Next Article in Journal
Electronic Structures of Kitaev Magnet Candidates RuCl3 and RuI3
Previous Article in Journal
M-Encapsulated Be12O12 Nano-Cage (M = K, Mn, or Cu) for CH2O Sensing Applications: A Theoretical Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Controlled Growth of WO3 Photoanode under Various pH Conditions for Efficient Photoelectrochemical Performance

Department of Energy Engineering, Dankook University, Cheonan 31116, Republic of Korea
*
Author to whom correspondence should be addressed.
Nanomaterials 2024, 14(1), 8; https://doi.org/10.3390/nano14010008
Submission received: 22 November 2023 / Revised: 15 December 2023 / Accepted: 18 December 2023 / Published: 19 December 2023
(This article belongs to the Section Energy and Catalysis)

Abstract

:
Herein, the effects of the precursor solution’s acidity level on the crystal structure, morphology, nucleation, and growth of WO3·nH2O and WO3 nanostructures are reported. Structural investigations on WO3·nH2O using X-ray diffraction and Fourier–transform infrared spectroscopy confirm that the quantity of hydrate groups increases due to the interaction between H+ and water molecules with increasing HCl volume. Surface analysis results support our claim that the evolution of grain size, surface roughness, and growth direction on WO3·nH2O and WO3 nanostructures rely on the precursor solution’s pH value. Consequently, the photocurrent density of a WO3 photoanode using a HCl-5 mL sample achieves the best results with 0.9 mA/cm2 at 1.23 V vs. a reversible hydrogen electrode (RHE). We suggest that the improved photocurrent density of the HCl-5 mL sample is due to the efficient light absorption from the densely grown WO3 nanoplates on a substrate and that its excellent charge transport kinetics originate from the large surface area, low charge transfer resistance, and fast ion diffusion through the photoanode/electrolyte interface.

1. Introduction

As the climate crisis has been aggravated because of the excessive use of fossil fuels, the demand for green and renewable energy has steadily increased worldwide [1]. As a result, it is crucial to secure carbon-free energy resources to meet the growing energy consumption. Green hydrogen production is an efficient way of generating high energy density, clean, and storable fuel [2,3]. One of the promising methods of achieving green hydrogen is to adopt a photoelectrochemical route that operates without an external bias [4,5,6,7].
Tungsten trioxide hydrates (WO3·nH2O, n = 0.33, 1, or 2) and tungsten trioxides (WO3) have been widely investigated due to their diverse crystal structures and intriguing electrochemical and photovoltaic properties [8,9,10,11,12,13,14,15,16,17,18,19,20]. In particular, WO3 has been considered as a promising semiconductor material for water splitting applications via a photoelectrochemical method, because it is stable and non-toxic, and its band gap is also suitable for visible light absorption [12,14,16,21,22,23]. In general, WO3·nH2O has been prepared through the solution-phase synthesis, and the WO3 has been obtained after removing water molecules from WO3·nH2O using calcination processes [24,25,26]. Thus, many efforts have been exerted on WO3·nH2O and WO3 nanostructures to develop suitable preparation methods for efficient photoelectrode applications, including acid precipitation and a hydrothermal process, by controlling the morphology, crystal structure, and growth direction at room or moderate temperatures [14,15,16,17,18,19,20,21,27,28]. However, to the best of our knowledge, there has been no report on the relationship between the structural evolution of WO3·nH2O and its effect on the electrochemical performance of WO3 photoanodes while controlling the crystal structure of WO3·nH2O.
In the present study, we prepared WO3·nH2O nanostructures by acid precipitation with various levels of hydrochloric (HCl) acid concentration and elucidated the influence of the precursor solution’s acidity level on the crystal structure, morphology, nucleation, and initial growth mode of the WO3·nH2O samples. Furthermore, we investigated the relationship between the structural evolution of WO3·nH2O nanostructures and their effect on the water oxidation properties of WO3 photoanodes.

2. Experiments

2.1. Synthesis of WO3·nH2O and WO3 Materials

In order to improve the statistical significance of the results, three samples were prepared by repeating the following experimental procedure three times for each sample. Firstly, WO3·nH2O was prepared by simple acid precipitation at room temperature with the following processes: 0.4 mmol of sodium tungstate dihydrate (Na2WO4∙2H2O) was dissolved in 15 mL of deionized (DI) water and various amounts of 3 M HCl were dropped in the prepared solution while it was stirred. The samples obtained after adding each volume of HCl solution were labeled HCl-1 mL, HCl-2.5 mL, HCl-5 mL, HCl-7.5 mL, and HCl-10 mL, respectively. Then, 0.8 mmol of ammonium oxalate was dissolved in 15 mL of DI and added to the above solution with stirring to achieve the polymeric oxide. The prepared solution was poured into a Teflon-lined autoclave for hydrothermal synthesis at 140 °C for 3 h. Lastly, the as-grown WO3·nH2O samples were rinsed with DI and ethyl alcohol several times, and WO3 materials were obtained after they were calcined using a furnace at 500 °C in air for 2 h. The preparation procedures of WO3·nH2O and WO3 materials are presented as a schematic illustration in the Supplementary Materials (Figure S1).

