Cs-Doped WO3 With Enhanced Conduction Band For Eff
Cs-Doped WO3 With Enhanced Conduction Band For Eff
Cs-Doped WO3 With Enhanced Conduction Band For Eff
1 School of Materials Science & Engineering, North Minzu University, Yinchuan 750021, China;
[email protected] (S.T.); [email protected] (Q.Q.); [email protected] (H.S.);
[email protected] (F.H.)
2 National and Local Joint Engineering Research Center of Advanced Carbon-Based Ceramics Preparation
[email protected]
* Correspondence: [email protected]
Abstract: Cesium doped WO3 (Cs-WO3) photocatalyst with high and stable oxidation activity was
successfully synthesized by a one-step hydrothermal method using Cs2CO3 as the doped metal ion
source and tungstic acid (H2WO4) as the tungsten source. A series of analytical characterization tools
and oxygen precipitation activity tests were used to compare the effects of different additions of
Cs2CO3 on the crystal structure and microscopic morphologies. The UV–visible diffuse reflectance
spectra (DRS) of Cs-doped material exhibited a significant red shift in the absorption edge with new
shoulders appearing at 440–520 nm. The formation of an oxygen vacancy was confirmed in Cs-WO3
by the EPR signal, which can effectively regulate the electronic structure of the catalyst surface and
contribute to improving the activity of the oxygen evolution reaction (OER). The photocatalytic OER
results showed that the Cs-WO3-0.1 exhibited the optimal oxygen precipitation activity, reaching
58.28 µmol at 6 h, which was greater than six times higher than that of WO3-0 (9.76 µmol). It can be
Citation: Li, D.; Tian, S.; Qian, Q.; attributed to the synergistic effect of the increase in the conduction band position of Cs-WO3-0.1
Gao, C.; Shen, H.; Han, F. (0.11 V) and oxygen vacancies compared to WO3-0, which accelerate the electron conduction rate
Cs-Doped WO3 with Enhanced and slow down the rapid compounding of photogenerated electrons–holes, improving the water-
Conduction Band for Efficient catalytic oxygen precipitation activity of WO3.
Photocatalytic Oxygen Evolution
Reaction Driven by Keywords: Cs-doped; tungsten trioxide; oxygen vacancy; band energy; oxygen evolution reaction
Long-Wavelength Visible Light.
Molecules 2024, 29, 3126. https://
doi.org/10.3390/molecules29133126
Cs2W6O19 compound began to be seen in the Cs-WO3-0.5 alongside the main peaks of the
cubic phase WO3·0.5H2O while overdoping with Cs+ ions. In Figure 1B,C, obvious shifts
of XRD peaks are observed, which are ascribed to the existence of structural strain [34].
Additionally, compared with the WO3-0, the crystallinity of the Cs-WO3 samples de-
creased with increasing the doping concentration of Cs ions. This may be due to the dis-
ruption of the crystal structure of WO3, when the Cs ions with larger ionic radii (1.67 Å)
are doped into the WO3 lattice, resulting in the formation of a distorted structure.
Figure 1. The XRD patterns of Cs-WO3-0.1 samples prepared at (A) different temperatures and (B)
different reaction times. (C) The XRD patterns of the pure WO3 and Cs-WO3 samples prepared with
different doping concentrations of Cs+ ions.
The average crystal grain size was calculated from the (311) peak (2θ = 28.9˚) of cubic
WO3·0.5H2O according to Scherrer’s equation [35], as shown in Table 1. The crystallite di-
ameter of the WO3-0 (23.6 nm) was larger than those of Cs-WO3-0.1 (22.6 nm), Cs-WO3-0.3
(20.9 nm) and Cs-WO3-0.5 (19.1 nm). The calculated lattice parameters and average crys-
talline size of WO3 were decreased (Table 1) by Cs doping. The results indicate that Cs-
doped WO3 has a strong restraining function with the increase in crystal size owning to
the dopant cations Cs+ preventing the growth of crystal grains in the nanoparticles [36].
tallite diameters were calculated from XRD data according to the Scherrer equation and expressed
as average values calculated based on the (311) peak. (c) The surface areas were provided from N2
sorption isotherms.
