiang, Z. Experimental and DFT insights of the Zn-doping effects on the visible-light photocatalytic water splitting and dye decomposition over Zn-doped BiOBr photocatalysts. Appl. Catal. B 2019, 243, 502–512. [Google Scholar] [CrossRef]
  22. Behzadifard, Z.; Shariatinia, Z.; Jourshabani, M. Novel visible light driven CuO/SmFeO3 nanocomposite photocatalysts with enhanced photocatalytic activities for degradation of organic pollutants. J. Mol. Liq. 2018, 262, 533–548. [Google Scholar] [CrossRef]
  23. Li, T.L.; Cai, C.D.; Yeh, T.F.; Teng, H. Capped CuInS2 quantum dots for H2 evolution from water under visible light illumination. J. Alloys Compd. 2013, 550, 326–330. [Google Scholar] [CrossRef]
  24. You, D.; Pan, B.; Jiang, F.; Zhou, Y.; Su, W. CdS nanoparticles/CeO2 nanorods composite with high-efficiency visible-light-driven photocatalytic activity. Appl. Surf. Sci. 2016, 363, 154–160. [Google Scholar] [CrossRef]
  25. Jin, X.; Chen, F.; Jia, D.; Cao, Y.; Duan, H.; Long, M.; Yang, L. Influences of synthetic conditions on the photocatalytic performance of ZnS/graphene composites. J. Alloys Compd. 2019, 780, 299–305. [Google Scholar] [CrossRef]
  26. Zhang, S.; Wang, L.; Liu, C.; Luo, J.; Crittenden, J.; Liu, X.; Ca, T.; Yuan, J.; Pei, Y.; Liu, Y. Photocatalytic wastewater purification with simultaneous hydrogen production using MoS2 QD-decorated hierarchical assembly of ZnIn2S4 on reduced graphene oxide photocatalyst. Water Res. 2017, 121, 11–19. [Google Scholar] [CrossRef]
  27. Fan, B.; Chen, Z.H.; Liu, Q.; Zhang, Z.G.; Fang, X.M. One-pot hydrothermal synthesis of Ni-doped ZnIn2S4 nanostructured film photoelectrodes with enhanced photoelectrochemical performance. Appl. Surf. Sci. 2016, 370, 252–259. [Google Scholar] [CrossRef]
  28. Ding, Y.; Gao, Y.; Li, Z. Carbon quantum dots (CQDs) and Co(dmgH)2PyCl synergistically promote photocatalytic hydrogen evolution over hexagonal ZnIn2S4. Appl. Surf. Sci. 2018, 462, 255–262. [Google Scholar] [CrossRef]
  29. Ding, Y.; Wei, D.; He, R.; Yuan, R.; Xie, T.; Li, Z. Rational Design of Z-Scheme PtS-ZnIn2S4/WO3-MnO2 for Overall Photo-catalytic Water Splitting under Visible Light. Appl. Catal. B Environ. 2019, 258, 117948. [Google Scholar] [CrossRef]
  30. Zhao, C.; Zhang, Y.; Jiang, H.; Chen, J.; Liu, Y.; Liang, Q.; Zhou, M.; Li, Z.; Zhou, Y. Combined Effects of Octahedron NH2-UiO-66 and Flowerlike ZnIn2S4 Microspheres for Photocatalytic Dye Degradation and Hydrogen Evolution Under Visible Light. J. Phys. Chem. C 2019, 123, 18037–18049. [Google Scholar] [CrossRef]
  31. Liu, A.; Yu, C.; Lin, J.; Sun, G.; Xu, G.; Huang, Y.; Liu, Z.; Tang, C. Construction of CuInS2@ZIF-8 nanocomposites with enhanced photocatalytic activity and durability. Mater. Res. Bull. 2019, 112, 147–153. [Google Scholar] [CrossRef]