2.2. Preparation of a WO3 Photoanode on Substrate

Fluorine-doped tin oxide (FTO) glass was used as a substrate cut to 2 cm x 2 cm and then cleaned by sequential sonication in ethyl alcohol, isopropyl alcohol, and DI water for 10 min in each solution. The cleaned FTO glass was vertically aligned on the jig and placed in the autoclave for the growth of the WO3 film on the substrate. The detailed preparation procedures of the WO3 photoanode on the FTO substrate are provided as an illustration in Figure S2.

2.3. Characterizations

The crystal structure was analyzed by X-ray diffraction (XRD, Miniflex 600) using a Cu Kα source. Fourier–transform infrared (FTIR, Perkin Elmer, Waltham, MA, USA) spectroscopy was adopted to study the chemical compositions. Surface analyses were conducted using scanning electron microscopy (SEM, Gemini II, Oberkochen, Germany) and atomic force microscopy (AFM, XE-100). The pH value of the solution was measured with a pH meter (HI2214, HANNA Instruments, Smithfield, RI, USA) which was calibrated with three kinds of pH buffer solutions for use in the high acidity condition. The pH controller was sequentially calibrated in pH 7, pH 4, and pH 1.68 buffer solutions (Thermo Scientific, Waltham, MA, USA). The acidity of the precursor suspension was varied within a wide pH range from 0.05 ± 0.01. to 1.12 ± 0.02 by adjusting the HCl volume. The names of the samples according to the HCl volume and average pH values for precursor suspensions are provided in Table 1. All electrochemical performances were measured with a three-electrode system using platinum as a counter electrode, Ag/AgCl as a reference electrode, and 0.5 M Na2SO4 (pH 6.9) used as an electrolyte. A Xenon lamp (LS150, ABET technologies) with a power of 150 W was used as the light source, and the illumination area of the photoanode was 2.54 cm2. The light intensity of a solar simulator with an air mass (AM) 1.5 G filter was calibrated to a 1 sun condition (100 mW/cm2) using a Si photodiode (RR_227 KG5, ABET technologies). And, the photoelectrochemical measurements on the WO3 photoanode were conducted by using back-side illumination. An electrochemical workstation (Versastat3, Ametek) was used for linear scan voltammetry (LSV) measurements with a scan rate of 10 mV/s. Electrochemical impedance spectroscopy (EIS) was conducted in the frequency range from 1 MHz to 0.1 Hz with a 10 mV voltage amplitude. The obtained potential of the working electrode was converted into the potential of the reversible hydrogen electrode (RHE) using the following equation:
ERHE = EAg/AgCl + 0.197 + 0.059 pH
where ERHE is the potential referred to for the RHE and EAg/AgCl is the achieved potential against the Ag/AgCl (3 M KCl saturated) reference electrode.

2.4. Data Acquisition

To demonstrate the reproducibility of all the samples, XRD, LSV, and EIS data were obtained. Three samples with different HCl volumes were measured under the same condition and showed almost identical results as shown in the Supplementary Materials (Figures S3–S6); therefore, samples denoted as #1 among the replica were reported in the Results and Discussion section to provide their typical results.

