Next Article in Journal
Solvent-Exchange Triggered Solidification of Peptide/POM Coacervates for Enhancing the On-Site Underwater Adhesion
Next Article in Special Issue
Preparation of Oily Sludge-Derived Activated Carbon and Its Adsorption Performance for Tetracycline Hydrochloride
Previous Article in Journal
Fragmentation Pathway of Organophosphorus Flame Retardants by Liquid Chromatography–Orbitrap-Based High-Resolution Mass Spectrometry
Previous Article in Special Issue
Improvement in the Chromium(VI)-Diphenylcarbazide Determination Using Cloud Point Microextraction; Speciation of Chromium at Low Levels in Water Samples
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synergistic Effect of Co3(HPO4)2(OH)2 Cocatalyst and Al2O3 Passivation Layer on BiVO4 Photoanode for Enhanced Photoelectrochemical Water Oxidation

Guangxi Key Laboratory of Multidimensional Information Fusion for Intelligent Vehicles, School of Electronic Engineering, Guangxi University of Science and Technology, Liuzhou 545000, China
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(3), 683; https://doi.org/10.3390/molecules29030683
Submission received: 29 December 2023 / Revised: 27 January 2024 / Accepted: 30 January 2024 / Published: 1 February 2024
(This article belongs to the Collection Green Energy and Environmental Materials)

Abstract

:
Bismuth vanadate (BVO) is regarded as an exceptional photoanode material for photoelectrochemical (PEC) water splitting, but it is restricted by the severe photocorrosion and slow water oxidation kinetics. Herein, a synergistic strategy combined with a Co3(HPO4)2(OH)2 (CoPH) cocatalyst and an Al2O3 (ALO) passivation layer was proposed for enhanced PEC performance. The CoPH/ALO/BVO photoanode exhibits an impressive photocurrent density of 4.9 mA cm−2 at 1.23 VRHE and an applied bias photon-to-current efficiency (ABPE) of 1.47% at 0.76 VRHE. This outstanding PEC performance can be ascribed to the suppressed surface charge recombination, facilitated interfacial charge transfer, and accelerated water oxidation kinetics with the introduction of the CoPH cocatalyst and ALO passivation layer. This work provides a novel and synergistic approach to design an efficient and stable photoanode for PEC applications by combining an oxygen evolution cocatalyst and a passivation layer.

