Next Article in Journal
Effect of Heat Treatment on Structure of Carbon Shell-Encapsulated Pt Nanoparticles for Fuel Cells
Previous Article in Journal
Stability of Graphene/Nafion Composite in PEM FC Electrodes
Previous Article in Special Issue
WO3 Nanoplates Decorated with Au and SnO2 Nanoparticles for Real-Time Detection of Foodborne Pathogens
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Bulk Oxygen Vacancy Dominating WO3−x Photocatalyst for Carbamazepine Degradation

1
GRINM (Guangdong) Institute for Advanced Materials and Technology, Foshan 528000, China
2
Guangdong Provincial Key Laboratory of Environmental Pollution Control and Remediation Technology, School of Environmental Science and Engineering, Sun Yat-sen University, Guangzhou 510006, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2024, 14(11), 923; https://doi.org/10.3390/nano14110923
Submission received: 24 March 2024 / Revised: 15 May 2024 / Accepted: 17 May 2024 / Published: 24 May 2024
(This article belongs to the Special Issue The Application of Nanosensors in Energy and Environment)

Abstract

:
Creating oxygen vacancy in tungsten trioxide (WO3) has been considered as an effective strategy to improve the photocatalytic performance for degrading organic pollutants. In this study, oxygen vacancies were introduced into WO3 by thermal treatment under Ar atmosphere and their proportion was changed by setting different treatment times. WO3−x samples show better photoelectric properties and photocatalytic degradation performance for carbamazepine (CBZ) than an oxygen-vacancy-free sample, and WO3−x with the optimal proportion of oxygen vacancies is obtained by thermal treatment for 3 h in 550 °C. Furthermore, it discovers that the surface oxygen vacancies on WO3−x would be recovered when it is exposed to air, resulting in a bulk oxygen vacancy dominating WO3−x (bulk-WO3−x). The bulk-WO3−x exhibited much higher degradation efficiency for CBZ than WO3−x with both surface and bulk oxygen vacancies. The mechanism study shows bulk-WO3−x mainly degrades the CBZ by producing OH radicals and superoxide radicals, while oxygen-vacancy-free sample mainly oxidizes the CBZ by the photoexcited hole, which requires the CBZ to be adsorbed on the surface for degradation. The radical generated by bulk-WO3−x exhibits stronger oxidizing capacity by migrating to the solution for CBZ degradation. In summary, the influence of oxygen vacancy on photocatalytic degradation performance depends on both the proportion and location distribution and could lie in the optimization of the photodegradation mechanism. The results of this study could potentially broaden our understanding of the role of oxygen vacancies and provide optimal directions and methods for oxygen vacancy regulation for photocatalysts.

1. Introduction

Along with industrialization and urbanization, organic pollutants have caused severe environmental problems and resulted in increasingly severe shortage of water resources [1]. Therefore, there is urgent demand for developing new techniques for resolving the problem of organic pollutants. In the past decades, research attention has been focused on developing and cost-effective photocatalytic technique as it is a green and efficient approach for degradation of organic pollutants [2,3,4,5,6,7]. In a photocatalytic degradation system, photocatalyst plays a vital role in the degradation efficiency of pollutants. Among all investigated photocatalysts, tungsten trioxide (WO3) exhibits a competitive advantage for the degradation of organic pollutants. Owing to its relatively narrow band gap (2.4–2.8 eV) and positive valence band (VB), WO3 can utilize visible light energy and have stronger oxidation ability [8,9]. They are both beneficial to oxidize organic pollutants. However, WO3, as a photocatalyst, also has the shortcomings of low photogenerated carries density, poor charge transfer efficiency and electron–hole separation [9].
Previous studies have suggested that introducing defect structure into WO3 could be an effective way to overcome the above shortcomings, and oxygen vacancy is a simple but effective defect structure for improving the photocatalytic performance of WO3 [10]. According to the previous literature, the promotion from oxygen vacancy can be reflected in several aspects. Some research demonstrates that oxygen vacancy improves the utilization of light energy due to extending light absorption of WO3 in the visible regions [11]. Some studies demonstrated that oxygen vacancy can effectively promote the separation of photogenerated electron–hole so that it can maximize the utilization rate of photogenerated carriers [12]. Other reports verify that the existence of oxygen vacancy can increase the charge transfer efficiency [13]. However, some reports showed that oxygen vacancies could cause negative influence on the photocatalytic performance because they may act as traps for photogenerated charge carriers [14], become recombination centers of electron and hole and weaken the crystal structure [15].
Oxygen vacancies can be introduced to WO3 by thermal treatment in inert or reducing atmosphere [16,17,18,19]. It is investigated whether thermal treatment with different atmospheres, temperatures or durations can result in different impacts on the distribution of oxygen vacancies and, thereby, the photocatalytic performance. When the thermal treatment temperature rises from 350 °C to 550 °C, the formation rate of surface oxygen vacancies is fast at lower temperatures, but it would slow down at higher temperatures because of the migration of oxygen atoms from bulk phase to surface [20]. Higher temperatures may be more conducive to the formation of bulk oxygen vacancies. It is considered that surface and bulk oxygen vacancies generated by thermal treatment play different roles on photocatalytic performance of samples [9]. Bulk oxygen vacancies could format intermediate electronic bands [21,22], which can inhibit the recombination of photogenerated electron–hole pairs [23,24]. Meanwhile, surface oxygen vacancies could result in a decrease in VB due to the local band caused by surface defect [25,26]. In addition, wet chemical reduction, including sol-gel method [27] and solvothermal method [28], and electrochemical reduction [29] are also investigated to construct oxygen vacancies in WO3.
This study aims to investigate the influence of oxygen vacancies on the photocatalytic degradation performance of WO3. Here, carbamazepine (CBZ) will be used as a representative pollutant for studying their photocatalytic degradation performance. The proportion of introduced oxygen vacancies into WO3 are changed by setting different thermal treatment times under Ar atmosphere and optimized by analyzing their structures and morphology properties, photoelectrochemistry properties and photocatalytic degradation performance for CBZ. Furthermore, the influence of oxygen vacancy location is also considered by comparing the WO3−x samples with different distributions of surface and bulk oxygen vacancy. Finally, the CBZ degradation mechanism by WO3 and optimized WO3−x are investigated.