3. Results and Discussion

Figure 1 illustrates the structural evolution of WO3·nH2O by varying the HCl concentration during the acid precipitation process using Na2WO4 as a precursor, DI water as a solvent, and HCl solution as a precipitant. When HCl was introduced to Na2WO4, NaCl and tungstic acid (H2WO4) were generated by the following chemical reaction (1) [18]:
2HCl + Na2WO4 → H2WO4 + 2NaCl
In general, H2WO4 can exist in the form of neutral WO2(OH)2. Under highly acidic conditions, namely, a high H+ ion concentration and low pH value, nucleophilic water molecules tend to interact with [WO2(OH)2] by the following reaction (2) [18,29,30]:
[WO2(OH)2] + 2H2O → [WO(OH)4(OH2)]
According to the above reaction, WO(OH)4(OH2) metal complexes were developed, and the aggregation between adjacent octahedral WO(OH)4(OH2) occurred in the form of corner-sharing. Moreover, the 2D stacked layers of WO3·H2O and WO3·2H2O were gradually created with increasing interaction between seed WO(OH)4(OH2) and water molecules owing to van der Waals forces [15,31].
According to the chemical reactions of (1) and (2), the structural change of the WO3·nH2O caused by the H+ concentration was clearly observed in the XRD patterns as shown in Figure 2. The XRD peaks of each sample corresponded to the hexagonal WO3∙0.33H2O (JCPDS no. 35-1001), orthorhombic WO3∙H2O (JCPDS no. 43-0679), and monoclinic WO3∙2H2O phases (JCPDS no. 18-1420). For the HCl-1 mL sample, 1 mL of HCl was added into a Na2WO4 solution during acid precipitation, and the hexagonal WO3∙0.33H2O and orthorhombic WO3∙H2O phases developed because of the weak interaction between a low H+ concentration and nucleophilic water molecules. However, the peak intensity at 22.9°, 24.3°, 26.9°, and 28.3° assigned to the lattice planes of (001), (110), (101), and (200) of the hexagonal structure greatly decreased with an increasing HCl volume, suggesting that the WO3∙0.33H2O hexagonal phase transformed into the WO3∙H2O orthorhombic phase as shown in Figure 2b [32]. Likewise, Figure 2b clearly shows that the crystal structure of the HCl-7.5 mL and HCl-10 mL samples were transformed from the orthorhombic WO3∙H2O phase to the monoclinic WO3∙2H2O phase owing to the promoted interaction of H+ and water molecules leading to the formation of WO3∙2H2O. Furthermore, based on XRD measurements, we claim that the reproducibility of WO3·nH2O can be demonstrated as shown in Figure S3. Therefore, we suggest that the crystal structure of WO3·nH2O (n = 0.33, 1.00, or 2.00) can be precisely engineered by changing the volume of HCl precipitants during room temperature wet synthesis.
Figure 3 shows the FTIR spectra of the prepared WO3·nH2O samples under different HCl concentrations. The absorption bands in the range of 500–1000 cm−1 are characteristic of the various WO3 crystal lattices, i.e., W–O, O–W–O, and W=O bonds. The strong peak around 630 cm−1 was assigned as the stretching mode of metal–oxygen (O–W–O) [15,33]. In addition, the strong IR absorption observed at 945 cm−1 was due to the stretching mode of W=O bonds of hydrated WO3 with an increasing H+ concentration [33]. Two strong absorption peaks near 1610 and 3500 cm−1 arose from the H–O–H bending and O–H stretching vibrations of H2O, respectively, suggesting that WO3·nH2O can host a large number of water molecules either in the coordinated or interlayer form [33,34]. Thus, the FTIR results suggested that the quantity of hydrate groups increased due to the interaction of H+ and nucleophilic water molecules with an increasing HCl volume, which is consistent with the XRD results.
To elucidate the relationship between the crystal structure and the initial nucleation growth of the WO3·nH2O on an FTO substrate, AFM was used to achieve three-dimensional (3D) images of all the samples. The 3D AFM images of the WO3·nH2O nanostructures grown on FTO glass with different amounts of HCl during acid precipitation are presented in Figure 4. The AFM image of the bare FTO glass exhibits a smooth surface, while island-like grains can be observed for the samples with different HCl concentrations. The smooth surface of bare FTO glass exhibits a root mean square (RMS) of 38.4 nm (Figure 4a). However, as the amount of HCl increased during acid precipitation of WO3·nH2O nanostructures, sharp grains developed on the surfaces of the FTO glass (Figure 4b–f). The size and RMS of these grains increased up to a HCl volume of 5 mL as shown in Figure 4b–d. Among the prepared samples, the HCl-5 mL sample shows higher values of RMS surface roughness, indicating that the growth rate of WO3·nH2O nanostructures is the highest among the samples. In contrast, the growth rate of WO3·nH2O was decreased in the samples prepared with 7.5 mL and 10 mL of HCl. The driving force for nucleation is closely related to the supersaturation of WO(OH)4(OH2) seed molecules according to the chemical reaction of (2). At a larger pH value (low H+ concentration), the number of WO3 nuclei could be limited because of the small supersaturation of seed molecules. On the contrary, at a smaller pH value (high H+ concentration), a great deal of nuclei could be quickly generated, due to the larger supersaturation of seed molecules. However, Pham et al. report that the condensed H+ ion had a confining effect that generates a strong electric field, inhibiting long-ordered nucleation growth [18]. Therefore, the optimized nucleation growth of WO3 nanoplates can be obtained in the moderate pH value and it is found to be the HCl-5 mL sample in our experimental results. Therefore, we suggest that the evolution of grain size, surface roughness, and inter-grain space distribution on WO3·nH2O nanostructures rely on the precursor solution’s acidity level.
After the hydrothermal growth and calcination process, WO3·nH2O nanostructures can be dehydrated and transformed to achieve the WO3 phase [26,27]. To study the crystal structure of WO3 materials on FTO glass, we conducted the XRD measurements, and the diffraction results are provided in Figure 5a. Regardless of the WO3·nH2O nanostructures and HCl concentration, the crystal structure of all the samples was identified as monoclinic WO3 (JCPDS no.43-1035) after calcination [35]. It can be seen from Figure S4 that these results are demonstrated via repeated XRD measurements using a dummy sample. It should be noted that the peak intensity of the crystal plane gradually increased from the (002) to the (200) plane, as the HCl volume increased from 1 to 5 mL. The highest diffraction peak intensity of the crystal plane (200) was obtained from the WO3 nanoplate on FTO glass with the incorporation of 5 mL of HCl as shown in Figure 5b.
