Next Article in Journal
Design of Amine-Modified Zr–Mg Mixed Oxide Aerogel Nanoarchitectonics with Dual Lewis Acidic and Basic Sites for CO2/Propylene Oxide Cycloaddition Reactions
Next Article in Special Issue
The Role of Nitrogen-Doped TiO2 Supported by Platinum Catalyst Synthesized via Various Mode Preparations for Photocatalytic Enhancement
Previous Article in Journal
Graphene-Based Composites with Silver Nanowires for Electronic Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Simple Fabrication of Sb2S3/TiO2 Photo-Anode with Long Wavelength Visible Light Absorption for Efficient Photoelectrochemical Water Oxidation

1
School of Chemistry and Chemical Engineering, Hefei University of Technology, Hefei 230009, China
2
School of Material Science and Engineering, North Minzu University, Yinchuan 750021, China
3
Key Laboratory of Polymer Materials and Manufacturing Technology, North Minzu University, Yinchuan 750021, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(19), 3444; https://doi.org/10.3390/nano12193444
Submission received: 31 August 2022 / Revised: 26 September 2022 / Accepted: 28 September 2022 / Published: 1 October 2022

Abstract

:
An Sb2S3-sensitized TiO2 (Sb2S3/TiO2) photo-anode (PA) exhibiting a high photo-electrochemical (PEC) performance in water oxidation has been successfully prepared by a simple chemical bath deposition (CBD) technique. Herein, the Raman spectra and XPS spectrum of Sb2S3/TiO2 confirmed the formation of Sb2S3 on the TiO2 coatings. The Sb2S3/TiO2 photo-anode significantly shifted the absorption edge from 395 nm (3.10 eV) to 650 nm (1.90 eV). Furthermore, the Sb2S3/TiO2 photo-anode generated a photo-anodic current under visible light irradiation below 650 nm due to the photo-electrochemical action compared with the TiO2 photo-anode at 390 nm. The incident photon-to-current conversion efficiency (IPCE = 7.7%) at 400 nm and −0.3 V vs. Ag/AgCl was 37 times higher than that (0.21%) of the TiO2 photo-anodes due to the low recombination rate and acceleration of the carriers of Sb2S3/TiO2. Moreover, the photo-anodic current and photostability of the Sb2S3/TiO2 photo-anodes improved via adding the Co2+ ions to the electrolyte solution during photo-electrocatalysis.

1. Introduction

PEC water splitting using semiconductors has been expected to be one of the ideal ways to realize the conversion of solar radiative energy into hydrogen energy [1,2,3]. In the PEC system, water oxidation is considered the kinetic control step of the whole water splitting process due to the sluggish kinetics of PEC water oxidation through photo-anodes (pAs). Therefore, developing a robust, stable, and economical photo-anode is critical for highly efficient solar water-splitting.
PEC water oxidation using n-type TiO2 semiconductor pAs has gained considerable attention since 1972 due to their chemical and electrochemical stability in oxidation conditions, easy preparation, and noticeable incident light-to-current conversion efficiencies [4]. However, TiO2 pAs-based PEC water oxidation is limited because of two serious drawbacks. First, TiO2 has a wide bandgap (3.0–3.2 eV) that solely responds to the ultraviolet fraction of the solar spectrum (accounts for just 5% of solar irradiation) [5,6]. Second, a high recombination probability of electron-hole pairs leads to decreased incident light-to-current conversion efficiency.
Considering these disadvantages, developing highly active TiO2 pAs is integral to absorbing solar irradiations from ultraviolet to visible light. So far, many efforts have been devoted to noble metal loading [7,8,9,10], semiconductor recombination [11,12,13,14,15], doping [16,17,18,19] and sensitization [13,19,20,21,22,23,24,25,26,27,28,29,30] to enlarge the spectral response range due to improve its PEC performance for efficient water oxidation and to promote electron-hole pair separation. Among these, the sensitization of TiO2 utilizing chalcogenide-based (Sb2S3) catalysts has received significant attention [19,20,21,31,32]. Sb2S3 possesses a suitable bandgap (1.6–1.9 eV) and a high absorption coefficient (105 cm−1) in the visible region, which benefits from absorbing the whole visible and near-infrared range of the solar spectrum. Moreover, Sb2S3 has low toxicity and moisture, making it one of the most potential sensitizers to investigate the sensitization application. Due to these phenomena, Sb2S3/TiO2 pAs were fabricated by a sensitization process with Sb2S3 as a sensitizer, exhibiting enhanced PEC water oxidation properties.
Meanwhile, different strategies have been noted for preparing the Sb2S3/TiO2 pAs, such as atomic layer deposition (ALD), chemical bath deposition (CBD), successive ionic layer adsorption and reaction, etc. However, the costs associated with these physical methods are comprehensive. CBD has been selected in this article due to its low cost and simple procedure [33,34,35,36]. Moreover, several synthetic strategies need subsequent complex steps involving a somewhat complicated procedure for reproducible Sb2S3/TiO2 and developing a simple and reproducible fabrication technique for Sb2S3/TiO2 pAs synthesis. Herein, we describe a simple, low-cost, straightforward strategy for fabricating Sb2S3/TiO2 pAs using a CBD technique. Sb2S3 were coated and deposited on the TiO2 coated previously squeezed onto the ITO substrate utilizing a commercially available TiO2 paste (PST-18NR). The Sb2S3/TiO2 pAs show a higher PEC execution in water oxidation than TiO2 pAs due to the charge separation and transportation acceleration phenomena.

