Next Article in Journal
Infrared Spectrum and UV-Induced Photochemistry of Matrix-Isolated Phenyl 1-Hydroxy-2-Naphthoate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

V-Substituted ZnIn2S4: A (Visible+NIR) Light-Active Photocatalyst

Instituto de Catálisis y Petroleoquímica, Consejo Superior de Investigaciones Científicas, 28049 Madrid, Spain
*
Author to whom correspondence should be addressed.
Submission received: 2 December 2020 / Revised: 20 December 2020 / Accepted: 23 December 2020 / Published: 7 January 2021

Abstract

:
ZnIn2S4 is known to be a visible light-active photocatalyst. In this work, it is shown that by substituting part of the In atoms with vanadium, the visible light range of photocatalytic activity of such material can be extended, using the so-called in-gap band scheme that has been shown to enhance photovoltaic characteristics. Characterization of this material using several techniques, complemented by DFT calculations, will support this statement. While here only the degradation of aqueous HCOOH in well-aerated conditions is discussed, the same material may be used, with an adequate sacrificial reagent, for photocatalytic H2 generation.

1. Introduction

Use of sulphides as photocatalysts is a well-studied subject, as shown by several very recent reviews [1,2,3,4,5,6]. In particular, ZnIn2S4 is known to be active in a significant part of the visible light spectrum; it has been used for H2 production [7,8], microorganism inactivation [9] or dye degradation [10]. On the other hand, ZnIn2S4 has been combined with other semiconductors to improve photocatalytic properties. Thus, it has been combined with oxides, like in [11,12,13,14,15]; it has also been combined with other sulphides, as in [16,17,18,19], or even nitrides and phosphides [20,21]. Additionally, some reviews on its use, or on the use of similar materials, have appeared [22,23,24,25], and our group has done a recent publication including references to the ZnIn2S4 system [26].
Here, the possibility of enlarging the spectral range in which ZnIn2S4 is photocatalytically active will be addressed, in particular, through substitution of indium by vanadium. The idea is based on a claim made by Luque and Martí [27] that by introducing intermediate levels in photovoltaic cells it may be possible, in theory, to enhance by more than 50% the solar efficiency of such systems. In the past, our group has addressed, using DFT calculations, the possibility of achieving such an electronic structure [28,29,30,31,32,33]; using photocatalysis, this has then proven to be the case in a couple of examples [31,34], while promising experimental results indicating that a similar electronic structure has been achieved have been provided recently [35].
In the following, a recall is presented of the data obtained by our group with undoped ZnIn2S4, and then relevant data on V-substituted ZnIn2S4 (including DFT results) and on its photocatalytic properties are given in detail.

2. Experimental Section

The reactants used were ZnCl2, InCl3, VCl3 and Na2S as the source of sulfur. The solvent used, in order to avoid as much as possible the oxidation of VCl3, was one with reducing characteristics, being composed of ethyleneglycol (in the following, shortened to EG) and MilliQ distilled water (to facilitate the solubility of the salts), water being present at 20% by weight. All reactants were supplied by Aldrich; the In and Zn salts had purity of 99.999%, while the others had purity of 99%, except EG, which was supplied by Panreac with purity of 99%.
The solids were made, after adjusting the pH to 3 with concentrated HCl (also from Panreac), using solvothermal treatments in the specified solvent at 180 °C during 65 h, with the objective of obtaining at the end 1 g of solid. The latter was filtered out from the solution and washed repeatedly with water and methanol, and finally with water. The objective was to prepare V-substituted ZnIn2S4 with a V–In ratio of 1:9.
Chemical analysis data, BET-specific surfaces, X-ray diffractograms, UV-Vis-NIR spectra and photocatalytic tests of aqueous HCOOH degradation (including spectral response profiles using a collection of filters coupled to a Xenon lamp) were obtained and analyzed as reported previously [36,37]. EPR data were obtained and analyzed as reported in the Supporting Information of Ref. [38]. DFT calculations were carried out with the VASP code [39], using the PAW method to represent the cores [40]; the cutoff energy for electron functions was set at 300 eV and the Brillouin zone was sampled with a (12 × 12 × 2) Monkhorst-Pack grid. Because this material has layers interacting only via van der Waals forces, a correction to the energy, using the method proposed in [41], was added and used in the relaxations.
Irradiation was provided by an ozone-free 450 W Xe lamp provided with a water filter (to decrease IR light) and, where necessary, with bandpass filters having FWHM = 50 nm. The magnetically stirred and well-aerated suspension, adjusted with a phosphate buffer to pH = 2.5 (the natural pH of the HCOOH 1.5 mM solution utilized) and using 40 mg of photocatalyst in 80 mL of the EG-water mixture, was sampled periodically and, after filtering out the catalyst, the HCOOH concentration was measured using the UV absorption of HCOOH at λ = 205 nm.