Figure 6 shows the SEM images of the monoclinic WO3 nanoplates grown on the FTO glass substrate after hydrothermal growth and the calcination process. The low-magnification top-view SEM images (Figure 6a–e) reveal that the HCl volume significantly affects the density of the WO3 nanoplate. As the HCl was incorporated into the precursor solution, the density of the WO3 nanoplates increased up to the HCl-5 mL sample; afterward, the density of the WO3 nanoplates drastically decreased in the WO3 nanoplates prepared with 7.5 mL and 10 mL of HCl, according to the AFM results as provided in Figure 4. Even though the WO3 nanoplate samples prepared by 1 mL of HCl had the lowest density among all the samples, due to the low supersaturation of seed molecules, a high-magnification SEM image (Figure 6f) revealed that the largest WO3 nanoplates were created by subsequent oxolation reactions on WO3 nuclei. Meanwhile, as the HCl volume increases, the seed molecules become more supersaturated, which makes it possible to generate a lot of nuclei on the substrate. However, the condensed H+ ion generates a strong electric field, inhibiting long-ordered nucleation and subsequent WO3 nanoplate growth as shown in Figure 6j. These results support our claim that the crystal growth of the WO3 nanoplate was greatly influenced by the initial nuclei density of WO3·nH2O nanostructures on the substrate; therefore, the optimum pH value for the growth of WO3 nanoplates should lie in the moderate pH range and it is revealed to be the HCl-5 mL sample as shown in Figure 6h. Figure 6k–o show the cross-sectional SEM images of the WO3 nanoplates on the substrates. Vertically aligned WO3 nanostructures with a length of 1–2 μm were well-grown on the FTO layer as indicated in Figure 6k. As confirmed from the magnified top-view SEM images, the largest WO3 nanoplates were observed in the HCl-1 mL sample (Figure 6k). On the other hand, it should be noted that the rest four samples have a similar growth thickness, but they have a different density of WO3 nanoplates. Moreover, we clearly observed that the growth direction of the WO3 nanoplates changed from a slanted to a vertical direction for the substrates according to the proportion of the WO3 nanoplates (Figure 6k–o). When the WO3 nanoplates were sparsely distributed on the substrate under a low acidity level, they were grown on FTO glass with an inclined direction against the substrate. However, as the number of WO3 nanoplates increased at high acidity levels, they densely grew on the substrate and the density of WO3 nanoplates was maximized in the HCl-5 mL sample as shown in Figure 6m.
Consequently, the preferential growth direction of the WO3 nanoplate changed according to the initial growth conditions, which resulted in a change in the strongest XRD peak from (002) to (200) as shown in Figure 5b. Thus, we claim that the HCl volume plays a crucial role in the growth process of WO3·nH2O nanostructures and WO3 nanoplates on FTO substrates.
The PEC performance of the prepared WO3 photoanodes was investigated under AM 1.5 G conditions with an illuminated area of 2.54 cm2. Figure 7 shows the LSV results of the WO3 photoanodes with different amounts of HCl in a 0.5 M Na2SO4 electrolyte. According to the LSV curve, the dark currents of all the samples were less than 20 μA/cm2 and were regarded as negligible. When the HCl concentration increased, the photocurrent density was enhanced as shown in Figure 7. The photocurrent density of the photoanode reached the highest value with 0.9 mA/cm2 at 1.23 V vs. RHE for the HCl-5 mL sample, while that of the HCl-1 mL sample exhibited the lowest with 0.3 mA/cm2 at the same voltage. However, as the incorporated amount of HCl further increased to more than 5 mL, the photocurrent densities decreased from 0.65 to 0.57 mA/cm2 at 1.23 V vs. RHE for the WO3 photoanodes prepared with 7.5 mL and 10 mL of HCl. As a result of repeated measurement using a replica, we suggest that the WO3 photoanodes with different HCl volumes can be identically reproduced as shown in Figure S5. Therefore, we claim that the improved photocurrent is due to the enhanced light absorption from the dense and large volume of WO3 nanoplates in a HCl-5 mL photoanode, which facilitates the charge transport kinetics by producing more photogenerated electron-hole pairs.
EIS measurements were conducted to prove the charge transfer kinetics at the WO3 photoanodes–electrolyte interface. The EIS results are presented as a Nyquist plot at an open-circuit voltage under AM 1.5 G illumination as shown in Figure 8. As seen in Figure 8, the HCl-5 mL photoanode sample represents the smallest semicircle compared to the other electrodes, indicating a much lower photoelectrode–electrolyte interfacial resistance compared to the other electrodes [36]. In this work, the obtained EIS data were fitted into the RC circuit model as provided in the inset of Figure 8, containing a resistance and a capacitance in parallel. Because the water oxidation takes place at the WO3 photoanode–electrolyte interface, the charge transfer resistance (Rct) and double layer capacitance (CPE) can be observed at the interface [37,38]. The charge transfer resistances for all the samples were obtained by using the ZSimpWin software (version 3.21), and their values for the HCl-1 mL, HCl-2.5 mL, HCl-5 mL, HCl-7.5 mL, and HCl-10 mL samples were 1445, 1372, 687, 987, and 2704 Ω, respectively. The fitting results clearly confirm that the derived Rct values for the HCl-5 mL photoanode were the smallest among all the samples. From the results of EIS data for three samples under the same condition, we confirm that the reproducibility of photoanodes–electrolyte interfacial resistance can be demonstrated as shown in Figure S6. Therefore, we suggest that the superior PEC performance of the HCl-5 mL sample is attributed to the electrode–electrolyte accessibility of the WO3 nanoplate owing to the densely grown nanostructure, efficient charge transfer, and fast ion diffusion between the photoanode and electrolyte interface.