2. Materials and Methods

2.1. Materials

Antimony (III) chloride (SbCl3), Marpolose (60MP-50), HCl, and Polyethylene glycol (PEG, molecular weight = 2000) were purchased from Aladdin’s Reagent (Shanghai Aladdin Bio-Chem Technology Co., Ltd., Shanghai, China) and TiO2 paste (18 NR-T) was purchased from Macklin Reagent (Shanghai Macklin Biochemical Co., Ltd. Shanghai, China). An Indium tin oxide (ITO)-coated glass substrate was obtained from Dalian HeptaChroma Co., Ltd., (Dalian, China); Millipore water (DIRECT-Q 3UV, Merck Ltd. Shanghai, China) was used for all the experiments. All other chemicals were of analytical grade and were used as received unless mentioned otherwise.

2.2. Fabrication of TiO2/ITO Electrode

In a typical procedure, the TiO2 pAs was synthesized as follows. Before coating, the FTO-glass substrate (1.0 cm−2 area) was cleaned up using deionized water, acetone, and ethanol by sonication for 15 min. Then TiO2 paste was squeezed over an ITO-glass substrate by a doctor-blade coater and dried at 80 °C for 15 min. After repeating the procedure twice, the electrode was calcined at 550 °C under an air atmosphere for 2 h to create a TiO2 film-coated ITO electrode (TiO2/ITO).

2.3. Fabrication of Sb2S3/TiO2/ITO Electrode

SbCl3 (787.1 mg, 3.46 mmol) was solved in acetone (3.03 mL); after the Na2S2O3 (30.3 mL, 1 mol/L) solution was added, the Sb2S3 suspension was obtained. Then, the TiO2/ITO was immersed into the resulting Sb2S3 suspension for 24 h in the refrigerator at 10 °C. The electrode was washed thoroughly with water and air-dried to obtain the Sb2S3/TiO2/ITO electrode.

2.4. Measurement

A Shimadzu XRD-6000 diffractometer (Shimadzu International Trade (Shanghai) Co., Ltd., Shanghai, China) with Cu Kα (λ = 0.15406 nm) radiation was utilized to determine the structure of samples and to obtain the X-ray diffraction (XRD) patterns. Raman microspectroscopic apparatus (Horiba-Jobin-Yvon LabRAM HR, Paris, France) was used to obtain the Raman spectra. At room temperature, a silicon reference sample was calibrated with a theoretical position of 520.7 cm−1. The spectrophotometer (Shimadzu UV-2700, Shimadzu International Trade (Shanghai) Co., Ltd., Shanghai, China) was used to test the UV–visible diffuse reflectance spectra (DRS). An electron probe microanalysis (JED-2300, JEOL, Tokyo, Japan) operating at 10 kV (acceleration voltage) was used to obtain the energy-dispersive X-ray spectroscopic (EDS) data. The XPS spectra were performed using a Thermo Fisher Scientific ESCALAB Xi+ instrument (Thermo Fisher Scientific (China) Co., Ltd., Shanghai, China) and calibrated in reference to C 1 s peak fixed at 284.2 eV.
All PEC measurements were examined using an electrochemical workstation in a single-compartment PEC cell (Shanghai Chenhua Instrument Co., Ltd. Shanghai, China, CHI760E). The prepared electrode was used as the working electrode. An Ag/AgCl electrode was used as the reference electrode, and a Pt wire as the counter electrode. An aqueous 0.1 M phosphate solution (pH 6.0) was used as an electrolyte in the cell compartments and Ar-gas environments. The cyclic voltammogram (CV) was recorded at a scan rate of 50 mV s−1 at room temperature. The output of light intensity was calibrated as 100 mW cm−2 using a spectroradiometer (USR-40, Ushio Shanghai Inc., Shanghai, China). Light (λ > 500 nm, 100 mW/cm2) was irradiated from an irradiation source of a 500 W xenon lamp (Optical Module X, Ushio Shanghai Inc., Shanghai, China) with a UV-cut filter (λ > 500), and a 0.2 M CuSO4 was used as a liquid filter for cutting the heat ray. The linear sweep voltammograms (LSV) were measured at a scan rate of 5 mV s−1. A monochromic light with a 10 nm bandwidth was used with a 500 W xenon lamp using a monochromator for incident photon-to-current conversion efficiency (IPCE) measurements.
Photo-electrocatalysis was executed under the potentiostatic conditions at −0.3 V at 25 °C with the illumination of light (λ > 500 nm, 100 mW/cm2) for 1 h. 0.1 mM Co(NO3)2·6H2O was dissolved in the electrolyte solution to test the effect of Co2+ ions on photo-electrocatalysis.

3. Results

3.1. Optimization of Immersion Time

The UV–vis absorption spectrum of Sb2S3/TiO2/ITO electrodes immersing at different times is shown in Figure 1a. No response was observed for the TiO2/ITO electrode in the visible region. On the contrary, a dramatic response was observed for Sb2S3/TiO2/ITO electrodes, which are extended over the whole visible region (450–800 nm), demonstrating that the hole-electron pairs can be separated efficiently after sensitizing with Sb2S3. Figure 1b shows the plots of the immersion times with the absorbance at 460 nm for the Sb2S3/TiO2/ITO electrodes. The absorbance exhibited an increasing tendency with immersion time increased; the highest absorbance was observed when the immersion time was 24 h. After that, the absorbance decreased when the immersion time increased, suggesting that the optimum immersion condition was 24 h, which was employed in the following experiments.

