Next Article in Journal
Influence of SiO2 Embedding on the Structure, Morphology, Thermal, and Magnetic Properties of Co0.4Zn0.4Ni0.2Fe2O4 Particles
Next Article in Special Issue
Photocatalytic Activity of MoS2 Nanoflower-Modified CaTiO3 Composites for Degradation of RhB under Visible Light
Previous Article in Journal
Nanomaterials for Molecular Detection and Analysis of Extracellular Vesicles
Previous Article in Special Issue
Fluorometric Sensing and Detection of p-Nitroaniline by Mixed Metal (Zn, Ni) Tungstate Nanocomposite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cobalt-Phosphate (Co-Pi)-Modified WO3 Photoanodes for Performance-Enhanced Photoelectrochemical Wastewater Degradation

College of Chemistry and Chemical Engineering, Qingdao University, Qingdao 266071, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2023, 13(3), 526; https://doi.org/10.3390/nano13030526
Submission received: 26 December 2022 / Revised: 20 January 2023 / Accepted: 25 January 2023 / Published: 28 January 2023
(This article belongs to the Special Issue Advanced Nanocomposite Materials for Water and Wastewater Treatment)

Abstract

:
Photocatalytic technology, with features of wide applicability, mild reaction conditions and sunlight availability, satisfies the requirements of “green chemistry”. As the star photoanode material for photoelectrochemical catalysis, WO3 has a suitable band gap of 2.8 eV and a strong oxidation capacity, as well as displaying great potential in organic wastewater degradation. However, its performance is usually hindered by competition with water oxidation to generate peroxides, rapid charge complexation caused by surface defect sites, and so on. Herein, WO3 films modified with cobalt–phosphate (Co-Pi/WO3) film were prepared and involved in photocatalytic organic wastewater degradation. A degradation rate constant of 0.63311 h−1 was obtained for Co-Pi/WO3, which was much higher than that of WO3, 10.23 times that of direct photocatalysis (DP) and 23.99 times that of electrocatalysis (EC). After three cycles of degradation, the film can maintain a relatively good level of stability and a degradation efficiency of 93.79%.