4. Conclusions

In summary, we demonstrated the influence of the precursor solution’s pH level on the structural evolution, crystal growth, and nucleation of WO3·nH2O and WO3 nanostructures. The XRD and FTIR results on WO3·nH2O verified that the quantity of hydrate groups increased due to the interaction of H+ and water molecules with an increasing HCl volume. Surface analyses using AFM and SEM confirmed that the evolution of grain size, surface roughness, and grain growth on WO3·nH2O and WO3 nanostructures were affected by the precursor solution’s pH value. The photoelectrochemical performance confirmed that the best photocurrent density and charge transfer resistance were obtained using the HCl-5 mL sample due to its structural benefits, which facilitated efficient charge transfer and ion diffusion between the photoanode and the electrolyte. Our findings offer a facile approach to obtaining superior performance in PEC water splitting.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano14010008/s1, Figure S1: Schematics of the preparation procedures of WO3⋅nH2O and WO3 materials. Figure S2: Illustrations of the preparation procedures of WO3 photoanode on FTO substrate. Figure S3: XRD spectra of the WO3⋅nH2O powders obtained by acid precipitation process with 2theta range from 10° to 70°. (a) HCl-1 mL, (b) HCl-2.5 mL, (c) HCl-5 mL, (d) HCl-7.5 mL, (e) HCl-10 mL samples. Figure S4: XRD spectra of the WO3 nanoplates grown on FTO glass with different HCl volume. (a) HCl-1 mL, (b) HCl-2.5 mL, (c) HCl-5 mL, (d) HCl-7.5 mL, (e) HCl-10 mL samples. Figure S5: LSV curves obtained from the WO3 photoanodes under AM 1.5G condition. (a) HCl-1 mL, (b) HCl-2.5 mL, (c) HCl-5 mL, (d) HCl-7.5 mL, (e) HCl-10 mL samples. Figure S6: EIS Nyquist plots of all the photoanodes under AM 1.5G condition. (a) HCl-1 mL, (b) HCl-2.5ml, (c) HCl-5 mL, (d) HCl-7.5 mL, (e) HCl-10 mL samples.

Author Contributions

Conceptualization, S.-H.B. and S.-J.Y.; methodology, S.-H.B.; software, S.-H.B.; validation, S.-H.B. and S.-J.Y. and D.K.; formal analysis, S.-J.Y.; investigation, S.-H.B.; resources, S.-J.Y.; data curation, S.-J.Y. and D.K.; writing—original draft preparation, S.-H.B. and S.-J.Y.; writing—review and editing, S.-H.B.; visualization, S.-J.Y. and D.K.; supervision, S.-H.B.; project administration, S.-H.B.; funding acquisition, S.-H.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Government of the Republic of Korea (MSIT) (No. 2023R1A2C1004725 and No. RS-2023-00236572). And, the present research was supported by the research fund of Dankook University in 2023 (Outstanding Junior Faculty Researcher Support Program).

Data Availability Statement

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

Conflicts of Interest

The authors declare that there are no conflict of interest.