3. Results and Discussion

3.1. Previous Data on Undoped ZnIn2S4

As reported in [26], the diffractogram, UV-Vis diffuse reflectance spectrum and spectral dependence of the photocatalytic activity for the degradation of aqueous HCOOH against wavelength were as depicted in Figure 1. It can be seen that the diffractogram presents well-resolved peaks. This contrasts with the diffractograms shown below in Figure 2, the reason being that, because there was no need to keep a reducing environment, only distilled water was used in that case for the solvothermal synthesis. The bandgap measured with the Tauc plot (2.6 eV) allows a significant use of visible light; the spectral response for photocatalytic degradation of aqueous HCOOH matches well the spectrum of ZnIn2S4, once the width of the filters used (FWHM = 50 nm) and the possibility of a few defects narrowing a bit the bandgap are taken into account.

3.2. New Data on V-Doped ZnIn2S4

3.2.1. Characterization of V-Substituted ZnIn2S4

Chemical analysis yielded a result close to the intended one (V–In ratio = 1:8.6). However, its diffractogram, shown in Figure 2, displayed peaks less resolved than in Figure 1a above; indeed, they are much closer to those of undoped ZnIn2S4 prepared using Na2S as well and the same EG-water mixture. This probably means that these systems contain more defects than the undoped ZnIn2S4 material. In any case, these diffractograms are very close to those reported in Figure 3 of Ref. [10] and Figure 10A of Ref. [22]; all of them correspond to the hexagonal phase of ZnIn2S4. On the other hand, the peaks for the system containing V appear displaced to slightly higher angles, as expected considering the smaller ionic radius of V in comparison with In.
The diffuse reflectance spectrum of V-substituted ZnIn2S4 prepared solvothermally with Na2S and the EG-water mixture, after obtaining the Kubelka-Munk transform and plotting it against the photon energy, is presented in Figure 3. It can be seen that, in comparison with the same data presented above for undoped ZnIn2S4, absorption is much extended into the visible and even IR range (it must be noted, however, that the small peaks below 0.7 eV are overtones of vibrations). No data are available on diffuse reflectance spectra upon substitution by V in this sulphide; the closest examples are probably those found in Figure 5b of Ref. [31] and Figure 3a of Ref. [38].
EPR spectra were also obtained for this V-substituted sample; they are shown below in Figure 4. Several things must be noted: (a) The shape of the spectra, showing hyperfine peaks, is typical of V4+ (V3+ is normally not seen in EPR, since the high fine structure leads to having all features outside the normal magnetic field range); indeed, it is very similar, except for the sharp peak at g ≈ 2.00 (surely due to some type of defect), to that found in the Supporting Information of Ref. [38], which shows the V4+ spectrum in V-substituted In2S3. (b) The ratio between doubly integrated areas of the spectra at 77 K and ambient temperature is close to 4:1, as can be expected for a normal paramagnetic system. (c) Most importantly: quantification of the absolute number of spins in comparison with a CuSO4 standard indicates that not more than 14% of vanadium is present as V4+, which means that the rest is present as V3+, as expected considering the reducing power of EG at the temperature of the solvothermal treatment.
DFT calculations were carried out for this V-substituted ZnIn2S4. Here, V3+ species substitute for In3+ octahedral cations. Indeed, V3+ ions usually prefer octahedral coordination, and it must be noted that the layered structure of ZnIn2S4 has all Zn cations in tetrahedral coordination while 50% of In cations are in octahedral coordination and the other 50% are in tetrahedral coordination. In this case, a V–In ratio of 1:7 was chosen, to approach the 1:8.6 ratio found experimentally. The result of this calculation is presented in Figure 5. It can be clearly seen that an in-gap band appears (separated from the valence and conduction bands) that is partially occupied, i.e., is crossed by the Fermi level (represented by the dashed vertical line). This in-gap band is rather close to the valence band, which explains that the transitions from its empty part to the wide valence band may cover completely the visible range and a substantial amount of the IR range, as shown in Figure 3. On the other hand, the distance between the conduction band and the filled part of the in-gap band justifies the small hump appearing in Figure 3 around 1.9 eV. The main rise in the absorption spectrum corresponds, obviously, to the main bandgap of ZnIn2S4. On the other hand, it is well known that DFT calculations at the PBE level significantly underestimate bandgaps in semiconductors; where the in-gap band induced by vanadium might appear once the bandgap widens is still an unresolved question.