As shown in Figure 2a, the elliptical-like WO3-0 is composed of nanosheets and na-
noparticles with a major axis of about 55 µm and a semimajor axis of 40 µm. Notably, the
Cs-WO3-0.1 is made up of a microsphere of about 41 µm in diameter which mainly con-
sists of nanoparticles and nanorods (Figure 2b). The diameter of the microsphere (43 µm)
Molecules 2024, 29, 3126 4 of 13
increased and the amount of nanorods gradually decreased with the increasing doping
amount of Cs+ (Figure 2c). A smooth microsphere with a diameter of 45 µm was formed
for the Cs-WO3-0.5 (Figure 2d), suggesting that the formation of micrometer spheres
mainly relies on self-assembly effects. As shown in Table 1, it was found that the specific
surface area of the sample strongly depends on its morphology. The Cs-WO3-0.1 pos-
sessed a higher specific surface area (16.1 m2/g) relative to those of WO3-0 (9.2 m2/g) and
Cs-WO3-0.3 (10.6 m2/g) owing to the formation of nanorods on the surface of the micro-
sphere, which is beneficial to form more accessible active sites for the OER.
Figure 2. The SEM images of (a) WO3-0, (b) Cs-WO3-0.1, (c) Cs-WO3-0.3, (d) Cs-WO3-0.5.
The elemental maps of the EDX for the Cs-WO3-0.1 are shown in Figure 3. The map-
ping signals of the W and O (Figure 3c,d) and of the Cs (Figure 3e) were detected. The
results showed a uniform distribution of Cs, O and W, in good chemical agreement with
Cs-WO3.
Figure 3. The SEM_EDX element distribution mapping images of the (a) Cs-WO3-0.1; (b) layered W,
O and Cs; (c) W; (d) O; (e) Cs.
The chemical composition and valence states of Cs-WO3 samples were investigated
by XPS. The spectra were calibrated with the C 1s peak as reference. As shown in Figure
4A, the XPS survey spectrum of WO3-0 (a) depicts that no other impurity phases were
detected except W and O elements. W, O and Cs elements were co-present in the Cs-WO3-
0.1 (b) and Cs-WO3-0.3 (c). In Figure 4B, the high-resolution XPS spectrum of W 4f exhibits
two peaks at 37.7 eV and 35.5 eV that are, respectively, ascribed to the spin–orbit doublet
Molecules 2024, 29, 3126 5 of 13
of W 4f5/2 and W 4f7/2, respectively, for a W6+ state in WO3 [37]. The XPS spectra of W 4f for
Cs-WO3-0.1 and Cs-WO3-0.3 (Figure 3b) exhibit two characteristic peaks at 38.1 eV and
35.9 eV, corresponding to 4f5/2 and W 4f7/2 of the WO3 lattice, respectively. As shown in
Figure 3c, the binding energy at 531.5 eV and 530.4 eV in the XPS spectrum of O 1s for the
WO3-0 can be assigned to the H2O and lattice oxygen, respectively [38]. The main peaks in
the high-resolution O 2p spectra for the Cs-WO3-0.1 and Cs-WO3-0.3 located at 530.8 eV
can be assigned to the lattice oxygen of the W-O bond in the crystalline WO3. The banding
energies at 531.9 eV correspond to the adsorbed oxygen ions and hydroxyl groups on the
surface, respectively. The ratios of adsorbed O increased to 12.7% and 13.5% in Cs-WO3-
0.1 and Cs-WO3-0.3 due to the surface oxygen vacancies, while it was 10.3% in WO3. This
can serve as indirect evidence for the presence of oxygen vacancies. The high-resolution
XPS spectra of the Cs element for both Cs-WO3-0.1 and Cs-WO3-0.3 (Figure 3d) exhibited
two peaks at 738.4 eV and 724.7 eV, which are assigned to the spin orbits of Cs 3d3/2 and
Cs 3d5/2, respectively [32,39]. Notably, the positive shifts of 0.4 eV and 0.8 eV in W 4f and
O 1s for Cs-WO3 can be seen after Cs+ doping. This is due to the occurrence of ion exchange
on the WO3 surface, which also confirmed that the Cs was successfully introduced into
the lattice of WO3 and formed W-Cs bonds in the doped samples The positive shifts im-
prove the PEC performance of WO3. Furthermore, it can be seen that the peak intensities
of W 4f and O 1s for Cs-WO3 decreased with increasing the doped contents of Cs element
because of the replacement of W6+ by Cs+ leading to the contamination of oxygen. Gener-
ally, the larger the ionic radius is, the more difficult it is for doping to occur due to the
requirement of high formation energy. Therefore, the replacement of W6+ by Cs+ is more
favorable than replacing O2- with Cs+.