3.2. Characterization Structure of Different Samples

The crystalline structures of the ITO, TiO2 and Sb2S3/TiO2 electrodes were characterized by XRD diffraction (Figure 2). Figure 2Ia shows the peaks of the ITO substrate. The diffracted peaks of TiO2/ITO (Figure 2Ib) and Sb2S3/TiO2/ITO electrodes (Figure 2Ic) were clearly detected at 2θ = 25.1°, 37.7°, 47.8°, 53.9° and 53.9°, respectively (JSPDF number: 89–4921) [5]. No peaks in terms of Sb2S3 were observed for both electrodes. However, it should be noted that two slightly broadened peaks were detected around 2θ = 12 and 55° for the Sb2S3/TiO2/FTO electrode. Raman spectroscopy was carried out for different powders scratching from the electrodes. Both TiO2 (Figure 2IIb) and Sb2S3/TiO2 (Figure 2IIc) samples exhibited the typical peaks located at 147, 199, 400, 517, and 638 cm−1, which correspond to the vibrational modes of pure anatase TiO2 [20,37]. Figure 2IIa shows the peak of 289 cm−1 for Sb2S3, which could be assigned to the Sb-S stretching modes [20,31,38,39,40]. Moreover, this broad peak was observed in Sb2S3/TiO2, while it was not detected for pure TiO2 (Figure 2IIb), confirming the Sb2S3/TiO2 composite formation.
The valence state of Sb2S3/TiO2 was investigated utilizing XPS. As shown in the survey scan in Figure 3a, Sb, Ti, S, and O elements were verified. In the high-resolution Ti 2p spectrum (Figure 3b), two peaks at 458.8 and 464.5 eV were observed, assigned to the binding energies of Ti 2p3/2 and Ti 2p1/2 of the TiO2 lattice, respectively [16,17,37]. The XPS spectrum of Sb2S3/TiO2 exhibited two peaks in the S 2p region at 161.4 and 162.4 eV correspond to S 2p1/2 and S 2p3/2 of divalent S in Figure 3c [16,19,41]. The peaks at 529.3 and 538.7 eV in the Sb 3d region for Sb2S3/TiO2, respectively, are assigned the Sb3+ oxidation state of Sb and are consistent with the earlier-reported values of Sb2S3 [19,42,43,44]. The relatively weak peak at 531.7 can be assigned to the formation of the Sb-O band at the surface, possibly associated with the oxygen atom of TiO2.
The FE-SEM observed the morphologies of the TiO2 and Sb2S3/TiO2 electrodes. Figure 4I(a) exhibited the top view of the TiO2 electrode at low magnification, in which a continuous uniform and smoothed surface were formed. This would be beneficial to improve the PEC performance of the Sb2S3/TiO2 electrode due to accelerating the water oxidation reaction at the surface of the electrolyte solution. The FE-SEM images at high magnification (Figure 4I(b)) showed that the average particle size of TiO2 was 10 nm. The surface FE-SEM image of the Sb2S3/TiO2 electrode at low magnification is shown in Figure 4I(c), where the larger Sb2S3 particles were deposited on the TiO2 surface. After that, the 60–100 nm Sb2S3 particles were clearly observed in Figure 4I(d). EDX analyses of the Sb2S3/TiO2 electrode were taken to confirm the Sb2S3 deposition and complete the chemical composition information. The elemental maps of the EDX for the Sb2S3/TiO2 electrode are shown in Figure 4II. The Ti and O mapping signals (Figure 4II(c,d)) and S and Sb (Figure 4II(e,f)) were detected. The distribution of Sb2S3 onto the TiO2 in the Sb2S3/TiO2 electrode appeared clearly.

3.3. The Optical Properties of Sb2S3/TiO2

The diffuse reflection spectrum of TiO2 and Sb2S3/TiO2 are presented in Figure 5a. It was observed that the Sb2S3/TiO2 absorb the visible light below 650 nm, exhibiting significant redshift in a broad wavelength region compared to TiO2 (395 nm). Tauc plots based on UV–vis DRS are exhibited in Figure 5b. The TiO2 showed a single absorption edge of 3.20 eV, corresponding to the previous report relating to anatase TiO2 [5]. The Sb2S3/TiO2 provided two different slops due to the appearance of the new absorption edges, one of the band energies was 1.94 eV, and another was 2.37 eV. The photo-response of the Sb2S3/TiO2 was significantly improved in the visible region. This is attributed to the existence of Sb2S3, which significantly promotes the separation of the electron-hole pairs.