3.2.2. Photocatalytic Activity of V-Substituted ZnIn2S4

The photocatalytic activity of this material for the degradation of aqueous HCOOH was then examined. First, the decrease in HCOOH concentration was determined for a series of bandpass filters (and also with no filter); the result in each case is displayed in Figure 6a. Then, each curve was fitted assuming first order kinetics; the results of rate constant k, compared with the same result obtained for the undoped ZnIn2S4, and also compared with the absorption spectrum of each sample, is shown in Figure 6b. It is clearly seen that the spectral response of V-substituted ZnIn2S4, although somewhat smaller in the 400–500 nm range (which can be ascribed to the presence of defects, including those introduced by vanadium), is on the other hand widened, reaching even the NIR range (as observed for the small photoactivity found at λ = 750 nm). This is just what can be expected considering the DFT results in Figure 5, which indicate a transition of lower energy from the filled part of the in-gap band to the conduction band (one must recall that spin flips during light absorption events are not allowed). Therefore, the effect of the in-gap band on the general electronic structure is confirmed by the photocatalytic experiments.
A caveat must be mentioned here. As shown in Ref. [26], in the photodegradation of aqueous HCOOH using undoped ZnIn2S4 (in well-aerated conditions), a significant photocorrosion of that sulphide took place. This has not been verified here for V-substituted ZnIn2S4, but a similar effect might happen as well since this latter material has more defects (as evidenced by its diffractogram); the increase in Madelung-type stability due to the smaller radius of the vanadium ion may compensate this only partially. However, this might not be the case when using this material for the photocatalytic generation of H2, especially when utilizing a sacrificial reagent, as was the case in Refs. [7,8]. Therefore, the interest of the work presented here, having as main objective the extension to higher wavelengths of the photocatalytic response when using V-substituted ZnIn2S4, might be applied to the situations when photoproduction of H2 with visible light is the main objective.

3.2.3. Likely Degradation Products and Reaction Mechanism

As stated in Ref. [36], the most likely degradation mechanism corresponds to reaction
HCOOH + 1 2 O 2 CO 2 + H 2 O
Other degradations, like
HCOOH CO 2 + H 2
or
2 HCOOH HCOO 2 + H 2
are unlikely since such degradations usually require Pt or a similar metal to generate H2, and the only other alternative
HCOOH CO + H 2 O
is also unlikely since such reaction requires usually a rather acidic catalyst, which is not the case here.
Concerning the reaction mechanism, there are two main possibilities: via OH or O2 radicals. The former are likely to be present, since, as shown in Ref. [34], both V-substituted and V-free In2S3 (sulphides very similar to that used here) generate OH radicals, as evidenced by the terephtalic acid test; these radicals are indeed very aggressive and reactive. The participation of superoxide (or a HO2 radical, for that matter) is less clear; however, it cannot be discarded, since, as shown in Ref. [26], where the degradation mechanism of the rhodamine B dye was studied using In2S3 as photocatalyst (again, a sulphide similar to that used here), there was a large difference between bubbling O2 and bubbling N2: the last steps of rhodamine B degradation proceeded much more slowly when N2 was bubbled. Therefore, both OH and O2 radicals might participate in the mechanism of the final degradation, which should proceed according to the first reaction above. Concerning singlet oxygen, it does not seem to play a large role in photocatalysis [42]; the present authors only know one reference in which it plays a significant role in photocatalysis including a visible light-active sulphide [43]. H2O2, on the other hand, is much less reactive than OH or O2/HO2 radicals; it is not likely to have a significant role here.

3.2.4. General Discussion

It is evidenced that including vanadium in the solvothermal synthesis of ZnIn2S4 leads to a noticeable extension in the visible light range of the photocatalytic response, at least for the HCOOH photodegradation; comparison of this result with that found in Ref. [26] (the only other case in which photocatalysis involving aqueous HCOOH and ZnIn2S4 has been addressed) makes this rather clear. As said in the last paragraph of Section 3.2.3, in these conditions, ZnIn2S4 (V-substituted or not) may suffer photocorrosion. In most cases, this photocorrosion effect can be avoided or reduced by combining with other materials [44,45,46]. However, it may be that as said above such photocorrosion does not occur when trying to photogenerate H2 using a sacrificial reagent, as is the case not only of Refs. [7,8] (in the latter case, using photoelectrochemistry), which are the first ones that the present authors could detect; additionally, other more recent cases have appeared [24,47,48,49,50]. In fact, not only H2 can be produced by ZnIn2S4; in recent years, it has also been found possible to photoreduce CO2 or produce syngas, in some cases combining two semiconductors [51,52,53,54]. In many of these cases, stable and reproducible behavior was always observed.