1. Introduction

Photocatalytic organic wastewater degradation has always been a research hotspot in the field of environmental protection [1,2,3,4,5,6,7,8,9]. Photocatalytic technology, with features of wide applicability, mild reaction conditions and sunlight availability, satisfies the requirements of “green chemistry” [10]. However, the photo-generated electrons/holes in photocatalytic process are extremely easy to recombine, which affects the photon quantum efficiency and becomes the main barrier for practical application [11,12]. Sensible solutions are urgently needed in the field of photocatalytic organic wastewater degradation.
Photoelectrochemical catalysis (PEC), which can effectively promote the rapid transfer of electrons, inhibit the photogenerated electron–hole pair recombination, and facilitate the generation of strong oxidative free radicals, displays great potential in organic wastewater degradation [13,14,15]. More importantly, the catalyst is immobilized on substrate for electrodes to effectively avoid secondary pollution [16,17,18]. Since the first attempt in 1972 by Fujishima A and Honda K, photoelectrochemical catalysis has experienced rapid development involving reactors and new catalytic materials [19]. Many n-type semiconductors, such as TiO2, BiVO4, ZnO, WO3, etc., have been widely studied and applied in the field of photoelectric catalysis as photoanodes.
As the star photoanode material for photoelectrochemical catalysis, WO3 has a suitable band gap of 2.8 eV, a strong oxidation capacity [20], can capture about 12% of the solar spectrum and can absorb up to 500 nm of visible light [21,22]. The strong oxidizing substances–hydroxyl radicals can also be obtained in the water body, and the complex and refractory organic pollutants in the water body can be quickly and efficiently mineralized into CO2 and H2O for the complete mineralization of organic pollutants. However, WO3 also has limitations in application, such as competition with water oxidation to generate peroxides and rapid charge complexation caused by surface defect sites [23,24].
It is generally known that co-catalysts can offer certain active sites on the photoelectrode’s surface, accelerating the trapping of carriers, making it easier to separate photogenerated electron-hole pairs, and enhancing the photoelectrode’s PEC performance [25]. Cobalt-phosphate (Co-Pi) is a crucial co-catalyst among many because of its noble-free nature, capacity to speed up hydrogen precipitation reactions and improve the rate of oxygen precipitation [26,27,28,29,30,31,32,33,34,35]. The Co-Pi-modified BiVO4/Cu2O photoanode exhibits high activity and potential for methylene blue (MB) and fluoroquinolones antibiotic (CIP) degradation when used to catalyze the degradation of organic pollutants, and it is anticipated to be used consistently to catalyze the degradation of refractory organic pollutants in water [36]. In research, the Ga-doped AgInS2 photoanodes modified with the Co-Pi co-catalyst exhibit superior visible-light-driven PEC performance [37]. However, the assembly method of Co-Pi displays a considerable impact on the performance. Electrodeposition is a quick and efficient way to thicken the Co-Pi layer. Previous studies have shown that Co-Pi via electrodeposition is easily formed on film vacancies, fractures, or even conductive substrates. When the Co-Pi layer is too thick, the photocatalytic activity of the film is compromised [26]. By employing solely their own photogenerated holes and enabling them to be oxidized where the holes are produced, photo-assisted electrodeposition is an efficient way to prepare semiconductor photoanodes without the need for applied voltage. The photodeposition method is very effective at depositing on the most active and most hole-prone locations in the semiconductor because the photodissolution of aquatic oxygen also necessitates the consumption of photogenerated holes. With the least amount of catalyst in the photolytic water, photo-assisted deposition can, therefore, self-select the best locations for selective deposition, resulting in efficient photolytic oxygenation.
Herein, to increase the practicality of WO3 thin film photoanodes in photoelectrochemical cells, a substantial improvement was made in the photoelectrochemical performance and electrode stability of WO3 thin film photoanodes. To avoid a high-pressure environment, the WO3 films were first prepared using a straightforward and practical atmospheric pressure solvothermal method. A low-cost Co-Pi was further selected to modify the WO3 films, using photo-assisted electrodeposition as opposed to traditional electrodeposition methods, which produced a more uniform and sparser layer of Co-Pi on the semiconductor electrodes. Compared to the bare WO3 films, a 35.95% increase in current density was obtained at 1.23 V vs. RHE. In the degradation of methylene blue-simulated waste solution, the PEC performance of the Co-Pi/WO3 was 1.19 times higher than that of the pure WO3 films. After three cycles of degradation, the film maintained a degradation efficiency of 93.79%. Through a combination of suitable co-catalysis, our work offers an efficient method for the design and manufacture of their highly active, affordable photocatalysts.

2. Experimental Section

2.1. Materials

Analytically pure grade ethanol and hydrochloric acid reagents were purchased from Sinopharm Chemical Regent Co., Ltd. (Shanghai, China). Analytically pure grade sodium tungstate, tungstic acid, hydrogen peroxide, ammonium oxalate, and cobalt nitrate hexahydrate reagents were purchased from Shanghai Aladdin Biochemical Technology Co., Ltd. (Shanghai, China).

2.2. Preparation of WO3 Films

In the process of [38], a simple atmospheric pressure solution thermal deposition process was designed for the preparation of WO3 films based on the specific chemical reaction process of sodium tungstate. First, the fluorine-doped tin oxide (FTO) conductive glass was ultrasonically cleaned in acetone, anhydrous ethanol and distilled water for 30–60 min, then blown dry and set aside. An amount of 0.002 mol of Na2WO4·2H2O and 1.3 g/L of ammonium oxalate was weighed, dissolved in distilled water and concentrated hydrochloric acid (36% to 38%) added to the solution while stirring vigorously to ensure the pH of the solution was less than 1 [39]. According to the reaction formula W O 4 2 + 2 H + H 2 W O 4 , the excess hydrochloric acid causes all the tungstate ions to be precipitated in the form of tungstic acid. After stirring for a while, hydrogen peroxide (60%) was gradually added dropwise, and the tungstic acid suspension gradually became clear and transparent. This was found in the form of peroxynitric acid, which forms polyperoxynitric acid under the bridging action of oxalate- ions. Then, a 35% volume ratio of ethanol (the volume of ethanol to the volume of the solution) was added as a reducing agent to 80 mL of liquid, the conductive glass was placed conductive surface downward in a beaker, and the reaction proceeded slowly at 85 °C. After 240 min of reaction, the bright yellow film was taken out, dried at 60 °C overnight, placed in a vacuum tube sintering furnace and calcined to obtain WO3 films.