3.4. Photoelectrocatalytic Properties

The action spectra of IPCE were recorded at −0.3 V by utilizing TiO2 and Sb2S3/TiO2 electrodes, as shown in Figure 5c. The IPCE value at 400 nm for Sb2S3/TiO2 electrode (7.7%) is 37 times higher than that for TiO2 electrode (0.21%). Further, the onset wavelength for photo-current generation on Sb2S3/TiO2 electrode was 650 nm (1.9 eV), which is noticeably longer than the TiO2 electrode (400 nm, 3.1 eV). The IPCE action spectra consisted of the UV-DRS data for Sb2S3/TiO2 electrode, indicating that the pC was generated based on the bandgap excitation through collateral excitation from Sb 3d orbital to CB.
Several PEC examinations were performed to reveal the intrinsic PEC performances of TiO2 and Sb2S3/TiO2 electrodes towards water oxidation, as shown in Figure 6. Figure 6a showed the CV of the TiO2 (black) and Sb2S3/TiO2 (red) electrodes in a 0.1M phosphate solution (pH = 6.0). Both electrodes did not have photo-anodic currents in the dark condition (dashed line). Contrary to the TiO2 electrode, which showed a 9.1 A/cm2 at 0.2 V, the Sb2S3/TiO2 electrode generated a stable photo-anodic current of 42.3 A/cm2 based on water oxidation over 0.6 V, showing 4.6 times higher value than that of the TiO2 electrode. Such remarkable enhancement of pC density for the Sb2S3/TiO2 electrode can be ascribed to the improvement of light absorption in the visible region. The LSV of two electrodes shown in Figure 6b was taken under chopped/simulated chopped visible light irradiation. Interestingly, it can be observed that the photo-responses of anodic and cathodic currents for the Sb2S3/TiO2 electrode (red) were clearly observed above or below −0.5 V, although hardly any pAs current was observed for the TiO2 electrode (black, insert).
The results suggest that the photo-current direction of the Sb2S3/TiO2 electrode can be modified according to the applied potential. When the potential is below −0.5 V, the charges separation based on bandgap excitation has caused upon irradiation with visible light, then the electrons from ITO are rapidly injected into the valence band of Sb2S3 through the conduction band of TiO2. Afterward, the electron acceptor in the conduction band of Sb2S3 is reduced by the electrons to generate a cathodic current. On the other condition of potential above −0.5 V, Sb2S3 absorbs photons to generate electron-hole pairs, the electrons are injected into the electron transport layer (TiO2) and collected by ITO, and the holes oxidize the water at the surface of Sb2S3 to generate an anodic current. As the light was turned on and off, instantaneous positive and negative transient photocurrents were observed under chopper illumination at −0.3 V (Figure 6c).
A negligible photo-current for the TiO2 electrode (black) was observed. However, the photo-current of 30 A/cm2 for the Sb2S3/TiO2 electrode (red) was observed. It is indicated that the charge carriers are very efficiently separated and limited to recombine after Sb2S3 sensitization. The photo-electrocatalysis was conducted under visible light irradiation (λ > 500 nm, 100 mW/cm2) at −0.3 V in a 0.1M phosphate solution (pH = 6.0) for 1 h (Figure 6d). The I-t profile of the Sb2S3/TiO2 electrode exhibited a photo-current of 30 A/cm2 over the period, while for the TiO2 electrode, almost no photo-current was observed. The PEC performance of hematite and WO3 photo-anodes for water oxidation can be enhanced by adding Co2+ ions into the electrolyte solution to act as catalysts at the electrode interface with the electrolyte solution without any deposition [45,46].
The influence of adding Co2+ ions on the PEC water oxidation for the Sb2S3/TiO2 electrode was measured to prove this assumption. The performance of the Sb2S3/TiO2 electrode for water oxidation was improved by the presence of 0.1 mm Co2+ ions in the phosphate buffer solution. It should be noted that the stability of the photo-current for the Sb2S3/TiO2 electrode by the presence of Co2+ ions was remarkably improved, which consequently generated a 2.7 times higher photo-current of 80 A/cm2 compared to the absence one. Such a lower decay rate (20%) of photo-current for the Sb2S3/TiO2 electrode by the addition of Co2+ ions demonstrates that the stability is ascribed to the acceleration of the water oxidation reaction between the Sb2S3/TiO2 surface and the electrolyte solution interface, which significantly reduce the recombination rate of photo-generated carriers.

4. Results and Discussion

Two main disadvantages are associated with high recombination of electron-hole pairs and a wide bandgap for TiO2. Many efforts have been devoted to noble metal loading, semiconductor recombination, doping and sensitization to enlarge the spectral response range due to improving its PEC performance for efficient water oxidation and promoting electron-hole pairs separation. Sb2S3 possesses a suitable bandgap and a high absorption coefficient in the visible region, which benefits from absorbing the whole visible and near-infrared range of the solar spectrum. Thus, the Sb2S3/TiO2 were fabricated by a sensitization process with Sb2S3 as a sensitizer, exhibiting enhanced PEC water oxidation properties. Herein, a simple, low-cost and straightforward strategy for fabricating Sb2S3/TiO2 using a CBD technique was reported in this paper.

5. Conclusions

An Sb2S3/TiO2 photo-anode was fabricated utilizing a simple and low-cost CBD strategy. XPS and elemental mapping detected Sb and S signals and Ti and O, demonstrating the formation of Sb2S3 and TiO2, respectively. The Sb2S3/TiO2 photo-anode can utilize visible light below 650 nm for PEC water oxidation, in contrast to utilization below 395 nm for the TiO2 photo-anode. The IPCE of 7.7% at 400 nm for Sb2S3/TiO2 photo-anode was higher than that of 0.21% for TiO2 photo-anode by 37 times. The Sb2S3/TiO2 photo-anode can generate a photo-anodic current and a photo-cathodic current in contrast to generating a photo-anodic current for the TiO2 photo-anode. The photo-electrocatalytic performance of Sb2S3/TiO2 photo-anode would be further improved by adding Co2+ ions in the electrolyte solution due to low recombination rates of charge carriers and facilitated electron transfers.