4. Conclusions

The effects of substituting part of the In atoms by vanadium are evidenced here: once a proper solvent and source of sulfur are chosen, an in-gap band partially filled is formed that leads to an extension of the wavelength range of photocatalytic activity, even entering a bit into the NIR range. The possibility of extending this effect to photocatalytic H2 production would be the succeeding step to follow.

Author Contributions

R.L. carried out the experimental work; R.L. and J.C.C. carried out the analysis and discussed the conclusions. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Spanish Ministerio de Economía y Competitividad, grant number ENE2016-77798-C4 (project SEHTOP) and by Comunidad de Madrid, grant number S2013/MAE-2780 (project MADRID-PV). It is related also to COST Action 18234, supported by COST (European Cooperation in Science and Technology).

Data Availability Statement

Data contained within the article is available on request from the authors.

Acknowledgments

Thanks are given to CSIC for the use of parallel computer trueno where the DFT calculations were made.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, S.; Sun, J.; Guan, J. Strategies to improve electrocatalytic and photocatalytic performance of two-dimensional materials for hydrogen evolution reaction. Chin. J. Catal. 2021, 42, 511–556. [Google Scholar] [CrossRef]
  2. Nasir, J.A.; ur Rehman, Z.; Ali Shah, S.N.; Khan, A.; Butler, I.S.; Catlow, C.R.A. Recent developments and perspectives in CdS based photocatalysts for water splitting. J. Mater. Chem. A 2020, 8, 20735–21342. [Google Scholar] [CrossRef]
  3. Zhang, W.; Mohamed, A.R.; Ong, W.J. Z-Scheme Photocatalytic Systems for Carbon Dioxide Reduction: Where Are We Now? Angew. Chem. Int. Ed. 2020, 59, 2–24. [Google Scholar] [CrossRef] [PubMed]
  4. Palchoudhury, S.; Ramasamy, K.; Gupta, A. Multinary copper-based chalcogenide nanocrystal systems from the perspective of device applications. Nanoscale Adv. 2020, 2, 3069–3082. [Google Scholar] [CrossRef]
  5. Li, J.; Jiménez-Calvo, P.; Paineau, E.; Ghazzal, M.N. Metal Chalcogenides Based Heterojunctions and Novel Nanostructures for Photocatalytic Hydrogen Evolution. Catalysts 2020, 10, 89. [Google Scholar] [CrossRef] [Green Version]
  6. Chandrasekaran, S.; Yao, L.; Deng, L.; Bowen, C.; Zhang, Y.; Chen, S.; Lin, Z.; Peng, F.; Zhang, P. Recent advances in metal sulfides: From controlled fabrication to electrocatalytic, photocatalytic and photoelectrochemical water splitting and beyond. Chem. Soc. Rev. 2019, 48, 4178–4280. [Google Scholar] [CrossRef]
  7. Lei, Z.; You, W.; Liu, M.; Zhou, G.; Takata, T.; Hara, M.; Domen, K.; Li, C. Photocatalytic water reduction under visible light on a novel ZnIn2S4 catalyst synthesized by hydrothermal method. Chem. Commun. 2003, 17, 2142–2143. [Google Scholar] [CrossRef]
  8. Li, M.; Su, J.; Guo, L. Preparation and characterization of ZnIn2S4 thin films deposited by spray pyrolysis for hydrogen production. Int. J. Hydrogen Energy 2008, 33, 2891–2896. [Google Scholar] [CrossRef]
  9. Yu, H.; Quan, X.; Zhang, Y.; Ma, N.; Chen, S.; Zhao, H. Electrochemically Assisted Photocatalytic Inactivation of Escherichia coli under Visible Light Using a ZnIn2S4 Film Electrode. Langmuir 2008, 24, 7599–7604. [Google Scholar] [CrossRef]
  10. Chen, Z.; Li, D.; Zhang, W.; Chen, C.; Li, W.; Sun, M.; He, Y.; Fu, X. Low-Temperature and Template-Free Synthesis of ZnIn2S4 Microspheres. Inorg. Chem. 2008, 47, 9766–9772. [Google Scholar] [CrossRef]