2. Materials and Methods

2.1. Preparation of Pristine WO3 and WO3 with Oxygen Vacancies

A pristine WO3 sample was prepared by a hydrothermal method, referring to reported reference [30]. First, 210 mg of sodium tungsten (Na2WO4, Sigma-Aldrich, Burlington, MA, USA) was dissolved in 30 mL of deionized (DI) water. Then, 10 M hydrochloric acid (HCl, Fischer Scientific, Waltham, MA, USA) was added dropwise into the solution until it became turbid, followed by the addition of 210 mg of ammonium oxalate ((NH4)2C2O4, Sigma-Aldrich, Burlington, MA, USA) into the mixture. After 10 min of stirring, the turbid solution became clear and was then transferred into a Teflon-lined stainless-steel autoclave, which was sealed and heated at 120 °C for 12 h. After that, the resulting yellow powder was filtered and washed with ethanol and water. The yellow powder was WO3∙H2O and transformed into WO3 after being calcinated at 550 °C for 3 h.
WO3−x samples: the pristine WO3 nanoparticles were thermally treated in pure Ar flow at 550 °C for 0.5 h, 3 h, 6 h or 10 h to prepare four WO3−x samples, which were denoted as WO3−x-0.5 h, WO3−x-3 h, WO3−x-6 h or WO3−x-10 h samples, respectively.

2.2. Preparation of Pristine WO3 and WO3 with Oxygen Vacancies

Scanning electron microscopy (SEM) images were taken on a Quanta 400 Thermal FE environmental scanning electron microscope (FEI, Eindhoven, The Netherlands). The diffuse reflectance absorption spectra of the samples were obtained by a UV-visible spectrophotometer equipped with an integrated sphere attachment (UV-3600, SHIMADZU, Kyoto, Japan). The X-ray diffraction (XRD) data were recorded on a Rigaku diffraction instrument using a Cu Kαradiation source with λ = 0.15418 nm (SmartLab, Rigaku, Tokyo, Japan). Photoelectrochemistry (PEC) measurements were performed on an electrochemical workstation (CHI760E, Chenhua, Shanghai, China) equipped with a three-electrode cell. The working electrode was made as follows: 2 mg of samples was fixed on FTO conducting glass mixed with 0.05 mL 0.05 wt% Nafion (Sigma-Aldrich, Gillingham, Dorset, UK), using graphite rod as counter electrodes and Ag/AgCl as the reference electrodes, and 0.1 M Na2SO4 was used as the electrolyte. The EPR spectra were detected by a CW/Pulse EPR system (A300, Bruker Co., Bremen, Germany) with a microwave frequency of 9.64 GHz, a microwave power of 0.94 mW, a modulation frequency of 100 kHz and a modulation amplitude of 2.0 G. X-Ray photoelectron spectroscopy (XPS) was performed using a Thermo Fisher Scientific K-Alpha (Thermo Scientific, Waltham, MA, USA) equipped with a VG CLAM 4 MCD electron energy analyzer (Thermo Scientific, Waltham, MA, USA) and twin-anode Mg Kα radiation (1253.6 eV) or Al Kα radiation (1496.3 eV) X-ray sources.

2.3. Photocatalytic Performance Experiments

All photocatalytic degradation experiments were carried out in a quartz tube, with 0.25 g/L photocatalyst and 5 ppm carbamazepine (CBZ, Macklin, Shanghai, China) under ultraviolet-visible irradiation from a 35 W xenon lamp. The control measurement of CBZ under illumination was tested as the blank sample. At an interval of 20 min, 0.5 mL of the solution was extracted. The residual concentrations of CBZ in reaction solution were detected every 20 min by HPLC (Shimadzu LC-20AD, Shimadzu Corporation, Kyoto, Japan) equipped with a Poroshell 120 EC-C18 column (4.6 mm × 100 mm, 2.7 μm, Agilent Technology, Santa Clara, CA, USA). The mobile phase for HPLC detection was a mixture of methanol and water with a ratio of 6:4 and at a 0.6 mL/min flow. The photocatalytic activity was evaluated using a time profile of Ct/C0, where Ct is the concentration of the pollutants at the irradiation time t and C0 is the concentration at the equilibrium before irradiation, respectively.

2.4. Detection and Measurements of Photogenerated Radicals

The formed active oxygen radicals, including ·OH, ·O2− and 1O2, in photocatalytic degradation systems were all identified by a Bruker A300 electron paramagnetic resonance (EPR) spectrometer (Billerica, MA, USA) using 5,5-dimethyl-1-pyrroline-N-oxide (DMPO), DMPO with methanol and TEMP as the trapping agent. Typically, 100 mM of DMPO was mixed with 20 mL solution (including 0.25 g∙L−1 catalyst, 5 ppm CBZ). All samples had been illuminated for 5 min before ERP measurements.