Author Contributions

Conceptualization, D.L. and F.H.; methodology, F.H.; investigation and data curation, M.M.A. and F.H.; formal analysis, S.M. and F.H.; supervision, D.L.; writing—original draft preparation, F.H.; writing—review and editing, D.L. and Z.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This project was supported by the Natural Science Foundation of Ningxia Province, grant number 2021AAC03170.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

F.H. thanks Hefei University of Technology for providing with a Ph.D. D.L. thanks the Natural Science Foundation of Ningxia Province (Grant No. 2021AAC03170).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Grätzel, M. Photoelectrochemical cells. Nature 2001, 414, 338–344. [Google Scholar] [CrossRef] [PubMed]
  2. Kment, S.; Riboni, F.; Pausova, S.; Wang, L.; Wang, L.; Han, H.; Hubicka, Z.; Krysa, J.; Schmuki, P.; Zboril, R. Photoanodes based on TiO2 and α-Fe2O3 for solar water splitting—superior role of 1D nanoarchitectures and of combined heterostructures. Chem. Soc. Rev. 2017, 46, 3716–3769. [Google Scholar] [CrossRef] [PubMed]
  3. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef] [PubMed]
  4. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  5. Chandra, D.; Tsuriya, R.; Sato, T.; Takama, D.; Abe, N.; Kajita, M.; Li, D.; Togashi, T.; Kurihara, M.; Saito, K.; et al. Characterization of Interfacial Charge-Transfer Photoexcitation of Polychromium-Oxo-Electrodeposited TiO2 as an Earth-Abundant Photoanode for Water Oxidation Driven by Visible Light. ChemPlusChem 2016, 81, 1116–1122. [Google Scholar] [CrossRef]
  6. El Rouby, W.M.A.; Antuch, M.; You, S.; Millet, P. Surface sensitization of TiO2 nanorod mats by electrodeposition of ZIF-67 for water photo-oxidation. Electrochim. Acta 2020, 339, 135882–135897. [Google Scholar] [CrossRef]
  7. Liu, R.; Fang, S.Y.; Tsai, K.C.; Yang, W.D. Enhancing hydrogen evolution of water splitting under solar spectra using Au/TiO2 heterojunction photocatalysts. Int. J. Hydrogen Energy 2021, 46, 28462–28473. [Google Scholar] [CrossRef]
  8. Park, H.; Choi, W.; Hoffmann, M.R. Effects of the preparation method of the ternary CdS/TiO2/Pt hybrid photocatalysts on visible light-induced hydrogen production. J. Mater. Chem. 2008, 18, 2379–2385. [Google Scholar] [CrossRef] [Green Version]
  9. Gupta, B.; Melvin, A.A.; Matthews, T.; Dash, S.; Tyagi, A.K. TiO2 modification by gold (Au) for photocatalytic hydrogen (H2) production. Renew. Sust. Energ. Rev. 2016, 58, 1366–1375. [Google Scholar] [CrossRef]
  10. Patra, K.K.; Gopinath, C.S. Harnessing Visible-Light and Limited Near-IR Photons through Plasmon Effect of Gold Nanorod with AgTiO2. J. Phys. Chem. C 2018, 122, 1206–1214. [Google Scholar] [CrossRef] [Green Version]
  11. Ibrahim, O.Y.; Gondal, A.M. Visible-light-driven photocatalytic performance of a Z-scheme based TiO2/WO3/g-C3N4 ternary heterojunctions. Mol. Catal. 2021, 505, 111494. [Google Scholar] [CrossRef]
  12. Ge, M.; Cai, J.; Cao, C.; Huang, J.; Zhang, X.; Shen, J.; Wang, S.; Zhang, S.; Zhang, K.Q.; Lai, Y.; et al. A review of TiO2 nanostructured catalysts for sustainable H2 generation. Int. J. Hydrogen Energy 2017, 42, 8418–8449. [Google Scholar] [CrossRef]
  13. Chen, S.; Li, C.; Hou, Z. A novel in situ synthesis of TiO2/CdS heterojunction for improving photoelectrochemical water splitting. Int. J. Hydrogen Energy 2019, 44, 25473–25485. [Google Scholar] [CrossRef]
  14. Ding, Q.; Gou, L.; Wei, D.; Xu, D.; Fan, W.; Shi, W. Metal-organic framework derived Co3O4/TiO2 heterostructure nanoarrays for promote photoelectrochemical water splitting. Int. J. Hydrogen Energy 2021, 46, 24965–24976. [Google Scholar] [CrossRef]
  15. Ismail, M.; Chebaane, M.M.; Bousselmi, L.; Zahraa, O.; Olivier, C.; Toupance, T. Photoelectrochemical properties of WO3 modified anatase TiO2 photoanodes and application for dye-sensitized solar cells. Surf. Interfaces 2021, 27, 101543–101551. [Google Scholar] [CrossRef]
  16. Le, T.T.T.; Tran, D.T.; Danh, T.H. Remarkable enhancement of visible light driven photocatalytic performance of TiO2 by simultaneously doping with C, N, and S. Chem. Phys. 2021, 545, 111144–111153. [Google Scholar] [CrossRef]