Figure 4. (A) The XPS survey spectra of (a) WO3-0, (b) Cs-WO3-0.1 and (c) Cs-WO3-0.3; (B) W 4f; (C)
O 2p; and (D) Cs 3d regions.
In situ electron paramagnetic resonance (EPR) measurements were carried out to fur-
ther prove the existence of surface oxygen vacancies and investigate their properties. No
signal was detected for the WO3-0; however, Cs-WO3 samples exhibited relatively stronger
EPR peaks intensity at g ≈ 2.002 under the same conditions, as shown in Figure 5. The
higher the doped amounts of Cs, the stronger the signal intensity. These results agreed
with results reported previously, confirming the presence of surface oxygen vacancies on
Cs-WO3 samples [40,41]. The oxygen vacancies may be beneficial to improve the photo-
catalytic performance of Cs-WO3.
Molecules 2024, 29, 3126 6 of 13
Figure 5. The EPR spectra of (a) WO3-0, (b) Cs-WO3-0.1 and (c) Cs-WO3-0.3 at 103 K in liquid N2.
Figure 6. (A) UV–visible DRS and (B) Tacu plots based on UV–visible DRS of (black) WO3-0, (red)
Cs-WO3-0.1 and (blue) Cs-WO3-0.3.
Molecules 2024, 29, 3126 7 of 13
Figure 7. (A) Photocatalytic O2 evolution over (black) WO3-0, (red) Cs-WO3-0.1 and (blue) Cs-WO3-
0.3 (10 mg of catalysts added into 30mL solution under visible-light irradiation.); (B) photocatalytic
O2 production stability of Cs-WO3-0.1.
Figure 8. Linear sweep voltammograms (LSVs) of the (black) WO3-0, (red) Cs-WO3-0.1 and (blue)
Cs-WO3-0.3 electrodes with visible-light irradiation chopped in a 0.1 M phosphate-buffered solution
of pH 6.0 with visible-light irradiation (λ > 420 nm, 100 mW cm−2).
Table 2. Summary of optical and electrochemical properties and energies of band structures of var-
ious WO3 samples.
The Tafel plots are useful to investigate the reaction kinetics of the OER. As shown in
Figure 9B, the Tafel slopes of the Cs-WO3 prominently decrease compared with those of
WO3-0. Cs-WO3-0.1 possesses a lower Tafel slope of 16.46 mVdec−1 than Cs-WO3-0.3 (32.78
mVdec−1) and WO3-0 (48.54 mVdec−1), indicating that Cs doping gives a faster kinetic re-
sponse in the OER and makes the Cs-WO3 catalysts have higher photocatalytic activities
for the OER.
The electrochemical impedance was utilized to give an insight into the kinetics of the
charge transfer process and to evaluate its effect on photocatalytic O2 evolution activity.
As seen in the results of the Nyquist plots in Figure 9C, Cs-WO3-0.1 exhibited smaller
Molecules 2024, 29, 3126 9 of 13
semicircles than WO3-0 and Cs-WO3-0.3. As is well known, the diameter of the semicircle
in the Nyquist plot corresponds to the impedance of the electrode, and the larger the ra-
dius, the larger the impedance [46]. This result indicates that the Cs-WO3-0.1 has a lower
charge transfer resistance and higher separation efficiency for photogenerated electron–
hole pairs than other electrodes and inhibits the recombination of photogenerated
charges. This is mainly due to the n-type doping of WO3 by cesium doping, which injects
electrons into the Fermi level, enhances the CB potential and also increases lattice defects
and oxygen vacancies in WO3, thereby improving the conductivity of the WO3. The Tafel
and electrochemical impedance results provide favorable evidence for the improvement
of photocatalytic activity for the OER.
Figure 9. (A) Mott–Schottky plots of the (black) WO3-0, (red) Cs-WO3-0.1 and (blue) Cs-WO3-0.3
electrodes in a 0.1 M phosphate-buffered solution of pH, 6.0; frequency, 0.1 Hz; amplitude potential,
10 mV. (B) Tafel plots and (C) Nyquist plots of the (black) WO3-0, (red) Cs-WO3-0.1 and (blue) Cs-
WO3-0.3 electrodes for photoelectrocatalytic water oxidation in a 0.1 M phosphate-buffered solution
(pH = 6).