References

  1. Armaroli, N.; Balzani, V. The future of energy supply: Challenges and opportunities. Angew. Chem. Int. Ed. 2007, 46, 52. [Google Scholar] [CrossRef] [PubMed]
  2. Chen, X.; Li, C.; Grätzel, M.; Kostecki, R.; Mao, S.S. Nanomaterials for renewable energy production and storage. Chem. Soc. Rev. 2012, 41, 7909. [Google Scholar] [CrossRef] [PubMed]
  3. Hosseini, S.E.; Wahid, M.A. Hydrogen production from renewable and sustainable energy resources: Promising green energy carrier for clean development. Renew. Sustain. Energy Rev. 2016, 57, 850. [Google Scholar] [CrossRef]
  4. Nikolaidis, P.; Poullikkas, A. A comparative overview of hydrogen production processes. Renew. Sustain. Energy Rev. 2017, 67, 597. [Google Scholar] [CrossRef]
  5. Møller, K.T.; Jensen, T.R.; Akiba, E.; Li, H.-W. Hydrogen—A sustainable energy carrier. Prog. Nat. Sci. 2017, 27, 34. [Google Scholar] [CrossRef]
  6. Kim, J.H.; Hansora, D.; Sharma, P.; Jang, J.W.; Lee, J.S. Toward practical solar hydrogen production—An artificial photosynthetic leaf-to-farm challenge. Chem. Soc. Rev. 2019, 48, 1908. [Google Scholar] [CrossRef] [PubMed]
  7. Ye, S.; Ding, C.; Liu, M.; Wang, A.; Huang, Q.; Li, C. Water oxidation catalysts for artificial photosynthesis. Adv. Mater. 2019, 31, 1902069. [Google Scholar] [CrossRef] [PubMed]
  8. Kuti, L.M.; Bhella, S.S.; Thangadurai, V. Transformation of proton-conducting perovskite-type into fluorite-type fast oxide ion electrolytes using a CO2 capture technique and their electrical properties. Inorg. Chem. 2009, 48, 6804. [Google Scholar] [CrossRef]
  9. Gotić, M.; Ivanda, M.; Popović, S.; Musić, S. Synthesis of tungsten trioxide hydrates and their structural properties. Mater. Sci. Eng. B 2000, 77, 193. [Google Scholar] [CrossRef]
  10. Jiao, Z.; Wang, J.; Ke, L.; Sun, X.W.; Demir, H.V. Morphology-tailored synthesis of tungsten trioxide (hydrate) thin films and their photocatalytic properties. ACS Appl. Mater. Interfaces 2011, 3, 229. [Google Scholar] [CrossRef]
  11. Amano, F.; Tian, M.; Wu, G.; Ohtani, B.; Chen, A. Facile preparation of platelike tungsten oxide thin film electrodes with high photoelectrode activity. ACS Appl. Mater. Interfaces 2011, 3, 4047. [Google Scholar] [CrossRef] [PubMed]
  12. Nayak, A.K.; Sohn, Y.; Pradhan, D. Facile green synthesis of WO3·H2O nanoplates and WO3 nanowires with enhanced photoelectrochemical performance. Cryst. Growth Des. 2017, 17, 4949. [Google Scholar] [CrossRef]
  13. Mitchell, J.B.; Lo, W.C.; Genc, A.; Lebeau, J.; Augustyn, V. Transition from Battery to Pseudocapacitor Behavior via Structural Water in Tungsten Oxide. Chem. Mater. 2017, 29, 3928. [Google Scholar] [CrossRef]
  14. Zhou, J.; Lin, S.; Chen, Y.; Gaskov, A.M. Facile morphology control of WO3 nanostructure arrays with enhanced photoelectrochemical performance. Appl. Surf. Sci. 2017, 403, 274. [Google Scholar] [CrossRef]
  15. Nayak, A.K.; Lee, S.; Choi, Y.I.; Yoon, H.J.; Sohn, Y.; Pradhan, D. Crystal phase and size-controlled synthesis of tungsten trioxide hydrate nanoplates at room temperature: Enhanced Cr(VI) photoreduction and methylene blue adsorption properties. ACS Sustain. Chem. Eng. 2017, 5, 2741. [Google Scholar] [CrossRef]
  16. Zheng, G.; Wang, J.; Liu, H.; Murugadoss, V.; Zu, G.; Che, H.; Lai, C.; Li, H.; Ding, T.; Gao, Q.; et al. Tungsten oxide nanostructures and nanocomposites for photoelectrochemical water splitting. Nanoscale 2019, 11, 18968. [Google Scholar] [CrossRef] [PubMed]
  17. Zhang, W.; Chen, H.; Peng, Q.; Liu, X.; Dian, L.; Ye, W.; Chang, Y.; Ma, X.; Wang, H. Controlled fabrication of WO3 nanoplate films and its photoelectrochemical properties. Mater. Res. Express. 2019, 6, 095901. [Google Scholar] [CrossRef]
  18. Pham, N.L.; Luu, T.L.A.; Nguyen, H.L.; Nguyen, C.T. Effects of acidity on the formation and adsorption activity of tungsten oxide nanostructures prepared via the acid precipitation method. Mater. Chem. Phys. 2021, 272, 125014. [Google Scholar] [CrossRef]
  19. Liang, F.; Lin, Y.; He, Z.; Chen, W.; Zhu, Y.; Chen, T.; Liang, L.; Ma, S.; Wu, Y.; Tu, B.; et al. Promising ITO-free perovskite solar cells with WO3-Ag-SnO2 as transparent conductive oxide. J. Mater. Chem. A 2018, 6, 19330. [Google Scholar] [CrossRef]