2.3. Preparation of CO-Pi/WO3 Films

The calcined WO3 film was modified with Co-Pi by the photo-assisted electro-deposition (PED) method. Specifically, the photoanode film was modified with 100 s as the unit under the three-electrode system, with 0.1 M of phosphate buffer as the electrolyte, 0.5 mM of cobalt nitrate concentration in solution and 0.1 V of voltage.

2.4. Characterization of Thin Films

The film structure was characterized by an X-ray diffractometer (XRD) (UltimaⅣ); scanning electron microscope (SEM) (INCA x-sight); high-resolution transmission electron microscope (HR-TEM) (FEI Tecnai G2 F20, USA); X-ray photoelectron spectroscopy (XPS) (Thermo Fisher, ESCALAB 250Xi, Waltham, MA, USA); ultraviolet-visible diffuse reflectance spectroscopy (UV-vis DRS) (Shimadzu UV3600); and energy-dispersive X-ray spectroscopy elemental analysis (EDS Mapping) (Zeiss Gemini 300, Birmingham, UK).

3. Results and Discussion

The crystallinity of Co-Pi/WO3 was investigated with X-ray diffraction (XRD) patterns (Figure 1A). The eleven peaks, located at 23.05°, 23.66°, 24.34°, 26.52°, 28.76°, 33.48°, 33.7°, 34.16°, 41.26°, 42.08° and 50.04°, were ascribed to the (002), (020), (200), (120), (112), (022), (−202), (202), (−222), (222) and (140) planes of WO3 (PDF#20-1323), respectively. No obvious characteristic peaks for Co-Pi were revealed in the XRD patterns for low loading. X-ray photoelectron spectroscopy (XPS) tests were further executed to investigate the valence state and chemical composition of WO3 and WO3/Co-Pi. The XPS survey spectra demonstrated the presence of W, O and Co elements (Figure 1B). As shown in Figure 1C, the W 4 f5/2 and W 4 f7/2 peaks are ~37.4 eV and ~35.3 eV, and the binding energy gap between the peaks is around 2.1 eV, indicating that W is in the film at +6 valence after calcination [40,41,42]. The peak values of O1 s are given in Figure 1D, with the main body at ~530.1 eV, indicating that the chemical state was mainly present in the metal oxides, attributed to lattice oxygen, while at ~532 eV (blue part of Figure 1D) this was related to the presence of surface hydroxyl groups and ligand unsaturated oxygen species [43,44]. Initially, the surface of the prepared film was loaded with elemental Co-Pi.
The morphologies and surface structure of WO3 and Co-Pi/WO3 were investigated by SEM and HR-TEM. As shown in Figure 2A–C, the WO3 films prepared and calcined by the isothermal method at atmospheric pressure were continuous and compact nanoflakes, with a measured lattice spacing of 0.38 nm for the WO3. Compared with the data of the standard comparison card (PDF#43-1035), the corresponding crystal plane of the material was plane (002). Surface (002) was the main exposure surface of the prepared WO3 film. As can be distinctly seen in Figure 2D,E, the Co-Pi-modified film had no change in morphology and was still tightly and continuously aligned with a lamellar structure. The measured lattice spacing of the WO3/Co-Pi films was 0.38 nm, according to the HR-TEM images (Figure 2F). No lattice spacing was detected for elemental Co-Pi for the small percentage due to the low loading, which was also similar to the published works [26]. The corresponding EDS mapping images (Figure 2G) showed the uniform distribution of W, O, P, and Co over the entire structure, which further confirms that the photoassisted electrodeposition method successfully deposited Co-Pi on the surface of the film, where it could be evenly distributed. The mass ratios of W, O, P and Co were 72.33%, 26.86%, 2.014% and 6.924%, respectively. The combination of XRD, XPS, SEM and HR-TEM results showed that the nano-sized tungsten trioxide films were prepared by the isothermal method and the Co element was successfully deposited on the surface of the films by photoassisted electrodeposition.
The photovoltaic properties were investigated using a three-electrode system in a pretreated 0.1 M NaSO4 electrolyte. Comparative experiments were conducted to explore the optimum process conditions for the WO3 films. Specifically, the photovoltaic properties of the WO3 films were investigated by controlling the amount of ammonium oxalate addition, reaction time, reaction temperature and ethanol dosage. As shown in Figures S1–S3, with the gradual increase in ammonium oxalate addition, the photoelectric properties showed a trend of first increasing and then decreasing. The optimum was reached at 1.3 g/L (0.54 mW·cm−2). In the impedance diagram (Figure S4), the resistance of the photocatalytic process showed a trend of decreasing and then increasing with the increase in ammonium oxalate addition. The reason for this change may be due to the hydrolysis of ammonium oxalate in solution to generate electron-rich groups (oxalic acid) [20], which form hydrogen bonds with the water molecules at the end of WO3·H2O and promote the orderly and continuous growth of the film. However, ammonium oxalate, as a common reducing agent, is prone to react with oxidants when heated to decomposition [45]. The addition of excess ammonium oxalate would lead to too-rapid precipitation of WO3·H2O, and some of the WO3·H2O would leak out from the bottom of the reactor in the form of precipitation, resulting in weak and discontinuous film growth, which affects the performance of the film. The reaction time and temperature control (Figures S5–S10) showed that at a reaction time of 240 min and a reaction temperature of 85 °C, the film had the lowest resistance, the highest photoresponse current and the best photovoltaic performance (0.54 mW·cm−2). After comparison (Figure S11), it was found that the (200) diffraction peak intensity of WO3 films increased and the (002) and (020) diffraction peak intensity decreased during the increase in ethanol dosage from 20 vol.% to 35 vol.%; then, when the dosage was gradually increased to 45 vol.%, the WO3 films’ (200) diffraction peak intensity decreased and the (002) and (020) diffraction peak intensity increased. The intensity of the (200) diffraction peak decreased and the (002) and (020) diffraction peaks increased when the dosage was gradually increased to 45 vol. The best photovoltaic performance was achieved at a specific dosage of 35 vol.% (Figures S12 and S13).
The results from the overall range of LSV plots (Figure 3A) and the photoresponse currents (Figure 3B) show that the optoelectronic properties of the Co-Pi-modified WO3 films were not only more stable, but also had a 35.95% increase in performance compared to the WO3 films alone at 1.23 V vs. improved performance. This was also confirmed by the film degradation performance tests. After exploring the optimum conditions, the photovoltaic properties of the optimized WO3 films were compared with those of WO3 films loaded with CO-Pi. The photocatalytic activity was investigated in a simulated waste solution of MB at an initial concentration of 5 mg/L using a three-electrode system. The degradation efficiency of the simulated waste solution was determined with the concentration of the simulated waste solution and the degradation rate at different times, based on the standard curve of MB staining solution (Figure S14). As shown in Figure 3C, the degradation concentration gradually decreased as the degradation time increased. After two hours of degradation, according to the degradation trend and degradation kinetic curves (Figure 3D,E), the degradation capacity of the WO3 film after CO-Pi modification was much greater than that of DP and EC. The highest degradation rate constant was 0.63311 h−1, which was 1.19 times higher than the degradation rate constant of the WO3 film before modification. Compared with previous works, superior performance was obtained for Co-Pi/WO3 (Table S1). In order to assess the stability of the films, three cycles of degradation were carried out (Figure 3F). The third time maintained a degradation efficiency of 93.79% compared to the first time and was higher than the pure WO3 film. This indicates that the film has good stability.