The energy positions were investigated to elucidate Cs-doping effects on the band
energy of WO3. Figure 10 shows energy positions for WO3-0, Cs-WO3-0.1 and Cs-WO3-0.3.
It has been well documented that the VB of Cs-WO3 consists of the hybridization between
O2p and Cs 4s, and the CB is from W 5d electronic components. The enhanced optical
absorption is illustrated for the contribution from the Cs 4s hybridization in VB.
Figure 10. The proposed band natures of WO3-0, Cs-WO3-0.1 and Cs-WO3-0.3.
Molecules 2024, 29, 3126 10 of 13
3. Experimental Section
3.1. Materials
Tungstic acid (H2WO4), Hydrogen peroxide (H2O2), Marpolose (60MP-50), Ethylene
glycol (EG, molecular weight = 300), Cs2CO3 and Fe2(SO4)3·9H2O were purchased from
Aladdin’s Reagent. A Fluorine-doped tin oxide (FTO)-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 used as received unless mentioned otherwise.
4. Conclusions
The Cs-doped WO3 with spatial charge separation was synthesized using a hydro-
thermal approach, which exhibited a significant enhancement of the photocatalytic per-
formance for water spitting to boost the production of oxygen. The addition of Cs depend-
ence on the physiochemical properties and the performance of the photocatalytic OER of
the WO3-0 and Cs-WO3 catalysts were investigated to characterize Cs doped into the WO3
lattice and reveal the mechanism of superior performance for the photocatalytic OER of
Cs-WO3. The Cs doping is responsible for the significant red shift in the absorption edge,
with a new shoulder appearing at 440–520 nm compared to that in WO3-0. The Cs-WO3
catalyst is able to utilize visible light at longer wavelengths below 520 nm for photocata-
lytic OER, in contrast to utilization below 440 nm for the WO3-0 catalyst. These results
demonstrate that Cs doping is an effective strategy for improving the photocatalytic per-
formance of WO3 photocatalysts for the OER, and thus it is expected to be applied for
photocatalytic OERs as artificial photosynthesis to improve the solar energy conversion
efficiency.
Author Contributions: Conceptualization, D.L., S.T. and F.H.; methodology, Q.Q., S.T., H.S. and
F.H.; investigation and data curation, Q.Q.; formal analysis, C.G.; supervision, D.L.; writing—orig-
inal draft preparation, D.L.; writing—review and editing, D.L. All authors have read and agreed to
the published version of the manuscript.
Funding: This project was supported by the National and Local Joint Engineering Research Center
of Advanced Carbon-Based Ceramics Preparation Technology, grant number 2024TJZX03. This pro-
ject was supported by the Natural Science Foundation of Ningxia Province, grant numbers
2023AAC03292; 2023AAC03306.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The original contributions presented in the study are included in the
article, further inquiries can be directed to the corresponding authors.
Acknowledgments: The authors thank the the National and Local Joint Engineering Research Cen-
ter of Advanced Carbon-Based Ceramics Preparation Technology (2024TJZX03) and the Natural Sci-
ence Foundation of Ningxia Province (Grant No. 2023AAC03292; 2023AAC03306).
Conflicts of Interest: The authors declare no conflicts of interest.
Molecules 2024, 29, 3126 12 of 13
References
1. Hisatomi, T.; Kubota, J.; Domen, K. Recent advances in semiconductors for photocatalytic and photoelectrochemical water split-
ting. Chem. Soc. Rev. 2014, 43, 7520–7535.
2. Zeng, F.; Mebrahtu, C.; Liao, L.; Beine, A.K.; Palkovits, R. Stability and deactivation of OER electrocatalysts: A review. J. Energy
Chem. 2022, 69, 301–329.
3. Bignozzi, C.A.; Caramori, S.; Cristino, V.; Argazzi, R.; Meda, L.; Tacca, A. Nanostructured photoelectrodes based on WO3: Ap-
plications to photooxidation of aqueous electrolytes. Chem. Soc. Rev. 2013, 42, 2228–2246.