  17. Mollavali, M.; Rohani, S.; Elahifard, M.; Behjatmanesh-Ardakani, R.; Nourany, M. Band gap reduction of (Mo+N) co-doped TiO2 nanotube arrays with a significant enhancement in visible light photo-conversion: A combination of experimental and theoretical study. Int. J. Hydrogen Energy 2021, 46, 21475–21498. [Google Scholar] [CrossRef]
  18. Imran, M.; Saeed, Z.; Pervaiz, M.; Mehmood, K.; Ejaz, R.; Younas, U.; Nadeem, H.A.; Hussain, S. Enhanced visible light photocatalytic activity of TiO2 co-doped with Fe, Co, and S for degradation of Cango red. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2021, 255, 119644–119655. [Google Scholar] [CrossRef]
  19. Cerdan-Pasaran, A.; López-Luke, T.; Mathewa, X.; Mathews, N.R. Effect of cobalt doping on the device properties of Sb2S3-sensitized TiO2 solar cells. Sol. Energy 2019, 183, 697–703. [Google Scholar] [CrossRef]
  20. Elbakkay, M.H.; El Rouby, W.M.A.; Mariño-López, A.; Sousa-Castillo, A.; Salgueiriño, V.; El-Dek, S.I.; Farghali, A.A.; Correa Duarte, M.A.; Millet, P. One-pot synthesis of TiO2/Sb2S3/RGO complex multicomponent heterostructures for highly enhanced photoelectrochemical water splitting. Int. J. Hydrogen Energy 2021, 46, 31216–31227. [Google Scholar]
  21. Ben Naceur, J.; Ouertani, R.; Chakhari, W.; Chtourou, R. Photo-electrochemical properties of Sb2S3/TiO2 heterostructures integrally synthesis by hydrothermal method. J. Mater. Sci. Mater. Electron. 2019, 30, 5631–5639. [Google Scholar] [CrossRef]
  22. Ortiz-Bustos, J.; Hierro, I.; de Sánchez-Ruiz, A.; García-Martínez, J.C.; Pérez, Y. Tuning of type-I and type-II mechanisms for visible light degradation in tris(styryl)benzene-sensitized TiO2 nanoparticles. Dyes Pigments 2021, 18, 108802–108816. [Google Scholar] [CrossRef]
  23. Yu, Y.-X.; Yu, Y.X.; Pan, L.; Son, M.K.; Mayer, M.; Zhang, W.D.; Hagfeldt, A.; Luo, J.; Grätzel, M. Solution Processed Cu2S Photocathodes for Photoelectrochemical Water Splitting. ACS Energy Lett. 2018, 3, 760–766. [Google Scholar] [CrossRef]
  24. Chandra, M.; Bhunia, K.; Pradhan, D. Controlled Synthesis of CuS/TiO2 Heterostructured Nanocomposites for Enhanced Photocatalytic Hydrogen Generation through Water Splitting. Inorg. Chem. 2018, 57, 4524–4533. [Google Scholar] [CrossRef]
  25. Hou, Y.; Li, X.; Zhao, Q.; Quan, X.; Chen, G. Fabrication of Cu2O/TiO2 nanotube heterojunction arrays and investigation of its photoelectrochemical behavior. App. Phys. Lett. 2009, 95, 093108. [Google Scholar] [CrossRef]
  26. Zaban, A.; Mićić, O.; Gregg, B.A.; Nozik, A. Photosensitization of Nanoporous TiO2 Electrodes with InP Quantum Dots. Langmuir 1998, 14, 3153–3156. [Google Scholar] [CrossRef]
  27. Li, Q.; Zheng, M.; Zhang, B.; Zhu, C.; Wang, F.; Song, J.; Zhong, M.; Ma, L.; Shen, W. InP nanopore arrays for photoelectrochemical hydrogen generation. Nanotechnology 2016, 27, 075704–075715. [Google Scholar] [CrossRef] [Green Version]
  28. Wang, T.; Gong, J. Single-Crystal Semiconductors with Narrow Band Gaps for Solar Water Splitting. Angew. Chem. Int. Edit. 2015, 54, 10718–10732. [Google Scholar] [CrossRef]
  29. Luo, Z.; Wang, T.; Zhang, J.; Li, C.; Li, H.; Gong, J. Dendritic Hematite Nanoarray Photoanode Modified with a Conformal Titanium Dioxide Interlayer for Effective Charge Collection. Angew. Chem. Int. Edit. 2017, 56, 12878–12882. [Google Scholar] [CrossRef]
  30. Yao, H.; Fu, W.; Yang, H.; Ma, J.; Sun, M.; Chen, Y.; Zhang, W.; Wu, D.; Lv, P.; Li, M. Vertical Growth of Two-Dimensional TiO2 Nanosheets Array Films and Enhanced Photoelectrochemical Properties Sensitized by CdS Quantum Dots. Electrochim. Acta 2014, 125, 258–265. [Google Scholar] [CrossRef]
  31. Sharma, V.; CDakshinamurthy, A.; Pandey, B.; Roy, S.C.; Sudakar, C. Highly efficient photoelectrochemical ZnO and TiO2 nanorod/Sb2S3 heterostructured photoanodes through one step thermolysis of Sb-MPA complex. Sol. Energy 2021, 225, 333–343. [Google Scholar] [CrossRef]
  32. Zhong, J.S.; Wang, Q.Y.; Liang, H.; Chen, D.Q.; Ji, Z.G. Improving the visible light photocatalytic activity of TiO2 nanotubes upon decoration with Sb2S3 microcrystalline. Mater. Charact. 2015, 109, 95–99. [Google Scholar] [CrossRef]