  20. Yadav, S.; Lohia, P.; Sahu, A. Enhanced performance of double perovskite solar cell using WO3 as an electron transport material. J. Opt. 2023, 52, 776. [Google Scholar] [CrossRef]
  21. Farhadian, M.; Sangpout, P.; Hosseinzadeh, G. Morphology dependent photocatalytic activity of WO3 nanostructures. J. Energy Chem. 2015, 24, 171. [Google Scholar] [CrossRef]
  22. Park, M.; Seo, J.H.; Song, H.; Nam, K.M. Enhanced visible light activity of single-crystalline WO3 microplates for photoelectrochemical water oxidation. J. Phys. Chem. C 2016, 120, 9192. [Google Scholar] [CrossRef]
  23. Ahmed, B.; Kumar, S.; Ojha, A.K.; Donfack, P.; Materny, A. Facile and controlled synthesis of aligned WO3 nanorods and nanosheets as an efficient photocatalyst material. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2017, 175, 250. [Google Scholar] [CrossRef] [PubMed]
  24. Ke, J.; Zhou, H.; Liu, J.; Duan, X.; Zhang, H.; Liu, S.; Wang, S. Crystal transformation of 2D tungstic acid H2WO4 to WO3 for enhanced photocatalytic water oxidation. J. Colloid Interface Sci. 2018, 514, 576. [Google Scholar] [CrossRef] [PubMed]
  25. Zhao, L.; Xi, X.; Liu, Y.; Ma, L.; Nie, Z. Facile synthesis of WO3 micro/nanostructures by paper-assisted calcination for visible-light-driven photocatalysis. Chem. Phys. 2020, 528, 110515. [Google Scholar] [CrossRef]
  26. Perfecto, T.M.; Zito, C.A.; Volanti, D.P. Room-temperature volatile organic compounds sensing based on WO3·0.33H2O, hexagonal-WO3, and their reduced graphene oxide composites. RSC Adv. 2016, 6, 105171. [Google Scholar]
  27. Wang, L.; Hu, H.; Xu, J.; Zhu, S.; Ding, A.; Deng, C. WO3 nanocubes: Hydrothermal synthesis, growth mechanism, and photocatalytic performance. J. Mater. Res. 2019, 34, 2955. [Google Scholar] [CrossRef]
  28. Ou, P.; Song, F.; Yang, Y.; Shao, J.; Hua, Y.; Yang, S.; Wang, H.; Luo, Y.; Liao, J. WO3·nH2O crystals with controllable morphology/phase and their optical absorption properties. ACS Omega 2022, 7, 8833. [Google Scholar] [CrossRef]
  29. Livage, J.; Guzman, G. Aqueous precursors for electrochromic tungsten oxide hydrates. Solid State Ion. 1996, 84, 205. [Google Scholar] [CrossRef]
  30. Hu, W.-H.; Han, G.-Q.; Dong, B.; Liu, C.-G. Facile Synthesis of Highly Dispersed WO3·H2O and WO3 Nanoplates for Electrocatalytic Hydrogen Evolution. J. Nanomater. 2015, 2015, 346086. [Google Scholar]
  31. Han, H.; Nayak, A.K.; Choi, H.; Ali, G.; Kwon, J.; Choi, S.; Paik, U.; Song, T. Partial dehydration in hydrated tungsten oxide nanoplates leads to excellent and robust bifunctional oxygen reduction and hydrogen evolution reactions in acidic media. ACS Sustain. Chem. Eng. 2020, 8, 9507. [Google Scholar] [CrossRef]
  32. Zhang, N.; Chen, C.; Mei, Z.; Liu, X.; Qu, X.; Li, Y.; Li, S.; Qi, W.; Zhang, Y.; Ye, J.; et al. Monoclinic tungsten oxide with {100} facet orientation and tuned electronic band structure for enhanced photocatalytic oxidations. ACS Appl. Mater. Interfaces 2016, 8, 10367. [Google Scholar] [CrossRef] [PubMed]
  33. Daniel, M.F.; Desbat, B.; Lassegues, J.C.; Gerand, B.; Figlarz, M. Infrared and Raman study of WO3 tungsten trioxides and WO3, xH2O tungsten trioxide tydrates. J. Solid State Chem. 1987, 67, 235. [Google Scholar] [CrossRef]
  34. Baek, S.-H.; Jeong, Y.-M.; Kim, D.Y.; Park, I.-K. Phase transformation of NiCo hydroxides derived from carbonate anion and its effect on electrochemical pseudocapacitor performance. Chem. Eng. J. 2020, 393, 124713. [Google Scholar] [CrossRef]
  35. Zheng, J.Y.; Song, G.; Hong, J.; Van, T.K.; Pawar, A.U.; Kim, D.Y.; Kim, C.W.; Haider, Z.; Kang, Y.S. Facile fabrication of WO3 nanoplates thin films with dominant crystal facet of (002) for water splitting. Cryst. Growth Des. 2014, 14, 6057. [Google Scholar] [CrossRef]
  36. Bredar, A.R.C.; Chown, A.L.; Burton, A.R.; Farnum, B.H. Electrochemical impedance spectroscopy of metal oxide electrodes for energy applications. ACS Appl. Energy Mater. 2020, 3, 66. [Google Scholar] [CrossRef]