4. Wang, L.; Lian, W.; Liu, B.; Lv, H.; Zhang, Y.; Wu, X.; Wang, T.; Gong, J.; Chen, T.; Xu, H. A transparent, high-performance, and
stable Sb2S3 photoanode enabled by heterojunction engineering with conjugated polycarbazole frameworks for unbiased pho-
toelectrochemical overall water splitting devices. Adv. Mater. 2022, 34, 2200723.
5. Ahmed, J.; Mao, Y. Ultrafine iridium oxide nanorods synthesized by molten salt method toward electrocatalytic oxygen and
hydrogen evolution reactions. Electrochim. Acta 2016, 212, 686–693.
6. Li, H.; Zhu, B.; Sun, J.; Gong, H.; Yu, J.; Zhang, L. Photocatalytic hydrogen production from seawater by TiO2/RuO2 hybrid
nanofiber with enhanced light absorption. J. Colloid Interface Sci. 2024, 654, 1010–1019.
7. Yang, M.; Li, J.; Ke, G.; Liu, B.; Dong, F.; Yang, L.; He, H.; Zhou, Y. WO3 homojunction photoanode: Integrating the advantages
of WO3 different facets for efficient water oxidation. J. Energy Chem. 2021, 56, 37–45.
8. Jiang, S.; Zhang, X.; Nawaz, M.; Fan, X.; Tao, R. Boosting the photoelectrochemical water splitting of Fe2O3 by surface-state
regulation. Inorg. Chem. Front. 2024, 11, 526–533.
9. del Olmo, L.; Dommett, M.; Oevreeide, I.H.; Walsh, A.; Di Tommaso, D.; Crespo-Otero, R. Water oxidation catalysed by quan-
tum-sized BiVO 4. J. Mater. Chem. A 2018, 6, 24965–24970.
10. He, Y.; Thorne, J.E.; Wu, C.H.; Ma, P.; Du, C.; Dong, Q.; Guo, J.; Wang, D. What limits the performance of Ta3N5 for solar water
splitting? Chem 2016, 1, 640–655.
11. Maeda, K.; Lu, D.; Domen, K. Solar-Driven Z-scheme water splitting using modified BaZrO3–BaTaO2N solid solutions as pho-
tocatalysts. Acs Catal. 2013, 3, 1026–1033.
12. Wang, Y.; Tian, W.; Chen, L.; Cao, F.; Guo, J.; Li, L. Three-dimensional WO3 nanoplate/Bi2S3 nanorod heterojunction as a highly
efficient photoanode for improved photoelectrochemical water splitting. ACS Appl. Mater. Interfaces 2017, 9, 40235–40243.
13. Liu, Y.; Wygant, B.R.; Mabayoje, O.; Lin, J.; Kawashima, K.; Kim, J.-H.; Li, W.; Li, J.; Mullins, C.B. Interface engineering and its
effect on WO3-based photoanode and tandem cell. ACS Appl. Mater. Interfaces 2018, 10, 12639–12650.
14. Chen, G.; Wang, Q.; Zhao, Z.; Gao, L.; Li, X. Synthesis and photocatalytic activity study of S-doped WO3 under visible light
irradiation. Environ. Sci. Pollut. Res. 2020, 27, 15103–15112.
15. Mi, Q.; Ping, Y.; Li, Y.; Cao, B.; Brunschwig, B.S.; Khalifah, P.G.; Galli, G.A.; Gray, H.B.; Lewis, N.S. Thermally stable N2-inter-
calated WO3 photoanodes for water oxidation. J. Am. Chem. Soc. 2012, 134, 18318–18324.
16. Kalanur, S.S.; Yoo, I.-H.; Eom, K.; Seo, H. Enhancement of photoelectrochemical water splitting response of WO3 by Means of
Bi doping. J. Catal. 2018, 357, 127–137.
17. Radecka, M.; Sobas, P.; Wierzbicka, M.; Rekas, M. Photoelectrochemical properties of undoped and Ti-doped WO3. Phys. B
Condens. Matter 2005, 364, 85–92.
18. Song, H.; Li, Y.; Lou, Z.; Xiao, M.; Hu, L.; Ye, Z.; Zhu, L. Synthesis of Fe-doped WO3 nanostructures with high visible-light-
driven photocatalytic activities. Appl. Catal. B Environ. 2015, 166, 112–120.