3. Results and Discussion

3.1. Structures and Morphologies Properties

The SEM images show that the WO3 powder particle is a cluster of two-dimensional nanosheets with the size of ca. 1 μm (Figure 1a,b). After being thermally treated for 0.5 h to 10 h, there is no remarkable visual change in particle surface morphology and size of the samples (Figure 1c–f), suggesting the morphology of WO3 is stable.
Figure 2 shows the optical properties of all samples. As compared with untreated WO3, the color of thermally treated WO3 samples changed from yellow to olive-green (0.5 h and 3 h) and then to indigo after 6 h and 10 h thermal treatment (Figure 2a). According to the literature, the change in color could be attributed due to the formation of bulk oxygen vacancies [9]. Correspondingly, their light absorption in the visible regions is extended with the increase in oxygen vacancy (Figure 2b). It can be found that more light can be absorbed when the WO3 samples are thermally treated for a longer time. Figure 2c shows the band gap of the samples calculated by the linear approximation of (αhν) 2 vs. photo energy [31]. Visibly, all WO3−x samples possess a narrower band gap than pristine WO3 due to oxygen vacancy [9,22,32]. The calculated value of pristine WO3 is around 2.77 eV, whereas the optical absorption of the WO3−x samples gradually shifts to the longer wavelengths with the band gap value (Eg) ranging from 2.53 to 2.71 eV. Among all WO3−x samples, WO3−x-10 h shows the most intense light absorption and the narrowest band gap.
The XRD pattern (Figure 3a) of all samples could be indexed to the monoclinic phase of WO3 (JCPDF20-1324). There is no observable shift for all characteristic diffraction peaks of WO3 after thermal treatment. It suggests that the main crystal structure is preserved even though oxygen vacancy is increased [33]. Nevertheless, the intensity of the diffraction peaks slightly weakens with the increase in their thermal-treated time, indicating the degradation of the WO3 crystallinity [34]. Crystalline degradation was mainly reflected in the orientations on (020), (022) and (202) planes (Figure 3b,c).
Based on the results of SEM, UV-vis diffuse reflectance spectra and XRD, it can be concluded that the introduction of oxygen vacancy has little influence on the surface morphology and size of WO3, except narrowing its band gap and weakening its crystal structures.

3.2. Photoelectrochemistry (PEC) Properties

PEC measurements were conducted in a three-electrode electrochemical cell in 0.1 M Na2SO4 solution (pH: ca. 7.0). The photocurrent density is measured at 0.6 V and the result is shown in Figure 4a. It illustrates that WO3−x samples can produce higher photocurrent density than pristine WO3, as reflected in the i-t patterns with on–off recycles of intermittent irradiation. This suggests that the introduction of oxygen vacancy, to some extent, can improve the photo-irradiated charge transfer efficiency of WO3. The improvement is considered to come from WO3/WO3−x heterojunction formed by the energy difference between WO3 and the oxygen-vacancy-rich layer [35]. Among all WO3−x samples, WO3−x-3 h exhibits the highest photocurrent density, which is about triple that of pristine WO3 at 0.6 V vs. RHE. Nevertheless, the photocurrent density of WO3−x samples decreases when the oxygen vacancies exceed the optimal proportion, indicating that the further increase in oxygen vacancies in WO3−x is not necessary.
A similar trend has been observed in EIS, which is the common analytical tool to estimate charge separation and migration. As displayed in Figure 4b, four WO3−x samples exhibit a smaller arc radius than pristine WO3, verifying that the existence of oxygen vacancy can reduce interfacial charge transport resistance, thus leading to more effective charge separation. The charge transport resistance reached the minimum value when WO3 had been thermally treated for 3 h. A trend from rising then declining generally means the charge transfer efficiency of WO3−x samples suffers from two opposite (positive and negative) influence factors. Based upon the result of XRD (Figure 3), the negative influence may arise from the crystalline destruction when the samples have excessive oxygen vacancies. As a result, the optimal proportion of oxygen vacancies appeared when WO3 was thermal treated for 3 h and, thus, WO3−x-3 h presents the best electrical performance.
The Mott–Schottky plots (Figure 4c) of pristine and WO3−x samples are plotted according to ref [30,36,37]. The flat-band potentials are calculated to be 0.03, 0.09, 0.13, 0.07 and 0.06 V vs. NHE for WO3, WO3−x-0.5 h, WO3−x-3 h, WO3−x-6 h and WO3−x-10 h, respectively. As their Eg were 2.77, 2.71, 2.67, 2.62 and 2.53 eV, their VB can be calculated according to the equation of Eg = EVBECB. The detail of band structures is illustrated in Figure 4d.

3.3. Photocatalytic Degradation of CBZ

The photocatalytic degradation of 5 ppm CBZ by as-prepared samples is investigated, and results are shown in Figure 5 and Figure 6 and photocatalytic degradation efficiency (%) can be calculated by Equation (1) [38]. Here, Ct is the concentration of the pollutants at the irradiation time t and C0 is the concentration at the equilibrium before irradiation, respectively.
Degradation efficiency (%) = (1 − Ct/C0) × 100%
As showed in Figure 5, only 36% CBZ has been removed by pristine WO3 after 100 min, while those of the WO3−x samples ranged from 49% to 94%. Obviously, the degradation efficiencies on WO3−x samples are remarkably higher than that of pristine WO3. Here, the kinetic constant (k) for CBZ is further estimated by the following equation [38]:
k = −ln(Ct/C0)/t
The plots of −ln(Ct/C0) vs. t are found to be a linear relationship, clarifying that the photocatalytic degradation reaction of CBZ can be described by using a pseudo-first-order model (Figure S1). From the kinetic curves, the kinetic constant k values on WO3 are calculated to 0.0060 min−1, while those of WO3−x-0.5 h, WO3−x-3 h, WO3−x-6 h and WO3−x-10 h are increased to 0.0106, 0.0340, 0.0212 and 0.0081 min−1, respectively (Figure 6a). Hence, the k value on WO3−x-3 h was the largest and 5.67 times that of on WO3, suggesting that the introduction of oxygen vacancy can significantly improve the photocatalytic degradation performance of WO3. However, the photocatalytic performance of WO3−x samples decreased when the oxygen vacancies exceeded the optimal proportion (WO3−x-3 h). The improved photocatalytic performance is considered to originate from the promotion of the dissociation of excitons and accelerates the transfer of charges.