  11. Bai, Z.; Yan, X.; Kang, Z.; Hu, Y.; Zhang, X.; Zhang, Y. Photoelectrochemical performance enhancement of ZnO photoanodes from ZnIn2S4 nanosheets coating. Nano Energy 2015, 14, 392–400. [Google Scholar] [CrossRef]
  12. Wei, N.; Wu, Y.; Wang, M.; Sun, W.; Li, Z.; Ding, L.; Cui, H. Construction of noble-metal-free TiO2 nanobelt/ZnIn2S4 nanosheet heterojunction nanocomposite for highly efficient photocatalytic hydrogen evolution. Nanotechnology 2019, 30, 045701. [Google Scholar] [CrossRef] [PubMed]
  13. Ben Assaker, I.; Gannouni, M.; Ben Naceur, J.; Almessiere, M.A.; Al-Otaibi, A.L.; Ghrib, T.; Shen, S.; Chtourou, R. Electrodeposited ZnIn2S4 onto TiO2 thin films for semiconductor-sensitized photocatalytic and photoelectrochemical applications. Appl. Surf. Sci. 2015, 351, 927–934. [Google Scholar] [CrossRef]
  14. Yang, C.; Li, Q.; Xia, Y.; Lv, K.; Li, M. Enhanced visible-light photocatalytic CO2 reduction performance of ZnIn2S4 microspheres by using CeO2 as cocatalyst. Appl. Surf. Sci. 2019, 464, 388–395. [Google Scholar] [CrossRef]
  15. Yuan, D.; Sun, M.; Tang, S.; Zhang, Y.; Wang, Z.; Qi, J.; Rao, Y.; Zhang, Q. All-solid-state BiVO4/ZnIn2S4 Z-scheme composite with efficient charge separations for improved visible light photocatalytic organics degradation. Chin. Chem. Lett. 2020, 31, 547–550. [Google Scholar] [CrossRef]
  16. Wei, L.; Chen, Y.; Zhao, J.; Li, Z. Preparation of NiS/ZnIn2S4 as a superior photocatalyst for hydrogen evolution under visible light irradiation. Beilstein J. Nanotech. 2013, 4, 949–955. [Google Scholar] [CrossRef] [Green Version]
  17. Hou, J.; Yang, C.; Cheng, H.; Wang, Z.; Jiao, S.; Zhu, H. Ternary 3D architectures of CdS QDs/graphene/ZnIn2S4 heterostructures for efficient photocatalytic H2 production. Phys. Chem. Chem. Phys. 2013, 15, 15660–15668. [Google Scholar] [CrossRef]
  18. Guo, X.; Peng, Y.; Liu, G.; Xie, G.; Guo, Y.; Zhang, Y.; Yu, J. An Efficient ZnIn2S4@CuInS2 Core-Shell p-n Heterojunction to Boost Visible-Light Photocatalytic Hydrogen Evolution. J. Phys. Chem. C 2020, 124, 5934–5943. [Google Scholar] [CrossRef]
  19. Pan, J.; Guan, Z.; Yang, J.; Li, Q. Facile fabrication of ZnIn2S4/SnS2 3D heterostructure for efficient visible-light photocatalytic reduction of Cr(VI). Chin. J. Catal. 2020, 41, 200–208. [Google Scholar] [CrossRef]
  20. Liu, H.; Jin, Z.; Xu, Z.; Zhang, Z.; Ao, D. Fabrication of ZnIn2S4-g-C3N4 sheet-on-sheet nanocomposites for efficient visible-light photocatalytic H2-evolution and degradation of organic pollutants. RSC Adv. 2015, 5, 97951–97961. [Google Scholar] [CrossRef]
  21. Li, X.; Wang, X.; Zhu, J.; Li, Y.; Zhao, J.; Li, F. Fabrication of two-dimensional Ni2P/ZnIn2S4 heterostructures for enhanced photocatalytic hydrogen evolution. Chem. Eng. J. 2018, 353, 15–24. [Google Scholar] [CrossRef]
  22. Pan, Y.; Yuan, X.; Jiang, L.; Yu, H.; Zhang, J.; Wang, H.; Guan, R.; Zeng, G. Recent advances in synthesis, modification and photocatalytic applications of micro/nano-structured zinc indium sulphide. Chem. Eng. J. 2018, 354, 407–431. [Google Scholar] [CrossRef]
  23. Li, Y.; Gao, C.; Long, R.; Xiong, Y. Photocatalyst design based on two-dimensional materials. Mater. Today Chem. 2019, 11, 197–216. [Google Scholar] [CrossRef]
  24. Mao, S.S.; Shena, S.; Guo, L. Nanomaterials for renewable hydrogen production, storage and utilization. Progr. Nat. Sci. Mater. Intern. 2012, 22, 522–534. [Google Scholar] [CrossRef] [Green Version]
  25. Yao, J.; Lu, H. Recent Advances in Liquid-phase Heterogeneous Photocatalysis for Organic Synthesis by Selective Oxidation. Curr. Org. Chem. 2014, 18, 1365–1372. [Google Scholar]