4. Conclusions

In conclusion, Co-Pi/WO3 film was prepared by light-assisted deposition, as demonstrated by XRD, XPS, SEM, TEM and HR-TEM, without altering the distinctive shape of the pure WO3 films involved in photocatalytic organic wastewater degradation. A degradation rate constant of 0.63311 h−1 was obtained for Co-Pi/WO3, which was much higher than that of WO3, 10.23 times that of DP and 23.99 times that of EC. After three cycles of degradation, the film can maintain a relatively good stability and a degradation efficiency of 93.79%. The superior performance was ascribed to the involvement of Co-Pi, which could reduce the compounding of photogenerated electron holes, increasing the efficiency of photogenerated electron transport, boosting the photocurrent and stabilizing the photoanodes. This work provides a new approach to the problem of wastewater pollution and a basis for energy conversion.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13030526/s1, Figure S1 LSV curves of WO3 films prepared at different ammonium oxalate contents under 100 mW·cm-2 light; Figure S2 Photoresponse current profiles for different ammonium oxalate additions; Figure S3 Trend of current density at 1.23V vs. RHE for films with different ammonium oxalate additions; Figure S4 Impedance mapping of different ammonium oxalate additions; Figure S5 LSV curves of WO3 films prepared at different reaction times under 100 mW·cm-2 light; Figure S6 Impedance mapping of WO3 films at different reaction times; Figure S7 Photoresponse current mapping of WO3 films at different reaction times; Figure S8 Trend of current density at 1.23V vs. RHE for WO3 films at different reaction times; Figure S9 LSV curves of WO3 films prepared at different reaction temperatures under 100 mW·cm-2 light; Figure S10 Impedance diagram of WO3 films at different temperatures; Figure S11 XRD patterns of WO3 films prepared at different ethanol dosages; Figure S12 LSV curves of WO3 prepared at different ethanol dosages under 100 mW·cm-2 light; Figure S13 Impedance diagram of WO3 films with different ethanol dosages; Figure S14 UV-Vis absorptions spectra (A) and calibration curve (B) used for calculation of methylene blue staining solution.