  33. Chang, J.A.; Im, S.H.; Lee, Y.H.; Kim, H.J.; Lim, C.S.; Heo, J.H.; Seok, S.I. Panchromatic photon-harvesting by hole-conducting materials in inorganic-organic heterojunction sensitized-solar cell through the formation of nanostructured electron channels. Nano Lett. 2012, 12, 1863–1867. [Google Scholar] [CrossRef] [PubMed]
  34. Maiti NIm, S.H.; Lima, C.-S.; Seok, S.I. A chemical precursor for depositing Sb2S3 onto mesoporous TiO2 layers in nonaqueous media and its application to solar cells. Dalton Trans. 2012, 41, 11569–11572. [Google Scholar] [CrossRef]
  35. Zhu, G.; Pan, L.; Xu, T.; Sun, Z. One-step synthesis of CdS sensitized TiO2 photoanodes for quantum dot-sensitized solar cells by microwave assisted chemical bath deposition method. Acs Appl. Mater. Inter. 2011, 3, 1472–1478. [Google Scholar] [CrossRef]
  36. Itzhaik, Y.; Itzhaik, Y.; Niitsoo, O.; Page, M.G.; Hodes, G. Sb2S3-Sensitized Nanoporous TiO2 Solar Cells. J. Phys. Chem. C 2013, 113, 4254–4256. [Google Scholar] [CrossRef]
  37. Sharavath, V.; Sarkar, S.; Ghosh, S. One-pot hydrothermal synthesis of TiO2/graphene nanocomposite with simultaneous nitrogen-doping for energy storage application. J. Electroanal. Chem. 2018, 829, 208–216. [Google Scholar] [CrossRef]
  38. Chen, L.; Zhu, W.; Han, Q.; Yang, X.; Lu, L.; Wang, X. Preparation of rod-like Sb2S3 dendrites processed in conventional hydrothermal. Mater. Lett. 2009, 63, 1258–1261. [Google Scholar] [CrossRef]
  39. Parize, R.; Cossuet, T.; Chaix-Pluchery, O.; Roussel, H.; Appert, E.; Consonni, V. In situ analysis of the crystallization process of Sb2S3 thin films by Raman scattering and X-ray diffraction. Mater. Design 2017, 121, 1–10. [Google Scholar] [CrossRef]
  40. Kozytskiy, A.V.; Stroyuk, O.L.; Skoryk, M.A.; Dzhagan, V.M.; Kuchmiy, S.Y.; Zahn, D.R.T. Photochemical formation and photoelectrochemical properties of TiO2/Sb2S3 heterostructures. J. Photochem. Photobiol. A 2015, 303–304, 8–16. [Google Scholar] [CrossRef]
  41. Pal, M.; Mathews, N.R.; Mathew, X. Surfactant-mediated self-assembly of Sb2S3 nanorods during hydrothermal synthesis. J. Mater. Res. 2017, 32, 530–538. [Google Scholar] [CrossRef]
  42. Avilez Garcia, R.G.; Meza Avendaño, C.A.; Pal, M.; Paraguay Delgado, F.; Mathews, N.R. Antimony sulfide (Sb2S3) thin films by pulse electrodeposition: Effect of thermal treatment on structural, optical and electrical properties. Mat. Sci. Semicon. Proc. 2016, 44, 91–100. [Google Scholar] [CrossRef]
  43. Lou, W.; Chen, M.; Wang, X.; Liu, W. Novel Single-Source Precursors Approach to Prepare Highly Uniform Bi2S3 and Sb2S3 Nanorods via a Solvothermal Treatment. Chem. Mater. 2007, 19, 872–878. [Google Scholar] [CrossRef]
  44. Ota, J.; Srivastava, S.K. Tartaric Acid Assisted Growth of Sb2S3 Nanorods by a Simple Wet Chemical Method. Cryst. Growth Des. 2007, 7, 343–347. [Google Scholar] [CrossRef]
  45. Kay, A.; Cesar, I.; Grätzel, M. New Benchmark for Water Photooxidation by Nanostructured α-Fe2O3 Films. J. Am. Chem. Soc. 2006, 128, 15714–15721. [Google Scholar] [CrossRef]
  46. Li, D.; Takeuchi Ryouchi Chandra, D.; Saito, K.; Yui, T.; Yagi, M. Visible Light-Driven Water Oxidation on an In Situ N2-Intercalated WO3 Nanorod Photoanode Synthesized by a Dual-Functional Structure-Directing Agent. ChemSusChem 2018, 11, 1151–1156. [Google Scholar] [CrossRef]
Figure 1. (a) UV–Vis spectral change of the Sb2S3/TiO2 at difference time for dipping; (b) Plots of Absorbance vs. immersion time for Sb2S3/TiO2, respectively.
Figure 1. (a) UV–Vis spectral change of the Sb2S3/TiO2 at difference time for dipping; (b) Plots of Absorbance vs. immersion time for Sb2S3/TiO2, respectively.
Nanomaterials 12 03444 g001
Figure 2. (I) X-RD data; (a) ITO substrate; (b) TiO2, and (c) Sb2S3/TiO2 photo-anodes, respectively; (II) Raman spectroscopy; (a) Sb2S3, (b) TiO2, and (c) Sb2S3/TiO2 powders, respectively.
Figure 2. (I) X-RD data; (a) ITO substrate; (b) TiO2, and (c) Sb2S3/TiO2 photo-anodes, respectively; (II) Raman spectroscopy; (a) Sb2S3, (b) TiO2, and (c) Sb2S3/TiO2 powders, respectively.
Nanomaterials 12 03444 g002