  26. Lucena, R.; Conesa, J.C. Photocatalysis with octahedral sulfides. In Current Developments in Photocatalysis and Photocatalytic Materials-New Horizons in Photocatalysis; Wang, D., Anpo, M., Fu, X., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; ISBN 978-0-12-819000-5. [Google Scholar]
  27. Luque, A.; Martí, A. Increasing the efficiency of ideal solar cells by photon induced transitions at intermediate levels. Phys. Rev. Lett. 1997, 78, 5014–5017. [Google Scholar] [CrossRef]
  28. Palacios, P.; Fernández, J.J.; Sánchez, K.; Conesa, J.C.; Wahnón, P. First-principles investigation of isolated band formation in half-metal TixGa1−xP compound at different dilutions. Phys. Rev. B 2006, 73, 085206. [Google Scholar] [CrossRef] [Green Version]
  29. Palacios, P.; Sánchez, K.; Wahnón, P.; Conesa, J.C. Characterization by Ab-initio Calculations of an Intermediate Band Material Based on Chalcopyrite Semiconductors Substituted by Several Transition Metals. J. Sol. Energy Eng. 2007, 129, 314–318. [Google Scholar] [CrossRef]
  30. Palacios, P.; Aguilera, I.; Sánchez, K.; Conesa, J.C.; Wahnón, P. Transition Metal Substituted Indium Thiospinels as Novel Intermediate Band Materials: Prediction and Understanding of their Electronic Properties. Phys. Rev. Lett. 2008, 101, 046403. [Google Scholar] [CrossRef] [Green Version]
  31. Wahnón, P.; Conesa, J.C.; Palacios, P.; Lucena, R.; Aguilera, I.; Seminovski, Y.; Fresno, F. V-doped SnS2: A new intermediate band material for a better use of the solar spectrum. Phys. Chem. Chem. Phys. 2011, 13, 20401–20407. [Google Scholar] [CrossRef] [Green Version]
  32. García, G.; Palacios, P.; Menéndez-Proupin, E.; Montero-Alejo, A.L.; Conesa, J.C.; Wahnón, P. Influence of chromium hyperdoping on the electronic structure of CH3NH3PbI3 perovskite: A first-principles insight. Sci. Rep. 2018, 8, 2511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Sánchez-Palencia, P.; García, G.; Conesa, J.C.; Wahnón, P.; Palacios, P. Spinel-Type Nitride Compounds with improved features as Solar Cell Absorbers. Acta Mater. 2020, 197, 316–329. [Google Scholar] [CrossRef]
  34. Lucena, R.; Conesa, J.C.; Aguilera, I.; Palacios, P.; Wahnón, P. V-substituted In2S3: An intermediate band material with photocatalytic activity in the whole visible light range. J. Mater. Chem. A 2014, 2, 8236–8245. [Google Scholar] [CrossRef] [Green Version]
  35. Serrano-Sánchez, F.; Conesa, J.C.; Rodrigues, J.E.; Marini, C.; Martínez, J.L.; Alonso, J.A. Divalent chromium in the octahedral positions of the novel hybrid perovskites CH3NH3Pb1−xCrx(Br,Cl)3 (x = 0.25, 0.5): Induction of narrow bands inside the bandgap. J. Alloys Comp. 2020, 821, 153414. [Google Scholar] [CrossRef]
  36. Lucena, R.; Fresno, F.; Conesa, J.C. Spectral response and stability of In2S3 as visible light active photocatalyst. Catal. Commun. 2012, 20, 1–5. [Google Scholar] [CrossRef] [Green Version]
  37. Lucena, R.; Fresno, F.; Conesa, J.C. Hydrothermally synthesized nanocrystalline tin disulphide as visible light-active photocatalyst: Spectral response and stability. Appl. Catal. A Gen. 2012, 415, 111–117. [Google Scholar] [CrossRef] [Green Version]
  38. Lucena, R.; Aguilera, I.; Palacios, P.; Wahnón, P.; Conesa, J.C. Synthesis and Spectral Properties of Nanocrystalline V-Substituted In2S3, a Novel Material for More Efficient Use of Solar Radiation. Chem. Mater. 2008, 20, 5125–5127. [Google Scholar] [CrossRef] [Green Version]
  39. Kresse, G.; Hafner, J. Abinitio molecular-dynamics for liquid-metals. Phys. Rev. B 1993, 47, 558–561. [Google Scholar] [CrossRef] [PubMed]