19. Song, Z.; Ma, J.; Sun, H.; Wang, W.; Sun, Y.; Sun, L.; Liu, Z.; Gao, C. Synthesis of NiWO4 nano-particles in low-temperature
molten salt medium. Ceram. Int. 2009, 35, 2675–2678.
20. Quyen, V.T.; Kim, J.; Park, P.-M.; Huong, P.T.; Viet, N.M.; Thang, P.Q. Enhanced the visible light photocatalytic decomposition
of antibiotic pollutant in wastewater by using Cu doped WO3. J. Environ. Chem. Eng. 2021, 9, 104737.
21. Cheng, X.; Leng, W.; Liu, D.; Zhang, J.; Cao, C. Enhanced photoelectrocatalytic performance of Zn-doped WO3 photocatalysts
for nitrite ions degradation under visible light. Chemosphere 2007, 68, 1976–1984.
22. Liew, S.; Zhang, Z.; Goh, T.G.; Subramanian, G.; Seng, H.D.; Hor, T.A.; Luo, H.-K.; Chi, D. Yb-doped WO3 photocatalysts for
water oxidation with visible light. Int. J. Hydrog. Energy 2014, 39, 4291–4298.
23. Mehtab, A.; Ahmed, J.; Alshehri, S.M.; Mao, Y.; Ahmad, T. Rare earth doped metal oxide nanoparticles for photocatalysis: A
perspective. Nanotechnology 2022, 33, 142001.
24. Sudrajat, H.; Babel, S. Rapid photocatalytic degradation of the recalcitrant dye amaranth by highly active N-WO 3. Environ.
Chem. Lett. 2016, 14, 243–249.
25. Sun, Y.; Murphy, C.J.; Reyes-Gil, K.R.; Reyes-Garcia, E.A.; Thornton, J.M.; Morris, N.A.; Raftery, D. Photoelectrochemical and
structural characterization of carbon-doped WO3 films prepared via spray pyrolysis. Int. J. Hydrog. Energy 2009, 34, 8476–8484.
26. Li, W.; Li, J.; Wang, X.; Chen, Q. Preparation and water-splitting photocatalytic behavior of S-doped WO3. Appl. Surf. Sci. 2012,
263, 157–162. https://doi.org/10.1016/j.apsusc.2012.09.021.
27. Sun, J.; Li, B.; Wang, Q.; Zhang, P.; Zhang, Y.; Gao, L.; Li, X. Preparation of phosphorus-doped tungsten trioxide nanomaterials
and their photocatalytic performances. Environ. Technol. 2021, 42, 4104–4114.
28. Liu, G.; Xu, J.; Li, R. Facile synthesis of Cs0. 3WO3 nanofibers by hydrothermal method and their optical properties. Opt. Mater.
2020, 107, 110147.
29. Kako, T.; Zou, Z.; Ye, J. Photocatalytic oxidation of 2-propanol in the gas phase over cesium bismuth niobates under visible
light irradiation. Res. Chem. Intermed. 2005, 31, 359–364.
Molecules 2024, 29, 3126 13 of 13
30. Friesen, D.A.; Morello, L.; Headley, J.V.; Langford, C.H. Factors influencing relative efficiency in photo-oxidations of organic
molecules by Cs3PW12O40 and TiO2 colloidal photocatalysts. J. Photochem. Photobiol. A Chem. 2000, 133, 213–220.
31. Qiao, X.; Seo, H.J. A visible-light-driven photocatalyst of cesium vanadate Cs2V4O11 nanowires by hydrothermal method. Mater.
Lett. 2014, 136, 322–324.
32. Huang, Y.; Qin, J.; Hu, C.; Liu, X.; Wei, D.; Seo, H.J. Cs-doped α-Bi2O3 microplates: Hydrothermal synthesis and improved
photochemical activities. Appl. Surf. Sci. 2019, 473, 401–408.
33. Rajput, H.; Changotra, R.; Sangal, V.K.; Mahla, S.K.; Dhir, A. A facile synthesis of Cs loaded TiO2 nanotube photoelectrode for
the removal of 4-chloroguaiacol. Chemosphere 2019, 218, 687–695.
34. Guan, D.Q.; Shi, C.L.; Xu, H.Y.; Gu, Y.X.; Zhong, J.; Sha, Y.C.; Hu, Z.W.; Ni, M.; Shao, Z.P. Simultaneously mastering operando
strain and reconstruction effects via phase-segregation strategy for enhanced oxygen-evolving electrocatalysis. J. Energy Chem.