Author Contributions

J.Z.: Investigation, Writin-Original Draft. W.S.: Investigation, Writing-Original Draft. X.D.: Supervision, Conceptualization, Methodology, K.X.: Methodology, Project administration, Visualization. T.L.: Supervision, Writing-Review & Editing, Resources, Methodology. X.Z.: Supervision, Writing-Review & Editing, Resources, Conceptualization. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Major Science and Technology Innovation Project of Shandong (No. 2019JZZY010507); the Qingdao Municipal Science and Technology Bureau (No. 17-1-1-86-jch); the Key Technology Research and Development Program of Shandong (No. 2018GGX108005); the China Postdoctoral Science Foundation (No. 2019T120571; No. 2018M632623); and Qingdao Postdoctoral Foundation (RZ2000003344).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, F.; Zhao, P.; Niu, M.; Maddy, J. The survey of key technologies in hydrogen energy storage. Int. J. Hydrog. Energy 2016, 41, 14535–14552. [Google Scholar] [CrossRef]
  2. Schultz, D.; Yoon, T. Solar Synthesis: Prospects in Visible Light Photocatalysis. Science 2014, 343, 985. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Hisatomi, T.; Domen, K. Reaction systems for solar hydrogen production via water splitting with particulate semiconductor photocatalysts. Nat. Catal. 2019, 2, 387–399. [Google Scholar] [CrossRef]
  4. Awfa, D.; Ateia, M.; Fujii, M.; Yoshimura, C. Photocatalytic degradation of organic micropollutants: Inhibition mechanisms by different fractions of natural organic matter. Water Res. 2020, 174, 11. [Google Scholar] [CrossRef]
  5. Lin, Z.; Li, J.; Shen, W.; Corriou, J.; Chen, X.; Xi, H. Different photocatalytic levels of organics in papermaking wastewater by flocculation-photocatalysis and SBR-photocatalysis: Degradation and GC? MS experiments, adsorption and photocatalysis simulations. Chem. Eng. J. 2021, 412, 14. [Google Scholar] [CrossRef]
  6. Rueda-Marquez, J.; Levchuk, I.; Ibanez, P.; Sillanpaa, M. A critical review on application of photocatalysis for toxicity reduction of real wastewaters. J. Clean. Prod. 2020, 258, 13. [Google Scholar] [CrossRef]
  7. Zhao, Y.; Wang, Y.; Chi, H.; Zhang, Y.; Sun, C.; Wei, H.; Li, R. Coupling photocatalytic water oxidation on decahedron BiVO4 crystals with catalytic wet peroxide oxidation for removing organic pollutions in wastewater. Appl. Catal. B Environ. 2022, 318, 9. [Google Scholar] [CrossRef]
  8. Garcia-Segura, S.; Brillas, E. Applied photoelectrocatalysis on the degradation of organic pollutants in wastewaters. J. Photochem. Photobiol. C Photochem. Rev. 2017, 31, 1–35. [Google Scholar] [CrossRef]
  9. Sreedhar, N.; Kumar, M.; Al Jitan, S.; Thomas, N.; Palmisano, G.; Arafat, H. 3D printed photocatalytic feed spacers functionalized with beta-FeOOH nanorods inducing pollutant degradation and membrane cleaning capabilities in water treatment. Appl. Catal. B Environ. 2022, 300, 120318. [Google Scholar] [CrossRef]
  10. Al-Nuaim, M.; Alwasiti, A.; Shnain, Z. The photocatalytic process in the treatment of polluted water. Chem. Pap. 2022. [Google Scholar] [CrossRef]
  11. Xu, P.; McCool, N.; Mallouk, T. Water splitting dye-sensitized solar cells. Nano Today 2017, 14, 42–58. [Google Scholar] [CrossRef]
  12. Henderson, M. A surface science perspective on TiO2 photocatalysis. Surf. Sci. Rep. 2011, 66, 185–297. [Google Scholar]