3.4. Optimization of Oxygen Vacancy

The activity of WO3−x can also be enhanced by optimizing the oxygen vacancy location distribution. As mentioned above, the thermal treatment causes oxygen vacancies both in bulk and on surface and leads to higher activity. Interestingly, it was found that the sample would exhibit even higher activity when it was placed in the air for a period of time. This interesting phenomenon can presumably be attributed to the recovery of the surface oxygen vacancies. It may produce bulk oxygen vacancy dominating WO3−x samples, along with a decrease in surface oxygen vacancies, and, hence, these were named as bulk-WO3−x. The degradation efficiency of bulk-WO3−x is higher and their kinetic constant increases to 0.0271, 0.0628, 0.0284 and 0.0208 min−1. They are 1.34~2.57 times that of WO3−x just after thermal treatment. This indicates oxygen exposure of the thermally treated sample will exhibit higher activity as compared.
The surface oxygen vacancies can be investigated by XPS. Figure 6b–d show the O 1s XPS spectra of WO3, WO3−x (represented by WO3−x-3 h) and bulk-WO3−x (represented by bulk-WO3−x-3 h) samples, which were deconvoluted in three peaks. The peaks at about 530.2, 531.7 and 533.2 eV correspond to surface lattice oxygen species (OL), hydroxyl species (OOH) and adsorbed water species (OH2O), respectively. It is considered that the percentage of OOH signal links to the quantity of surface oxygen vacancies [9,24,39]. The percentages of the three oxygen species of WO3, WO3−x and bulk-WO3−x samples are listed in Table 1. Their percentage of OOH increases obviously from 11.8% for WO3 to 15.3% for bulk-WO3−x, suggesting the thermal treatment causes surface oxygen vacancies.
Moreover, when the WO3−x sample was placed in the air for a period of time, the density of surface oxygen vacancies decreases apparently. As shown in Table 1, the percentage of OOH decreases from 15.3% to 11.3%. This demonstrates that many surface oxygen vacancies are recovered. It is reported that the surface oxygen vacancies will cause a shift of VB position [9,25,26]. However, the VB positions of WO3 and bulk-WO3−x samples are both 2.80 eV (Figure 6f), without any shift in VB positions between the two samples. It demonstrates that much of the surface oxygen vacancies are recovered and the density of surface oxygen vacancies in bulk-WO3−x sample has become similar to that of pristine WO3. Meanwhile, EPR spectra indicate the formation of bulk oxygen vacancies (Figure 6e). The EPR signal can still be observed on bulk-WO3−x sample but there is no signal in the pristine WO3 sample. It indicates the existence of bulk oxygen vacancy in the bulk-WO3−x sample.
Therefore, it can be concluded that the activity of samples increases when both bulk and surface oxygen vacancies are presented, while the activity is further increased when the surface oxygen vacancies are recovered. The procedures of preparation from pristine WO3 to WO3−x and bulk-WO3−x. are exhibited as Figure 7.

3.5. Mechanism of Photocatalytic Degradation

Inspired by this interesting phenomenon, the bulk oxygen vacancy dominating samples is considered as the optimal one for photocatalytic degradation. Subsequently, the mechanism for the degradation is studied.
To investigate the contribution of the various oxidative species generated in the photocatalytic system of WO3 and bulk-WO3−x samples for the degradation of CBZ, active species are tested by EPR. The generation of reactive oxygen species in the system are monitored though EPR technique after adding DMPO or TEMP as the trapping agent under UV-vis radiation. As shown in Figure 8a, a four-line EPR signal of DMPO-·OH adduct with an intensity ratio of 1:2:2:1 is also detected in both of WO3 and bulk-WO3−x systems, and the intensity of DMPO-·OH adduct signal in the system with WO3−x is much stronger than that of WO3. It indicates oxygen vacancy is significantly conducive to the generation of ·OH. As the VB of WO3 and WO3−x are both 2.80 eV, more positive than the standard redox potential of H2O/·OH (2.37 V) [40], both samples can oxidize H2O to ·OH under UV-vis thermodynamically. Therefore, ·OH can be generated from the oxidation of H2O directly (h+ + H2O → ·OH + H+). Then, methanol is added during the EPR measurement to catch ·OH and four very weak peaks with an intensity ratio of 1:1:1:1 are detected in bulk-WO3−x with oxygen vacancies (Figure 8b). It suggests that ·O2− is generated in bulk-WO3−x systems, as they are the characteristic quartet signal of DMPO-·O2− adduct. Here, ·O2− may be generated by the reduction of O2 (e + O2 → ·O2−). However, these characteristic signals cannot be observed in WO3 systems without oxygen vacancies. Therefore, this strongly indicates that the introduction of oxygen vacancies is favorable for the formation of ·O2−, which is beneficial to the degradation efficiency. In addition, there is no signal of TEMP-1O2 adduct detected in WO3 nor bulk-WO3−x systems (Figure 8c).
In addition, active species trapping tests are also executed by adding KI and K2Cr2O7 as the quenchers of photoexcited hole (h+) and e, respectively. When 10 mM KI is added into the WO3 system, the degradation efficiency of CBZ remarkably decreases to 66.2% of pristine value (Figure 8d). It suggests that photoexcited hole (h+) is one of the major reactive species contributing to the catalytic degradation of CBZ by WO3. However, the addition of KI in the bulk-WO3−x does not influence the degradation efficiency. This means that the hole does not directly cause the degradation of CBZ; instead, the hole mainly oxidized water into OH radicals for the degradation.
Meanwhile, the degradation efficiency of CBZ remarkably increase to 197.6% as compared with pristine value when 10 mM K2Cr2O7 is added. The increase may originate from the promotion of charge separation due to quenching of e. In comparison, the degradation efficiency of CBZ in the bulk-WO3−x photocatalytic system has little change by adding 10 mM K2Cr2O7. This should be due to the high charge transfer efficiency of WO3−x resulting in insignificant enhancement caused by the introduction of K2Cr2O7.
Based on the above analysis, the bulk oxygen vacancy has optimized the photodegradation mechanism of CBZ by WO3. The oxygen-vacancy-free sample mainly oxidizes the CBZ by the photoexcited hole (h+) (Figure 9a), requiring the CBZ to be adsorbed and cause low activity. By contrast, the bulk oxygen vacancy dominating WO3 mainly produces ·OH or ·O2− to degrade CBZ; the stronger oxidization power of radicals and their ability to migrate to the solution is beneficial to the degradation of CBZ (Figure 9b).