  40. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  41. Bucko, T.; Lebegue, S.; Angyan, J.G.; Hafner, J. Extending the applicability of the Tkatchenko-Scheffler dispersion correction via iterative Hirshfeld partitioning. J. Chem. Phys. 2014, 141, 034114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Rangarajan, G.; Yan, N.; Farnood, R. High-performance photocatalysts for the selective oxidation of alcohols to carbonyl compounds. Can. J. Chem. Eng. 2020, 98, 2259–2293. [Google Scholar] [CrossRef]
  43. Ali, M.M.; Sandhya, K.N.Y. One-step solvothermal synthesis of carbon doped TiO2-MoS2 heterostructure composites with improved visible light catalytic activity. New J. Chem. 2016, 40, 8123–8130. [Google Scholar] [CrossRef]
  44. Chen, J.; Zhang, H.; Liu, P.; Li, Y.; Liu, X.; Li, G.; Wong, P.K.; An, T.; Zhao, H. Cross-linked ZnIn2S4/rGO composite photocatalyst for sunlight-driven photocatalytic degradation of 4-nitrophenol. Appl. Catal. B Environ. 2015, 168, 266–273. [Google Scholar] [CrossRef]
  45. Tian, Q.; Wu, W.; Liu, J.; Wu, Z.; Yao, W.; Ding, J.; Jiang, C. Dimensional heterostructures of 1D CdS/2D ZnIn2S4 composited with 2D graphene: Designed synthesis and superior photocatalytic performance. Dalton Trans. 2017, 46, 2770–2777. [Google Scholar] [CrossRef] [PubMed]
  46. Li, L.; Che, H.; Yan, Y.; Liu, C.; Dong, H. Z-scheme AgVO3/ZnIn2S4 photocatalysts: “One Stone and Two Birds” strategy to solve photocorrosion and improve the photocatalytic activity and stability. Chem. Eng. J. 2020, 398, 125523. [Google Scholar] [CrossRef]
  47. Fan, W.-J.; Zhou, Z.-F.; Xu, W.-B.; Shi, Z.-F.; Ren, F.-M.; Ma, H.-H.; Huang, S.-W. Preparation of ZnIn2S4/fluoropolymer fiber composites and its photocatalytic H2 evolution from splitting of water using Xe lamp irradiation. Int. J. Hydrogen Energy 2010, 35, 6525–6530. [Google Scholar] [CrossRef]
  48. Shen, S.; Chen, X.; Ren, F.; Kronawitter, C.X.; Mao, S.S.; Guo, L. Solar light-driven photocatalytic hydrogen evolution over ZnIn2S4 loaded with transition-metal sulfides. Nanoscale Res. Lett. 2011, 6, 290. [Google Scholar] [CrossRef] [Green Version]
  49. An, H.; Li, M.; Liu, R.; Gao, Z.; Yin, Z. Design of AgxAu1−x alloy/ZnIn2S4 system with tunable spectral response and Schottky barrier height for visible-light-driven hydrogen evolution. Chem. Eng. J. 2020, 382, 122953. [Google Scholar] [CrossRef]
  50. Hu, J.; Chen, C.; Zheng, Y.; Zhang, G.; Guo, C.; Li, C.M. Spatially Separating Redox Centers on Z-Scheme ZnIn2S4/BiVO4 Hierarchical Heterostructure for Highly Efficient Photocatalytic Hydrogen Evolution. Small 2020, 16, 2002988. [Google Scholar] [CrossRef]
  51. Zhou, M.; Wang, S.; Yang, P.; Luo, Z.; Yuan, R.; Asiri, A.M.; Wakeel, M.; Wang, X. Layered Heterostructures of Ultrathin Polymeric Carbon Nitride and ZnIn2S4 Nanosheets for Photocatalytic CO2 Reduction. Chem. Eur. J. 2018, 24, 18529–18534. [Google Scholar] [CrossRef]
  52. Gao, W.; Wang, L.; Gao, C.; Liu, J.; Yang, Y.; Yang, L.; Shen, Q.; Wu, C.; Zhou, Y.; Zou, Z. Exquisite design of porous carbon microtubule-scaffolding hierarchical In2O3-ZnIn2S= heterostructures toward efficient photocatalytic conversion of CO2 into CO. Nanoscale 2020, 12, 14676–14681. [Google Scholar] [CrossRef] [PubMed]
  53. Zhu, Z.; Li, X.; Qu, Y.; Zhou, F.; Wang, Z.; Wang, W.; Zhao, C.; Wang, H.; Li, L.; Yao, Y.; et al. A hierarchical heterostructure of CdS QDs confined on 3D ZnIn2S4 with boosted charge transfer for photocatalytic CO2 reduction. NanoResearch 2021, 14, 81–90. [Google Scholar] [CrossRef]