  13. Jia, J.; Seitz, L.; Benck, J.; Huo, Y.; Chen, Y.; Ng, J.; Bilir, T.; Harris, J.; Jaramillo, T. Solar water splitting by photovoltaic-electrolysis with a solar-to-hydrogen efficiency over 30%. Nat. Commun. 2016, 7, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Jiang, C.; Moniz, S.; Wang, A.; Zhang, T.; Tang, J. Photoelectrochemical devices for solar water splitting—Materials and challenges. Chem. Soc. Rev. 2017, 46, 4645–4660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Guo, Q.; Zhou, C.; Ma, Z.; Yang, X. Fundamentals of TiO2 Photocatalysis: Concepts, Mechanisms, and Challenges. Adv. Mater. 2019, 31, 26. [Google Scholar] [CrossRef]
  16. Zhang, K.; Ma, M.; Li, P.; Wang, D.; Park, J. Water Splitting Progress in Tandem Devices: Moving Photolysis beyond Electrolysis. Adv. Energy Mater. 2016, 6, 16. [Google Scholar] [CrossRef]
  17. Jacobsson, T.; Fjällström, V.; Edoff, M.; Edvinsson, T. Sustainable solar hydrogen production: From photoelectrochemical cells to PV-electrolyzers and back again. Energy Environ. Sci. 2014, 7, 2056–2070. [Google Scholar] [CrossRef]
  18. Balberg, I. Energy conversion by photoelectrochemical cells: Principles and applications. Vacuum 1983, 33, 579–583. [Google Scholar] [CrossRef]
  19. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  20. Su, J.; Feng, X.; Sloppy, J.; Guo, L.; Grimes, C. Vertically Aligned WO3 Nanowire Arrays Grown Directly on Transparent Conducting Oxide Coated Glass: Synthesis and Photoelectrochemical Properties. Nano Lett. 2010, 11, 203–208. [Google Scholar] [CrossRef]
  21. Olejnicek, J.; Brunclikova, M.; Kment, S.; Hubicka, Z.; Kmentova, H.; Ksirova, P.; Cada, M.; Zlamal, M.; Krysa, J. WO3 thin films prepared by sedimentation and plasma sputtering. Chem. Eng. J. 2017, 318, 281–288. [Google Scholar] [CrossRef]
  22. Shin, S.; Han, H.; Kim, J.; Park, I.; Lee, M.; Hong, K.; Cho, I. A tree-like nanoporous WO3 photoanode with enhanced charge transport efficiency for photoelectrochemical water oxidation. J. Mater. Chem. A 2015, 3, 12920–12926. [Google Scholar] [CrossRef]
  23. Kim, W.; Tachikawa, T.; Monllor-Satoca, D.; Kim, H.; Majima, T.; Choi, W. Promoting water photooxidation on transparent WO3 thin films using an alumina overlayer. Energy Environ. Sci. 2013, 6, 3732–3739. [Google Scholar] [CrossRef]
  24. Choi, H.; Kim, E.; Hahn, S. Photocatalytic activity of Au-buffered WO3 thin films prepared by RF magnetron sputtering. Chem. Eng. J. 2010, 161, 285–288. [Google Scholar]
  25. Fernandes, E.; Gomes, J.; Martins, R.C. Semiconductors Application Forms and Doping Benefits to Wastewater Treatment: A Comparison of TiO2, WO3, and g-C3N4. Catalysts 2022, 12, 1218. [Google Scholar] [CrossRef]
  26. Zhong, D.; Cornuz, M.; Sivula, K.; Grätzel, M.; Gamelin, D.R. Photo-assisted electrodeposition of cobalt–phosphate (Co–Pi) catalyst on hematite photoanodes for solar water oxidation. Energy Environ. Sci. 2011, 4, 1759–1764. [Google Scholar] [CrossRef]
  27. Gamelin, D.R. Water splitting: Catalyst or spectator? Nat. Chem. 2012, 4, 965–967. [Google Scholar]
  28. Carroll, G.; Gamelin, D.R. Kinetic analysis of photoelectrochemical water oxidation by mesostructured Co-Pi/α-Fe2O3 photoanodes. J. Mater. Chem. A 2015, 4, 2986–2994. [Google Scholar] [CrossRef]