4. Conclusions

In summary, introduction of oxygen vacancy can improve photoelectric properties and photocatalytic degradation performance of WO3 samples. The photocurrent density and degradation efficiency of WO3−x increase firstly, followed by diminishing, along with an increase in the proportion of oxygen vacancy. The bulk and surface oxygen vacancies are produced in WO3−x just after thermal treatment. It is found that surface oxygen vacancies would be recovered when WO3−x was oxygen-exposed, resulting in bulk oxygen vacancies dominating WO3−x, whose photocatalytic degradation performance even shows further improvement. Therefore, both proportion and location distribution of oxygen vacancy could exert different influences on the photocatalytic performance of WO3−x. The recovered behavior of surface oxygen vacancy and more essential influence of bulk oxygen vacancy deserve to be further studied. The CBZ degradation mechanism was compared between WO3 and bulk oxygen vacancies dominating WO3−x. The results suggest that improvement in photocatalytic degradation may be due to the generation of OH radicals and superoxide radicals. The results of this study could potentially broaden our understanding of the role of oxygen vacancies and provide optimal directions and methods for oxygen vacancy regulation for photocatalysts.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano14110923/s1, Figure S1: The pseudo first-order kinetic fitting curves of as-prepared samples; Figure S2: The pseudo first-order kinetic fitting curves of WO3−x and bulk-WO3−x.

Author Contributions

W.G., F.W. and Z.H. conceived and designed the review article; W.G. wrote the first draft of the manuscript; F.W. and Z.H. reviewed and edited; Q.W. and G.L. helped with the data analysis. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by Key-Area Research and Development Program of Guangdong Province (2021B0909060001), Guangdong high level Innovation Research Institute (2021B0909050001), Youth Foundation of China GRINM Group Corporation Limited (Y20233102133), and the Guangzhou Basic and Applied Basic Research Foundation, China (202201011695).