  32. Shen, S.; Zhao, L.; Zhou, Z.; Guo, L. Enhanced Photocatalytic Hydrogen Evolution over Cu-Doped ZnIn2S4 under Visible Light Irradiation. J. Phys. Chem. C 2008, 112, 16148–16155. [Google Scholar] [CrossRef]
  33. Zhang, G.; Zhang, W.; John, C.C.; Chen, Y.; Minakata, D.; Wang, P. Photocatalytic hydrogen production under visible-light irradiation on (CuAg)0.15In0.3Zn1.4S2 systhesized by precipitation and calcination. Chin. J. Catal. 2013, 34, 1926–1935. [Google Scholar] [CrossRef]
  34. Shen, S.H.; Zhao, L.; Guo, L.J. ZnmIn2S3+m (m=1–5, integer): A new series of visible-light-driven photocatalysts for splitting water to hydrogen. Int. J. Hydrogen Energy 2010, 35, 10148–10154. [Google Scholar] [CrossRef]
  35. Sun, M.; Zhao, X.; Zeng, Q.; Yan, T.; Ji, P.G.; Wu, T.T.; Wei, D.; Du, B. Facile synthesis of hierarchical ZnIn2S4/CdIn2S4 microspheres with enhanced visible light driven photocatalytic activity. Appl. Surf. Sci. 2017, 407, 328–336. [Google Scholar] [CrossRef]
  36. Song, K.; Zhu, R.; Tian, F.; Cao, G.; Ouyang, F. Effects of indium contents on photocatalytic performance of ZnIn2S4 for hydrogen evolution under visible light. J. Solid State Chem. 2015, 232, 138–143. [Google Scholar] [CrossRef]
  37. Du, C.; Zhang, Q.; Lin, Z.; Yan, B.; Xia, C.; Yang, G. Half-unit-cell ZnIn2S4 monolayer with sulfur vacancies for photocatalytic hydrogen evolution. Appl. Catal. B. 2019, 248, 193–201. [Google Scholar] [CrossRef]
  38. Yang, X.; Li, Q.; Lv, K.L.; Li, M. Heterojunction construction between TiO2 hollowsphere and ZnIn2S4 flower for photocatalysis application. Appl. Surf. Sci. 2017, 398, 81–88. [Google Scholar]
  39. Yue, W.; Han, S.; Peng, R.; Shen, W.; Geng, H.; Wu, F.; Tao, S.; Wang, M. CuInS2 quantum dots synthesized by a solvothermal route and their application as effective electron acceptors for hybrid solar cells. J. Mater. Chem. 2010, 20, 7570–7578. [Google Scholar] [CrossRef]
  40. Zepeda, T.A.; Pawelec, B.; Díaz de León, J.N.; de los Reyes, J.A.; Olivas, A. Effect of gallium loading on the hydrodesulfurization activity of unsupported Ga2S3/WS2 catalysts. Appl. Catal. B 2012, 111–112, 10–19. [Google Scholar] [CrossRef]
  41. Chen, S.; Li, S.; Xiong, L.; Wang, G. In-situ growth of ZnIn2S4 decorated on electrospun TiO2 nanofibers with enhanced visible-light photocatalytic activity. Chem. Phys. Lett. 2018, 706, 68–75. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) ZnIn2S4, (b) Zn0.87In2S3.87, (c) Zn0.87Cu0.13In2S3.935 and (d) Zn0.74Cu0.13In2S3.805.
Figure 1. XRD patterns of (a) ZnIn2S4, (b) Zn0.87In2S3.87, (c) Zn0.87Cu0.13In2S3.935 and (d) Zn0.74Cu0.13In2S3.805.
Catalysts 09 00681 g001
Figure 2. X-ray photoelectron spectroscopy (XPS) narrow and survey spectra of Zn0.74Cu0.13In2S3.805. (a) survey, (b) Zn, (c) In, (d) Cu and (e) S.