  29. Seabold, J.; Choi, K.-S. Effect of a Cobalt-Based Oxygen Evolution Catalyst on the Stability and the Selectivity of Photo-Oxidation Reactions of a WO3 Photoanode. Chem. Mater. 2011, 23, 1105–1112. [Google Scholar] [CrossRef]
  30. Ostachavičiūtė, S.; Šulčiūtė, A.; Valatka, E. The morphology and electrochemical properties of WO3 and Se-WO3 films modified with cobalt-based oxygen evolution catalyst. Mater. Sci. Eng. B 2020, 260, 114630. [Google Scholar] [CrossRef]
  31. Surendranath, Y.; Kanan, M.; Nocera, D.G. Mechanistic Studies of the Oxygen Evolution Reaction by a Cobalt-Phosphate Catalyst at Neutral pH. J. Am. Chem. Soc. 2010, 132, 16501–16509. [Google Scholar] [CrossRef]
  32. Steinmiller, E.; Choi, K.-S. Photochemical deposition of cobalt-based oxygen evolving catalyst on a semiconductor photoanode for solar oxygen production. Proc. Natl. Acad. Sci. USA 2009, 106, 20633–20636. [Google Scholar] [CrossRef] [Green Version]
  33. Brodsky, C.; Bediako, D.; Shi, C.; Keane, T.; Costentin, C.; Billinge, S.; Nocera, D.G. Proton–Electron Conductivity in Thin Films of a Cobalt–Oxygen Evolving Catalyst. ACS Appl. Energy Mater. 2019, 2, 3–12. [Google Scholar] [CrossRef]
  34. Tsuneda, T.; Ten-no, S. Water–oxidation mechanism of cobalt phosphate co-catalyst in artificial photosynthesis: A theoretical study. Phys. Chem. Chem. Phys. 2022, 24, 4674–4682. [Google Scholar] [CrossRef] [PubMed]
  35. Kanan, M.; Nocera, D. In Situ Formation of an Oxygen-Evolving Catalyst in Neutral Water Containing Phosphate and Co2+. Science 2008, 321, 1072–1075. [Google Scholar]
  36. Li, X.; Wan, J.; Ma, Y.; Wang, Y.; Li, X. Study on cobalt-phosphate (Co-Pi) modified BiVO4/Cu2O photoanode to significantly inhibit photochemical corrosion and improve the photoelectrochemical performance. Chem. Eng. J. 2021, 404, 127054. [Google Scholar]
  37. Cai, Q.; Liu, Z.; Li, J.; Han, C.; Tong, Z. Ga-Doped AgInS2 Modified with Co–Pi Co–catalyst for Efficient Photoelectrochemical Water Splitting. Catal. Lett. 2020, 150, 1089–1097. [Google Scholar]
  38. Zeng, Q.; Li, J.; Bai, J.; Li, X.; Xia, L.; Zhou, B. Preparation of vertically aligned WO3 nanoplate array films based on peroxotungstate reduction reaction and their excellent photoelectrocatalytic performance. Appl. Catal. B: Environ. 2017, 202, 388–396. [Google Scholar] [CrossRef]
  39. Patrick, E.; Orazem, M.; Sanchez, J.; Nishida, T. Corrosion of tungsten microelectrodes used in neural recording applications. J. Neurosci. Methods 2011, 198, 158–171. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Kalanur, S.; Yoo, I.-H.; Eom, K.; Seo, H. Enhancement of photoelectrochemical water splitting response of WO3 by Means of Bi doping. J. Catal. 2017, 357, 127–137. [Google Scholar] [CrossRef]
  41. You, Y.; Tian, W.; Min, L.; Cao, F.; Deng, K.; Li, L. TiO2/WO3 Bilayer as Electron Transport Layer for Efficient Planar Perovskite Solar Cell with Efficiency Exceeding 20%. Adv. Mater. Interfaces 2019. [Google Scholar] [CrossRef]
  42. Zhan, F.; Li, J.; Li, W.; Yang, Y.; Liu, W.; Li, Y. In situ synthesis of CdS/CdWO4/WO3 heterojunction films with enhanced photoelectrochemical properties. J. Power Sources 2016, 325, 591–597. [Google Scholar] [CrossRef]
  43. Ramachandran, R.; Dendooven, J.; Detavernier, C. Plasma enhanced atomic layer deposition of Fe2O3 thin films. J. Mater. Chem. A 2014, 2, 10662–10667. [Google Scholar] [CrossRef]