Data Availability Statement

The data presented in this study are available on request from the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhao, T.; Xing, Z.; Xiu, Z.; Li, Z.; Chen, P.; Zhu, Q.; Zhou, W. Synergistic effect of surface plasmon resonance, Ti3+ and oxygen vacancy defects on Ag/MoS2/TiO2−x ternary heterojunctions with enhancing photothermal catalysis for low-temperature wastewater degradation. J. Hazard. Mater. 2019, 364, 117–124. [Google Scholar] [CrossRef] [PubMed]
  2. Zhao, H.; Liu, X.; Dong, Y.; Xia, Y.; Wang, H. A special synthesis of BiOCl photocatalyst for efficient pollutants removal: New insight into the band structure regulation and molecular oxygen activation. Appl. Catal. B Environ. 2019, 256, 117872. [Google Scholar] [CrossRef]
  3. Wang, H.; Zhang, L.; Chen, Z.; Hu, J.; Li, S.; Wang, Z.; Liu, J.; Wang, X. Semiconductor heterojunction photocatalysts: Design, construction, and photocatalytic performances. Chem. Soc. Rev. 2014, 43, 5234–5244. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, W.; Li, G.; Liu, H.; Chen, J.; Ma, S.; An, T. Micro/nano-bubble assisted synthesis of Au/TiO2@CNTs composite photocatalyst for photocatalytic degradation of gaseous styrene and its enhanced catalytic mechanism. Environ. Sci. Nano 2019, 6, 948–958. [Google Scholar] [CrossRef]
  5. Feng, Y.; Li, H.; Ling, L.; Yan, S.; Pan, D.; Ge, H.; Li, H.; Bian, Z. Enhanced Photocatalytic Degradation Performance by Fluid-Induced Piezoelectric Field. Environ. Sci. Technol. 2018, 52, 7842–7848. [Google Scholar] [CrossRef] [PubMed]
  6. Han, N.; Wang, S.; Yao, Z.; Zhang, W.; Zhang, X.; Zeng, L.; Chen, R. Superior three-dimensional perovskite catalyst for catalytic oxidation. EcoMat 2020, 2, e12044. [Google Scholar] [CrossRef]
  7. Jiang, M.; Zhang, M.; Wang, L.; Fei, Y.; Wang, S.; Núñez-Delgado, A.; Bokhari, A.; Race, M.; Khataee, A.; Klemeš, J.J.; et al. Photocatalytic degradation of xanthate in flotation plant tailings by TiO2/graphene nanocomposites. Chem. Eng. J. 2022, 432, 134104. [Google Scholar] [CrossRef]
  8. Kim, J.; Lee, C.; Choi, W. Platinized WO3 as an Environmental Photocatalyst that Generates OH Radicals under Visible Light. Environ. Sci. Technol. 2010, 44, 6849–6854. [Google Scholar] [CrossRef] [PubMed]
  9. Wang, Y.; Cai, J.; Wu, M.; Chen, J.; Zhao, W.; Tian, Y.; Ding, T.; Zhang, J.; Jiang, Z.; Li, X. Rational construction of oxygen vacancies onto tungsten trioxide to improve visible light photocatalytic water oxidation reaction. Appl. Catal. B Environ. 2018, 239, 398–407. [Google Scholar] [CrossRef]
  10. Li, Y.; Tang, Z.; Zhang, J.; Zhang, Z. Defect Engineering of Air-Treated WO3 and Its Enhanced Visible-Light-Driven Photocatalytic and Electrochemical Performance. J. Phys. Chem. C 2016, 120, 9750–9763. [Google Scholar] [CrossRef]
  11. Mi, Q.; Zhanaidarova, A.; Brunschwig, B.; Gray, H.B.; Lewis, N.S. A quantitative assessment of the competition between water and anion oxidation at WO3 photoanodes in acidic aqueous electrolytes. Energy Environ. Sci. 2012, 5, 5694–5700. [Google Scholar] [CrossRef]
  12. Wu, J.; Qiao, P.; Li, H.; Ren, L.; Xu, Y.; Tian, G.; Li, M.; Pan, K.; Zhou, W. Surface-oxygen vacancy defect-promoted electron-hole separation of defective tungsten trioxide ultrathin nanosheets and their enhanced solar-driven photocatalytic performance. J. Colloid Interface Sci. 2019, 557, 18–27. [Google Scholar] [CrossRef] [PubMed]
  13. Koo, B.; Ahn, H. Fast-switching electrochromic properties of mesoporous WO3 films with oxygen vacancy defects. Nanoscale 2017, 9, 17788–17793. [Google Scholar] [CrossRef] [PubMed]
  14. Petruleviciene, M.; Juodkazyte, J.; Parvin, M.; Tereshchenko, A.; Ramanavicius, S.; Karpicz, R.; Samukaite-Bubniene, U.; Ramanavicius, A. Tuning the Photo-Luminescence Properties of WO3 Layers by the Adjustment of Layer Formation Conditions. Materials 2020, 13, 2814. [Google Scholar] [CrossRef]
  15. Shi, W.; Li, H.; Chen, J.; Lv, X.; Shen, Y. Hierarchical WO3 nanoflakes architecture with enhanced photoelectrochemical activity. Electrochim. Acta 2017, 225, 473–481. [Google Scholar] [CrossRef]
  16. Solarska, R.; Królikowska, A.; Augustyński, J. Silver Nanoparticle Induced Photocurrent Enhancement at WO3 Photoanodes. Angew. Chem. Int. Ed. 2010, 49, 7980–7983. [Google Scholar] [CrossRef]
  17. Liu, G.; Han, J.; Zhou, X.; Huang, L.; Zhang, F.; Wang, X.; Ding, C.; Zheng, X.; Han, H.; Li, C. Enhancement of visible-light-driven O2 evolution from water oxidation on WO3 treated with hydrogen. J. Catal. 2013, 307, 148–152. [Google Scholar] [CrossRef]