Figure 3. XPS spectra of the Sb2S3/TiO2; (a) survey scan, (b) Ti 2p, (c) S 2p, and (d) Sb 3d, respectively. For the spectra in the Ti 2p, S 2p, and Sb 3d regions, the dashed lines are the deconvoluted spectrum bands, and the solid lines are the spectrum simulated by the deconvoluted bands.
Figure 3. XPS spectra of the Sb2S3/TiO2; (a) survey scan, (b) Ti 2p, (c) S 2p, and (d) Sb 3d, respectively. For the spectra in the Ti 2p, S 2p, and Sb 3d regions, the dashed lines are the deconvoluted spectrum bands, and the solid lines are the spectrum simulated by the deconvoluted bands.
Nanomaterials 12 03444 g003
Figure 4. (I) SEM images; (a,b) TiO2 and (c,d) Sb2S3/TiO2 electrode, respectively; (II) (a) SEM_EDX elements distribution mapping images of the Sb2S3/TiO2 electrode; (b) Ti, O S, Sb layered, (c) Ti, (d) O, (e) S, and (f) Sb, respectively.
Figure 4. (I) SEM images; (a,b) TiO2 and (c,d) Sb2S3/TiO2 electrode, respectively; (II) (a) SEM_EDX elements distribution mapping images of the Sb2S3/TiO2 electrode; (b) Ti, O S, Sb layered, (c) Ti, (d) O, (e) S, and (f) Sb, respectively.
Nanomaterials 12 03444 g004
Figure 5. (a) Diffuse reflection spectrum of TiO2 (black) and Sb2S3/TiO2 (Red); (b) Tauc plots of TiO2 (black) and Sb2S3/TiO2 (red). The dashed lines show tangent lines near the edges of the plots, respectively, (c) Action spectra of IPCE of theTiO2 (black) and Sb2S3/TiO2 (red) electrode in a 0.1 M phosphate buffer solution of pH 6.0 at −0.3 V vs. Ag/AgCl. The insets show the magnified spectra near the edges.
Figure 5. (a) Diffuse reflection spectrum of TiO2 (black) and Sb2S3/TiO2 (Red); (b) Tauc plots of TiO2 (black) and Sb2S3/TiO2 (red). The dashed lines show tangent lines near the edges of the plots, respectively, (c) Action spectra of IPCE of theTiO2 (black) and Sb2S3/TiO2 (red) electrode in a 0.1 M phosphate buffer solution of pH 6.0 at −0.3 V vs. Ag/AgCl. The insets show the magnified spectra near the edges.
Nanomaterials 12 03444 g005
Figure 6. (a) CVs measured in the dark (dash) and on irradiation (solid); the subfigures in Figure 6a is the magnification from -0.4 V to 0.2 V for TiO2.(b) LSV plots and; (c) Transient on/off chopped photo-current response of the Sb2S3/TiO2 (red) and TiO2 (the subfigures in Figure 6b) photo-anode in a 0.1 M phosphate buffer solution of pH 7.0 with visible light irradiation (λ > 500 nm, 100 mW/cm2) chopped. (d) photo-current density-time profiles for Sb2S3/TiO2 (red) and TiO2 photo-anode (black) in 0.1 m phosphate buffer solution of pH 6.0 during photo-electrocatalytic water oxidation at −0.3 V vs. Ag/AgCl. Sb2S3/TiO2 (blue) in the presence of 0.1 mm Co2+ ions in the electrolyte solution.
Figure 6. (a) CVs measured in the dark (dash) and on irradiation (solid); the subfigures in Figure 6a is the magnification from -0.4 V to 0.2 V for TiO2.(b) LSV plots and; (c) Transient on/off chopped photo-current response of the Sb2S3/TiO2 (red) and TiO2 (the subfigures in Figure 6b) photo-anode in a 0.1 M phosphate buffer solution of pH 7.0 with visible light irradiation (λ > 500 nm, 100 mW/cm2) chopped. (d) photo-current density-time profiles for Sb2S3/TiO2 (red) and TiO2 photo-anode (black) in 0.1 m phosphate buffer solution of pH 6.0 during photo-electrocatalytic water oxidation at −0.3 V vs. Ag/AgCl. Sb2S3/TiO2 (blue) in the presence of 0.1 mm Co2+ ions in the electrolyte solution.
Nanomaterials 12 03444 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

Han, F.; Ma, S.; Li, D.; Alam, M.M.; Yang, Z. A Simple Fabrication of Sb2S3/TiO2 Photo-Anode with Long Wavelength Visible Light Absorption for Efficient Photoelectrochemical Water Oxidation. Nanomaterials 2022, 12, 3444. https://doi.org/10.3390/nano12193444

AMA Style

Han F, Ma S, Li D, Alam MM, Yang Z. A Simple Fabrication of Sb2S3/TiO2 Photo-Anode with Long Wavelength Visible Light Absorption for Efficient Photoelectrochemical Water Oxidation. Nanomaterials. 2022; 12(19):3444. https://doi.org/10.3390/nano12193444

Chicago/Turabian Style

Han, Fei, Sai Ma, Dong Li, Md Mofasserul Alam, and Zeheng Yang. 2022. "A Simple Fabrication of Sb2S3/TiO2 Photo-Anode with Long Wavelength Visible Light Absorption for Efficient Photoelectrochemical Water Oxidation" Nanomaterials 12, no. 19: 3444. https://doi.org/10.3390/nano12193444

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