  54. Wang, X.; Chen, J.; Li, Q.; Li, L.; Zhuang, Z.; Chen, F.-F.; Yu, Y. Light-Driven Syngas Production over Defective ZnIn2S4 Nanosheets. Chem. Eur. J. 2020. [CrossRef]
Figure 1. (a) X-ray diffractogram of undoped ZnIn2S4, prepared hydrothermally with thiourea and water. (b) Determination of the bandgap, from the Kubelka-Munk transform of the diffuse reflectance spectra, using a Tauc plot. (c) Comparison of the spectral response in the photocatalytic degradation of aqueous HCOOH, using the first-order rate constant k, with the spectrum of ZnIn2S4 (adapted with permission from Ref. [26]. Copyright 2020 Elsevier Ltd.).
Figure 1. (a) X-ray diffractogram of undoped ZnIn2S4, prepared hydrothermally with thiourea and water. (b) Determination of the bandgap, from the Kubelka-Munk transform of the diffuse reflectance spectra, using a Tauc plot. (c) Comparison of the spectral response in the photocatalytic degradation of aqueous HCOOH, using the first-order rate constant k, with the spectrum of ZnIn2S4 (adapted with permission from Ref. [26]. Copyright 2020 Elsevier Ltd.).
Photochem 01 00001 g001
Figure 2. X-ray diffractogram of undoped and V-substituted ZnIn2S4, using Na2S and EG-water as solvent.
Figure 2. X-ray diffractogram of undoped and V-substituted ZnIn2S4, using Na2S and EG-water as solvent.
Photochem 01 00001 g002
Figure 3. Kubelka-Munk transform, plotted versus photon energy, of V-substituted ZnIn2S4 made solvothermally.
Figure 3. Kubelka-Munk transform, plotted versus photon energy, of V-substituted ZnIn2S4 made solvothermally.
Photochem 01 00001 g003
Figure 4. EPR spectra of V-substituted ZnIn2S4.
Figure 4. EPR spectra of V-substituted ZnIn2S4.
Photochem 01 00001 g004
Figure 5. Density of states plot showing the results of the DFT calculation carried out for V-substituted ZnIn2S4. Total DOS in black; colored traces correspond to the partial DOS centered on each element. The vertical dashed line marks the Fermi energy.
Figure 5. Density of states plot showing the results of the DFT calculation carried out for V-substituted ZnIn2S4. Total DOS in black; colored traces correspond to the partial DOS centered on each element. The vertical dashed line marks the Fermi energy.
Photochem 01 00001 g005
Figure 6. (a) Decrease in the HCOOH concentration for the V-substituted ZnIn2S4 photocatalyst using different filters (the numbers indicate the wavelength utilized). (b) Comparison of the HCOOH photodegradation rate constants found for undoped and V-substituted ZnIn2S4, including also the absorption spectra of each material.
Figure 6. (a) Decrease in the HCOOH concentration for the V-substituted ZnIn2S4 photocatalyst using different filters (the numbers indicate the wavelength utilized). (b) Comparison of the HCOOH photodegradation rate constants found for undoped and V-substituted ZnIn2S4, including also the absorption spectra of each material.
Photochem 01 00001 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lucena, R.; Conesa, J.C. V-Substituted ZnIn2S4: A (Visible+NIR) Light-Active Photocatalyst. Photochem 2021, 1, 1-9. https://0-doi-org.brum.beds.ac.uk/10.3390/photochem1010001

AMA Style

Lucena R, Conesa JC. V-Substituted ZnIn2S4: A (Visible+NIR) Light-Active Photocatalyst. Photochem. 2021; 1(1):1-9. https://0-doi-org.brum.beds.ac.uk/10.3390/photochem1010001

Chicago/Turabian Style

Lucena, Raquel, and José C. Conesa. 2021. "V-Substituted ZnIn2S4: A (Visible+NIR) Light-Active Photocatalyst" Photochem 1, no. 1: 1-9. https://0-doi-org.brum.beds.ac.uk/10.3390/photochem1010001

Article Metrics

Back to TopTop