  18. Moulzolf, S.; Ding, S.; Lad, R. Stoichiometry and microstructure effects on tungsten oxide chemiresistive films. Sens. Actuators B Chem. 2001, 77, 375–382. [Google Scholar] [CrossRef]
  19. Solarska, R.; Królikowska, A.; Augustyński, J. Silver Nanoparticle Induced Photocurrent Enhancement at WO3 Photoanodes. Angew. Chem. Int. Ed. 2010, 122, 8152–8155. [Google Scholar] [CrossRef]
  20. Liu, Q.; Wang, F.; Lin, H.; Xie, Y.; Tong, N.; Lin, J.; Zhang, X.; Zhang, Z.; Wang, X. Surface oxygen vacancy and defect engineering of WO3 for improved visible light photocatalytic performance. Catal. Sci. Technol. 2018, 8, 4399–4406. [Google Scholar] [CrossRef]
  21. Long, R.; English, N. Various 3d-4f spin interactions and field-induced metamagnetism in the Cr5+ system DyCrO4. Phys. Rev. B 2011, 83, 024416. [Google Scholar] [CrossRef]
  22. Ansari, S.; Khan, M.; Kalathil, S.; Nisar, A.; Lee, J.; Cho, M.H. Oxygen vacancy induced band gap narrowing of ZnO nanostructures by an electrochemically active biofilm. Nanoscale 2013, 5, 9238–9246. [Google Scholar] [CrossRef] [PubMed]
  23. Linsebigler, A.; Lu, G. Photocatalysis on TiO2 Surfaces: Principles, Mechanisms, and Selected Results. Chem. Rev. 1995, 95, 735–758. [Google Scholar] [CrossRef]
  24. Cai, J.; Zhu, Y.; Liu, D.; Meng, M.; Hu, Z.; Jiang, Z. Synergistic Effect of Titanate-Anatase Heterostructure and Hydrogenation-Induced Surface Disorder on Photocatalytic Water Splitting. ACS Catal. 2015, 5, 1708–1716. [Google Scholar] [CrossRef]
  25. Schmid, M.; Shishkin, M.; Kresse, G.; Napetschnig, E.; Varga, P.; Kulawik, M.; Nilius, N.; Rust, H.-P.; Freund, H.-J. Oxygen-Deficient Line Defects in an Ultrathin Aluminum Oxide Film. Phys. Rev. Lett. 2006, 97, 046101. [Google Scholar] [CrossRef] [PubMed]
  26. Weinhardt, L.; Blum, M.; Bär, M.; Heske, C.; Cole, B.; Marsen, B.; Miller, E.L. Electronic Surface Level Positions of WO3 Thin Films for Photoelectrochemical Hydrogen Production. J. Phys. Chem. C 2008, 112, 3078–3082. [Google Scholar] [CrossRef]
  27. Ramanavicius, S.; Petruleviciene, M.; Juodkazyte, J.; Grigucevičienė, A.; Ramanavičius, A. Selectivity of Tungsten Oxide Synthesized by Sol-Gel Method Towards Some Volatile Organic Compounds and Gaseous Materials in a Broad Range of Temperatures. Materials 2020, 13, 523. [Google Scholar] [CrossRef] [PubMed]
  28. Zheng, X.; Dong, X.; Zhang, S.; Yang, J. Defect-induced ferromagnetism in W18O49 nanowires. J. Alloys Compd. 2020, 818, 152894. [Google Scholar] [CrossRef]
  29. Liu, Y.; Yao, W.; Lin, J. Improved photoelectrocatalytic performances over electrochemically reduced WO3: Implications of oxygen vacancies. Phys. Chem. Chem. Phys. 2023, 25, 15248–15256. [Google Scholar] [CrossRef]
  30. Hu, Z.; Xu, M.; Shen, Z.; Yu, J. A nanostructured chromium(III) oxide/tungsten(VI) oxide p–n junction photoanode toward enhanced efficiency for water oxidation. J. Mater. Chem. A 2015, 3, 14046–14053. [Google Scholar] [CrossRef]
  31. Zhang, T.; Zhao, K.; Yu, J.; Jin, J.; Qi, Y.; Li, H.; Hou, X.; Liu, G. Photocatalytic water splitting for hydrogen generation on cubic, orthorhombic, and tetragonal KNbO3 microcubes. Nanoscale 2013, 5, 8375–8383. [Google Scholar] [CrossRef]
  32. Pan, X.; Yang, M.; Fu, X.; Zhang, N.; Xu, Y.-J. Defective TiO2 with oxygen vacancies: Synthesis, properties and photocatalytic applications. Nanoscale 2013, 5, 3601–3614. [Google Scholar] [CrossRef] [PubMed]
  33. Li, W.; Da, P.; Zhang, Y.; Wang, Y.; Lin, X.; Gong, X.; Zheng, G. WO3 Nanoflakes for Enhanced Photoelectrochemical Conversion. ACS Nano 2014, 8, 11770–11777. [Google Scholar] [CrossRef] [PubMed]
  34. Zhao, C.; Kong, X.; Liu, X.; Tu, L.; Wu, F.; Zhang, Y.; Liu, K.; Zeng, Q.; Zhang, H. Li+ ion doping: An approach for improving the crystallinity and upconversion emissions of NaYF4:Yb3+, Tm3+ nanoparticles. Nanoscale 2013, 5, 8084–8089. [Google Scholar] [CrossRef] [PubMed]
  35. Zhang, R.; Ning, F.; Xu, S.; Zhou, L.; Shao, M.; Wei, M. Oxygen vacancy engineering of WO3 toward largely enhanced photoelectrochemical water splitting. Electrochim. Acta 2018, 274, 217–223. [Google Scholar] [CrossRef]
  36. Yang, X.; Wolcott, A.; Wang, G.; Sobo, A.; Fitzmorris, R.C.; Qian, F.; Zhang, J.Z.; Li, Y. Nitrogen-Doped ZnO Nanowire Arrays for Photoelectrochemical Water Splitting. Nano Lett. 2009, 9, 2331–2336. [Google Scholar] [CrossRef] [PubMed]
  37. Wolcott, A.; Smith, W.; Kuykendall, T.; Zhao, Y.; Zhang, J.Z. Photoelectrochemical Water Splitting Using Dense and Aligned TiO2 Nanorod Arrays. Small 2009, 5, 104–111. [Google Scholar] [CrossRef]