  37. Kim, J.Y.; Magesh, G.; Youn, D.H.; Jang, J.-W.; Kubata, J.; Domen, K.; Lee, J.S. Single-crystalline, wormlike hematite photoanodes for efficient solar water splitting. Sci. Rep. 2013, 3, 2681. [Google Scholar] [CrossRef] [PubMed]
  38. Kong, Y.; Sun, H.; Fan, W.; Wang, L.; Zhao, H.; Zhao, X.; Yuan, S. Enhanced photoelectrochemical performance of tungsten oxide film by bifunctional Au nanoparticles. RSC Adv. 2017, 7, 15201. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the WO3·nH2O under various pH conditions during the acid precipitation process.
Figure 1. Schematic diagram of the WO3·nH2O under various pH conditions during the acid precipitation process.
Nanomaterials 14 00008 g001
Figure 2. XRD spectra of the WO3·nH2O powders obtained by an acid precipitation process. (a) 2theta range from 10° to 70° and (b) magnified XRD spectra in the 2theta range from 20° to 30°.
Figure 2. XRD spectra of the WO3·nH2O powders obtained by an acid precipitation process. (a) 2theta range from 10° to 70° and (b) magnified XRD spectra in the 2theta range from 20° to 30°.
Nanomaterials 14 00008 g002
Figure 3. FTIR spectra of the WO3·nH2O powders obtained by an acid precipitation process.
Figure 3. FTIR spectra of the WO3·nH2O powders obtained by an acid precipitation process.
Nanomaterials 14 00008 g003
Figure 4. AFM images of the WO3·nH2O powders grown on FTO substrates with different HCl volumes during acid precipitation.
Figure 4. AFM images of the WO3·nH2O powders grown on FTO substrates with different HCl volumes during acid precipitation.
Nanomaterials 14 00008 g004
Figure 5. XRD results of the WO3 nanoplates grown on FTO glass. (a) 2theta range from 20° to 80° and (b) magnified XRD spectra in the 2theta range from 20° to 30° (orange squares indicate the XRD peak positions of the FTO substrate).
Figure 5. XRD results of the WO3 nanoplates grown on FTO glass. (a) 2theta range from 20° to 80° and (b) magnified XRD spectra in the 2theta range from 20° to 30° (orange squares indicate the XRD peak positions of the FTO substrate).
Nanomaterials 14 00008 g005
Figure 6. SEM images of the WO3 nanoplates grown on FTO glass with different HCl volumes. (a) HCl-1 mL, (b) HCl-2.5 mL, (c) HCl-5 mL, (d) HCl-7.5 mL, (e) HCl-10 mL samples. Magnified (fj) top-view and (ko) cross-section view SEM images of the WO3 nanoplates corresponding to the upper panel.
Figure 6. SEM images of the WO3 nanoplates grown on FTO glass with different HCl volumes. (a) HCl-1 mL, (b) HCl-2.5 mL, (c) HCl-5 mL, (d) HCl-7.5 mL, (e) HCl-10 mL samples. Magnified (fj) top-view and (ko) cross-section view SEM images of the WO3 nanoplates corresponding to the upper panel.
Nanomaterials 14 00008 g006
Figure 7. LSV curves obtained from the WO3 photoanodes under AM 1.5 G conditions.
Figure 7. LSV curves obtained from the WO3 photoanodes under AM 1.5 G conditions.
Nanomaterials 14 00008 g007
Figure 8. EIS Nyquist plots of all the photoanodes under AM 1.5 G conditions. (Inset: the equivalent circuit model).
Figure 8. EIS Nyquist plots of all the photoanodes under AM 1.5 G conditions. (Inset: the equivalent circuit model).
Nanomaterials 14 00008 g008
Table 1. Average pH values of precursor solutions according to HCl volume.
Table 1. Average pH values of precursor solutions according to HCl volume.
SampleHCl-1 mLHCl-2.5 mLHCl-5 mLHCl-7.5 mLHCl-10 mL
pH value1.12 ± 0.020.86 ± 0.020.62 ± 0.010.34 ± 0.010.05 ± 0.01
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

Yoo, S.-J.; Kim, D.; Baek, S.-H. Controlled Growth of WO3 Photoanode under Various pH Conditions for Efficient Photoelectrochemical Performance. Nanomaterials 2024, 14, 8. https://doi.org/10.3390/nano14010008

AMA Style

Yoo S-J, Kim D, Baek S-H. Controlled Growth of WO3 Photoanode under Various pH Conditions for Efficient Photoelectrochemical Performance. Nanomaterials. 2024; 14(1):8. https://doi.org/10.3390/nano14010008

Chicago/Turabian Style

Yoo, Seung-Je, Dohyun Kim, and Seong-Ho Baek. 2024. "Controlled Growth of WO3 Photoanode under Various pH Conditions for Efficient Photoelectrochemical Performance" Nanomaterials 14, no. 1: 8. https://doi.org/10.3390/nano14010008

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