Figure 2. X-ray photoelectron spectroscopy (XPS) narrow and survey spectra of Zn0.74Cu0.13In2S3.805. (a) survey, (b) Zn, (c) In, (d) Cu and (e) S.
Catalysts 09 00681 g002
Figure 3. SEM images of (a) ZnIn2S4 and (b) Zn0.74Cu0.13In2S3.805.
Figure 3. SEM images of (a) ZnIn2S4 and (b) Zn0.74Cu0.13In2S3.805.
Catalysts 09 00681 g003
Figure 4. UV-visible spectra of (a) ZnIn2S4, (b) Zn0.87In2S3.87, (c) Zn0.87Cu0.13In2S3.935 and (d) Zn0.74Cu0.13In2S3.805.
Figure 4. UV-visible spectra of (a) ZnIn2S4, (b) Zn0.87In2S3.87, (c) Zn0.87Cu0.13In2S3.935 and (d) Zn0.74Cu0.13In2S3.805.
Catalysts 09 00681 g004
Figure 5. Photoluminescence spectra for (a) ZnIn2S4, (b) Zn0.87In2S3.87, (c) Zn0.87Cu0.13In2S3.935 and (d) Zn0.74Cu0.13In2S3.805. Excitation: 350 nm.
Figure 5. Photoluminescence spectra for (a) ZnIn2S4, (b) Zn0.87In2S3.87, (c) Zn0.87Cu0.13In2S3.935 and (d) Zn0.74Cu0.13In2S3.805. Excitation: 350 nm.
Catalysts 09 00681 g005
Figure 6. Photocatalytic hydrogen production rate with (a) ZnIn2S4, (b) Zn0.87In2S3.87, (c) Zn0.87Cu0.13In2S3.935 and (d) Zn0.74Cu0.13In2S3.805.
Figure 6. Photocatalytic hydrogen production rate with (a) ZnIn2S4, (b) Zn0.87In2S3.87, (c) Zn0.87Cu0.13In2S3.935 and (d) Zn0.74Cu0.13In2S3.805.
Catalysts 09 00681 g006
Scheme 1. H2 production mechanism.
Scheme 1. H2 production mechanism.
Catalysts 09 00681 sch001
Table 1. Hydrogen generation experimental conditions.
Table 1. Hydrogen generation experimental conditions.
PhotocatalystZnIn2S4, Zn0.87In2S3.87, Zn0.87Cu0.13In2S3.935, Zn0.74Cu0.13In2S3.805
Cocatalyst0.04 ppm H2PtCl6 10 mL (1.0 wt%)
Medium0.25 M Na2SO3 / 0.35 M Na2S 40 mL
ReactorPyrex glass vessel (volume: 123 mL)
TemperatureRoom temperature (25 °C)
Light sourceXenon lamp (λ ≥ 420 nm, 4500 µW/cm2)
Irradiation time6 h
AnalysisGas chromatography (TCD)

Share and Cite

MDPI and ACS Style

Tateishi, I.; Furukawa, M.; Katsumata, H.; Kaneco, S. Improvement of Photocatalytic H2-Generation under Visible Light Irradiation by Controlling the Band Gap of ZnIn2S4 with Cu and In. Catalysts 2019, 9, 681. https://0-doi-org.brum.beds.ac.uk/10.3390/catal9080681

AMA Style

Tateishi I, Furukawa M, Katsumata H, Kaneco S. Improvement of Photocatalytic H2-Generation under Visible Light Irradiation by Controlling the Band Gap of ZnIn2S4 with Cu and In. Catalysts. 2019; 9(8):681. https://0-doi-org.brum.beds.ac.uk/10.3390/catal9080681

Chicago/Turabian Style

Tateishi, Ikki, Mai Furukawa, Hideyuki Katsumata, and Satoshi Kaneco. 2019. "Improvement of Photocatalytic H2-Generation under Visible Light Irradiation by Controlling the Band Gap of ZnIn2S4 with Cu and In" Catalysts 9, no. 8: 681. https://0-doi-org.brum.beds.ac.uk/10.3390/catal9080681

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