  38. Yang, Y.; Zhang, C.; Huang, D.; Zeng, G.; Huang, J.; Lai, C.; Zhou, C.; Wang, W.; Guo, H.; Xue, W.; et al. Boron nitride quantum dots decorated ultrathin porous g-C3N4: Intensified exciton dissociation and charge transfer for promoting visible-light-driven molecular oxygen activation. Appl. Catal. B Environ. 2019, 245, 87–99. [Google Scholar] [CrossRef]
  39. Rahimnejad, S.; He, J.; Pan, F.; Lee, X.; Chen, W.; Wu, K.; Xu, G.Q. Enhancement of the photocatalytic efficiency of WO3 nanoparticles via hydrogen plasma treatment. Mater. Res. Express 2014, 1, 045044. [Google Scholar] [CrossRef]
  40. Ma, J.; Wang, C.; He, H. Enhanced photocatalytic oxidation of NO over g-C3N4-TiO2 under UV and visible light. Appl. Catal. B Environ. 2016, 184, 28–34. [Google Scholar] [CrossRef]
Figure 1. SEM images of (a,b) WO3; and WO3−x samples: (c) WO3−x-0.5 h; (d) WO3−x-3 h; (e) WO3−x-6 h; (f) WO3−x-10 h.
Figure 1. SEM images of (a,b) WO3; and WO3−x samples: (c) WO3−x-0.5 h; (d) WO3−x-3 h; (e) WO3−x-6 h; (f) WO3−x-10 h.
Nanomaterials 14 00923 g001
Figure 2. (a) The photographs; (b) UV-vis diffuse reflectance spectra; (c) band gaps calculated from Kubellka–Munk plots of the samples.
Figure 2. (a) The photographs; (b) UV-vis diffuse reflectance spectra; (c) band gaps calculated from Kubellka–Munk plots of the samples.
Nanomaterials 14 00923 g002
Figure 3. (a) XRD patterns of WO3, WO3−x-0.5 h, WO3−x-3 h, WO3−x-6 h and WO3−x-10 h; (b) their XRD patterns on (002), (020) and (200) planes and (c) on (020) and (200) planes.
Figure 3. (a) XRD patterns of WO3, WO3−x-0.5 h, WO3−x-3 h, WO3−x-6 h and WO3−x-10 h; (b) their XRD patterns on (002), (020) and (200) planes and (c) on (020) and (200) planes.
Nanomaterials 14 00923 g003
Figure 4. (a) Transient photocurrent responses; (b) EIS spectra; (c) Mott–Schottky plots; and (d) band structure diagram of as-prepared samples.
Figure 4. (a) Transient photocurrent responses; (b) EIS spectra; (c) Mott–Schottky plots; and (d) band structure diagram of as-prepared samples.
Nanomaterials 14 00923 g004
Figure 5. Photocatalytic degradation curve of as-prepared samples for CBZ under UV-vis.
Figure 5. Photocatalytic degradation curve of as-prepared samples for CBZ under UV-vis.
Nanomaterials 14 00923 g005
Figure 6. (a) Kinetic constant (k) of as-prepared samples for CBZ under UV-vis; O 1s XPS spectra of (b) WO3, (c) WO3−x and (d) bulk-WO3−x; (e) EPR spectra and (f) VB XPS spectra of WO3 and bulk-WO3−x.
Figure 6. (a) Kinetic constant (k) of as-prepared samples for CBZ under UV-vis; O 1s XPS spectra of (b) WO3, (c) WO3−x and (d) bulk-WO3−x; (e) EPR spectra and (f) VB XPS spectra of WO3 and bulk-WO3−x.
Nanomaterials 14 00923 g006
Figure 7. The schema of preparation from pristine WO3 to WO3−x and bulk-WO3−x.
Figure 7. The schema of preparation from pristine WO3 to WO3−x and bulk-WO3−x.
Nanomaterials 14 00923 g007
Figure 8. EPR spectra of (a) DMPO-·OH adduct; (b) DMPO-·O2− adduct and (c) TEMP-1O2 for WO3 and bulk-WO3−x under UV-vis irradiation; (d) the changes in degradation efficiencies of CBZ in WO3 and bulk-WO3−x photocatalytic systems with KI or K2Cr2O7 as quenchers.
Figure 8. EPR spectra of (a) DMPO-·OH adduct; (b) DMPO-·O2− adduct and (c) TEMP-1O2 for WO3 and bulk-WO3−x under UV-vis irradiation; (d) the changes in degradation efficiencies of CBZ in WO3 and bulk-WO3−x photocatalytic systems with KI or K2Cr2O7 as quenchers.
Nanomaterials 14 00923 g008
Figure 9. Illustration of photocatalytic mechanism of (a) WO3 and (b) bulk-WO3−x.
Figure 9. Illustration of photocatalytic mechanism of (a) WO3 and (b) bulk-WO3−x.
Nanomaterials 14 00923 g009
Table 1. Detailed results of O 1s XPS spectra of pristine WO3, OV-WO3-3 h and bulk-OV-WO3.
Table 1. Detailed results of O 1s XPS spectra of pristine WO3, OV-WO3-3 h and bulk-OV-WO3.
TotalOLOOHOH2O
AreaWO384,376.66272,615.11010,022.6001738.952
WO3−x174,771.533142,919.30026,817.0805035.153
bulk-WO3−x78,938.52064,014.5208907.0126016.988
PercentageWO3100%86.1%11.8%2.1%
WO3−x100%81.8%15.3%2.9%
bulk-WO3−x100%81.1%11.3%7.6%
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

Guo, W.; Wei, Q.; Li, G.; Wei, F.; Hu, Z. A Bulk Oxygen Vacancy Dominating WO3−x Photocatalyst for Carbamazepine Degradation. Nanomaterials 2024, 14, 923. https://doi.org/10.3390/nano14110923

AMA Style

Guo W, Wei Q, Li G, Wei F, Hu Z. A Bulk Oxygen Vacancy Dominating WO3−x Photocatalyst for Carbamazepine Degradation. Nanomaterials. 2024; 14(11):923. https://doi.org/10.3390/nano14110923

Chicago/Turabian Style

Guo, Weiqing, Qianhui Wei, Gangrong Li, Feng Wei, and Zhuofeng Hu. 2024. "A Bulk Oxygen Vacancy Dominating WO3−x Photocatalyst for Carbamazepine Degradation" Nanomaterials 14, no. 11: 923. https://doi.org/10.3390/nano14110923

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