  44. Ge, R.; Li, L.; Su, J.; Lin, Y.; Tian, Z.; Chen, L. Ultrafine Defective RuO2 Electrocatayst Integrated on Carbon Cloth for Robust Water Oxidation in Acidic Media. Adv. Energy Mater. 2019, 9, 9. [Google Scholar]
  45. Verma, A.; Johnson, G.; Corbin, D.; Shiflett, M. Separation of Lithium and Cobalt from LiCoO2: A Unique Critical Metals Recovery Process Utilizing Oxalate Chemistry. ACS Sustain. Chem. Eng. 2020, 8, 6100–6108. [Google Scholar] [CrossRef]
Figure 1. (A) XRD patterns of WO3 film and Co-Pi/WO3 film; (B) XPS patterns of WO3 film and Co-Pi/WO3 film; (C) W 4f XPS spectra of Co-Pi/WO3; (D) O 1s XPS spectra of Co-Pi/WO3.
Figure 1. (A) XRD patterns of WO3 film and Co-Pi/WO3 film; (B) XPS patterns of WO3 film and Co-Pi/WO3 film; (C) W 4f XPS spectra of Co-Pi/WO3; (D) O 1s XPS spectra of Co-Pi/WO3.
Nanomaterials 13 00526 g001
Figure 2. (AC) SEM, TEM, HR-TEM images of WO3; (DF) SEM, HR-TEM images of Co-Pi/WO3; (G) EDS elemental mapping images of W, O, P and Co in Co-Pi/WO3.
Figure 2. (AC) SEM, TEM, HR-TEM images of WO3; (DF) SEM, HR-TEM images of Co-Pi/WO3; (G) EDS elemental mapping images of W, O, P and Co in Co-Pi/WO3.
Nanomaterials 13 00526 g002
Figure 3. (A) LSV curves of WO3 films before and after Co-Pi modification under 100 mW·cm−2 light; (B) photoresponse current profiles of WO3 films before and after Co-Pi modification; (C) UV absorption spectra of MB degraded by Co-Pi/WO3 films under AM1.5 illuminations at 1 V applied voltage; (D) DP, EC and PEC degradation of MB under AM1.5 illuminations at 1 V applied bias voltage before and after Co-Pi modification of WO3 films; (E) corresponding kinetic curves before and after Co-Pi modification of WO3 films; (F) variation of the performance of MB degradation under PEC conditions with three cycles.
Figure 3. (A) LSV curves of WO3 films before and after Co-Pi modification under 100 mW·cm−2 light; (B) photoresponse current profiles of WO3 films before and after Co-Pi modification; (C) UV absorption spectra of MB degraded by Co-Pi/WO3 films under AM1.5 illuminations at 1 V applied voltage; (D) DP, EC and PEC degradation of MB under AM1.5 illuminations at 1 V applied bias voltage before and after Co-Pi modification of WO3 films; (E) corresponding kinetic curves before and after Co-Pi modification of WO3 films; (F) variation of the performance of MB degradation under PEC conditions with three cycles.
Nanomaterials 13 00526 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, J.; Sun, W.; Ding, X.; Xia, K.; Liu, T.; Zhang, X. Cobalt-Phosphate (Co-Pi)-Modified WO3 Photoanodes for Performance-Enhanced Photoelectrochemical Wastewater Degradation. Nanomaterials 2023, 13, 526. https://doi.org/10.3390/nano13030526

AMA Style

Zhang J, Sun W, Ding X, Xia K, Liu T, Zhang X. Cobalt-Phosphate (Co-Pi)-Modified WO3 Photoanodes for Performance-Enhanced Photoelectrochemical Wastewater Degradation. Nanomaterials. 2023; 13(3):526. https://doi.org/10.3390/nano13030526

Chicago/Turabian Style

Zhang, Jiakun, Weixu Sun, Xin Ding, Kai Xia, Tao Liu, and Xiaodong Zhang. 2023. "Cobalt-Phosphate (Co-Pi)-Modified WO3 Photoanodes for Performance-Enhanced Photoelectrochemical Wastewater Degradation" Nanomaterials 13, no. 3: 526. https://doi.org/10.3390/nano13030526

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