2023, 82, 572–580. https://doi.org/10.1016/j.jechem.2023.03.033.
35. Wang, X.; Zhao, Y.; Jiang, X.; Liu, L.; Li, X.; Li, H.; Liang, W. In-situ self-assembly of plant polyphenol-coated Fe3O4 particles for
oleaginous microalgae harvesting. J. Environ. Manag. 2018, 214, 335–345.
36. Sivakarthik, P.; Thangaraj, V.; Parthibavarman, M. A facile and one-pot synthesis of pure and transition metals (M=Co & Ni)
doped WO3 nanoparticles for enhanced photocatalytic performance. J. Mater. Sci. Mater. Electron. 2017, 28, 1–7.
37. Wang, Y.; Zhang, F.; Zhao, G.; Zhao, Y.; Ren, Y.; Zhang, H.; Zhang, L.; Du, J.; Han, Y.; Kang, D.J. Porous WO3 monolith-based
photoanodes for high-efficient photoelectrochemical water splitting. Ceram. Int. 2019, 45, 7302–7308.
38. Zou, D.; Yi, Y.N.; Song, Y.F.; Guan, D.Q.; Xu, M.G.; Ran, R.; Wang, W.; Zhou, W.; Shao, Z.P. The BaCe0.16Y0.04Fe0.8O3-δnanocom-
posite: A new high-performance cobalt-free triple-conducting cathode for protonic ceramic fuel cells operating at reduced tem-
peratures. J. Mater. Chem. A 2022, 10, 5381–5390. https://doi.org/10.1039/d1ta10652j.
39. Zhang, Y.; Guo, D.; Li, R.X. Synthesis ofCs0.3WO3 with visible transparency and near-infrared absorption from commercial WO3.
J. Solid State Chem. 2022, 306, 122768. https://doi.org/10.1016/j.jssc.2021.122768.
40. Windisch, C.F., Jr.; Exarhos, G.J.; Yao, C.; Wang, L.Q. Raman study of the influence of hydrogen on defects in ZnO. J. Appl. Phys.
2007, 101, 1012.
41. Wei, Z.; Wang, W.; Li, W.; Bai, X.; Zhao, J.; Tse, E.C.M.; Phillips, D.L.; Zhu, Y. Steering Electron-Hole Migration Pathways Using
Oxygen Vacancies in Tungsten Oxides to Enhance Their Photocatalytic Oxygen Evolution Performance. Angew. Chem. 2021, 60,
8236–8242.
42. Li, D.; Chandra, D.; Takeuchi, R.; Togashi, T.; Kurihara, M.; Saito, K.; Yui, T.; Yagi, M. Dual-Functional Surfactant-Templated
Strategy for Synthesis of an In Situ N2-Intercalated Mesoporous WO3 Photoanode for Efficient Visible-Light-Driven Water Ox-
idation. Chem. A Eur. J. 2017, 23, 6596–6604. https://doi.org/10.1002/chem.201700088.
43. Huang, H.; Li, D.; Lin, Q.; Zhang, W.; Shao, Y.; Chen, Y.; Sun, M.; Fu, X. Efficient degradation of benzene over LaVO4/TiO2
nanocrystalline heterojunction photocatalyst under visible light irradiation. Environ. Sci. Technol. 2009, 43, 4164–4168.
44. Wu, H.; Li, C.; Che, H.; Hu, H.; Hu, W.; Liu, C.; Ai, J.; Dong, H. Decoration of mesoporous Co3O4 nanospheres assembled by
monocrystal nanodots on g-C3N4 to construct Z-scheme system for improving photocatalytic performance. Appl. Surf. Sci. 2018,
440, 308–319.
45. Ye, M.; Gong, J.; Lai, Y.; Lin, C.; Lin, Z. High-efficiency photoelectrocatalytic hydrogen generation enabled by palladium quan-
tum dots-sensitized TiO2 nanotube arrays. J. Am. Chem. Soc. 2012, 134, 15720–15723.
46. Bertoluzzi, L.; Bisquert, J. Equivalent circuit of electrons and holes in thin semiconductor films for photoelectrochemical water
splitting applications. J. Phys. Chem. Lett. 2012, 3, 2517–2522.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual au-
thor(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.