1. Introduction

Photoelectrochemical (PEC) water splitting is recognized as an attractive approach for converting inexhaustible solar energy into hydrogen energy [1,2,3,4]. In a typical PEC system, the oxygen evolution reaction (OER) occurs at the photoanode, and the hydrogen evolution reaction (HER) occurs at the counter electrode. Owing to the characteristics of the material itself, there is currently no single semiconductor material that can satisfy these two half-reactions at the same time, so the anode and cathode in the PEC system often use different semiconductor materials. At the core of the PEC water splitting is the OER reaction, a key challenge in PEC water splitting which involves a complex four-proton coupled multi-electron process. Moreover, the high performance of a photoanode is predominantly determined by its ability to absorb solar light, facilitate charge separation, and exhibit high catalytic activity and stability during the PEC water splitting. Among the variety of photoanode materials, BiVO4 (BVO) is recognized as one of the most promising choices for PEC water splitting due to its unique qualities, such as low cost, coupled with a narrow band gap (~2.4 eV) that renders it an optimal absorber of visible light, essential for the efficient harnessing of solar energy. Moreover, BVO’s advantageous positioning of its appropriate valence band greatly aids in the water oxidation process, a key step in PEC. These characteristics elevate BVO as a promising candidate for sustainable hydrogen production [5,6]. However, the rapid charge recombination, slow surface water oxidation, and severe photocorrosion limit its PEC water splitting performance [7,8,9]. To overcome these limitations, a variety of strategies have been employed to enhance the PEC performance of BVO-based photoanodes, such as morphology control [10], element doping [11], heterojunctions formation [12,13], and oxygen evolution catalysts (OECs) modification [14,15].
A promising strategy to facilitate charge separation and enhance the oxidation reaction kinetics involves the incorporation of suitable OECs because it could facilitate the charge separation, accelerate the water oxidation kinetics, and restrain photocorrison [16]. To date, numerous transition metal-based materials have been considerably utilized as OECs due to their excellent catalytic properties and earth abundance [17]. For example, Khiarak et al. introduced electrochemically reduced graphene oxide (ERGO)/sulfur-doped copper oxide supported with carbon cloth (CC), serving as an extremely efficient OER electrocatalyst [18]. Among the multitude of reported OECs, cobalt-based materials have been widely explored owing to their excellent water oxidation activity [17]. Within these cobalt-based OECs, cobalt hydroxides have attracted significant interest due to their favorable combination of moderate electronic conductivity and high catalytic activity [19]. For instance, Ning et al. deposited the Co(OH)2 as a cocatalyst on the BVO photoanode to obtain a higher photocurrent density. This improvement is attributed to the incorporation of Co(OH)2 to rapidly migrate photogenerated holes to its surface and effectively inhibit interface recombination [20]. As one of the most important electrocatalytic oxygen evolution catalysts, CoPi is often employed to modify semiconductor photoanodes due to its superior activity, low overpotential and inherent self-repair properties [21]. Hernandez et al. introduced CoPi on the BVO photoanode, achieving a high photocurrent density of 3.0 mA cm−2 at 1.23 VRHE. This improvement is attributed to CoPi’s role as an effective oxygen evolution catalyst [22]. Reddy et al. used CoPi to modify the BVO photoanode to achieve a notable photocurrent density of 2.7 mA cm−2 and a negative shift in the onset potential to 0.32 VRHE, enhancing separation and suppressing recombination of charge carriers [23]. The phosphate group on the catalyst acts as a proton acceptor, which contributes to the self-repair of the catalyst and assists the proton–electron transfer process during metal oxidation, thereby improving the water oxidation activity [24]. However, the dissolution of CoPi under specific pH conditions significantly hinders its long-term stability, impacting its efficiency in PEC systems [25].
Recently, we reported that nickel nitrate hydroxides exhibit superior performance in electrocatalytic water oxidation processes compared to nickel hydroxides and nickel salts. According to our previous work, we found that transition metal basic salts have better electronic conductivity and more active sites, leading to the rapid charge transfer and fast water oxidation. Moreover, the presence of hydroxide groups could effectively enhance its chemical stability [26]. Consequently, even though the incorporation of transition metal basic salts in photoanode modification for PEC water splitting remains largely unexplored, their prospective role as competent and stable OECs is increasingly recognized. These materials, with their unique characteristics, are unitized to significantly enhance the PEC performance of photoanodes. Accordingly, they provide important advantages, including enhanced charge transfer capabilities and heightened durability under stringent operational conditions, which are critical for the PEC progression and technologies. This emerging recognition underscores the potential of transition metal basic salts to revolutionize solar fuel production. In addition, to suppress the photocorrosion and modify the surface state distribution of the BVO photoanode, a passivation layer was fabricated between the OECs and semiconductor [27]. The passivation layer is usually composed of an ultrathin layer of metal oxide, such as Al2O3 (ALO) [28], ZnO [29], NiOx [30], and so on. Among them, Al2O3 shows the best passivation effect and low cost [31].
Herein, we first report that the ALO passivation layer and Co3(HPO4)2(OH)2 (CoPH) OEC were synthesized on BVO (synthesized based on previous report [14]) surfaces by a convenient immersion–calcination method and a hydrothermal method, respectively (Figure 1). CoPH demonstrated superior enhanced properties in comparison to CoPi and Co(OH)2. Further PEC analysis indicated that the CoPH/ALO/BVO photoanode exhibits an impressive photocurrent density, reaching 4.9 mA cm−2 at 1.23 VRHE. This performance makes a significant advancement in the field of PEC water splitting. In addition, we observed that the stability of the CoPH/ALO/BVO photoanode was substantially superior to that of the bare BVO photoanode. This improvement in stability is critical for practical applications. These findings provide new avenues for the development of highly efficient photoanodes for solar energy conversion, providing the way toward more sustainable and eco-friendly energy solutions.

2. Results and Discussion

The crystal structures of as-prepared BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes were determined by X-ray diffraction (XRD) in Figure 2a [32]. The diffraction peaks at 18.9°, 28.9°, 30.5°, 35.2°, 40.0°, 47.2°, 50.3°, 53.4°, and 59.4° have been assigned to the (011), (112), (004), (020), (−121), (024), (301), (220), and (132) planes, indicating successful preparation of standard monoclinic BVO (PDF#83-1700). The diffraction peaks at 26.4°, 34.7°, 37.7°, 51.8°, and 65.9° have been assigned to the (110), (101), (200), (211), and (301) planes, which correspond to standard monoclinic SnO2 (PDF#77-0452). Those analyses indicate that the BVO had successful growth on FTO substrate. In addition, the diffraction peaks at 37.8° correspond to the (110) facet of standard ALO (PDF#77-2135), suggesting the deposition of the ALO layer on the BVO. Furthermore, no other diffraction peaks were detected, suggesting the low amount of CoPH is consistent with previous findings about metal complexes/semiconductor photoanodes [33]. Though the existence of CoPH is difficult to distinguish, the optical images of these photoanodes (as shown in Figure S1) exhibit the color change (bright yellow for BVO, dark yellow for ALO/BVO, bright yellow for CoPH/BVO, and brown for CoPH/ALO/BVO), indicating the successful loading of ALO and CoPH. To evaluate the phase and presence of CoPH, the CoPH powder was analyzed by an XRD test (Figure S2). The diffraction peaks at 26.0°, 26.9°, 27.4°, 29.4°, 33.8°, 36.1°, 38.5°, 42.6°, 53.6°, 55.3°, 57.8°, 61.9°, and 71.1° have been matched well with the (−112), (−202), (021), (−212), (−221), (−222), (−131), (−313), (−224), (−404), (400), (−503), and (−442) planes of standard monoclinic Co3(HPO4)2(OH)2 (PDF#80-1997). It suggests the low content of CoPH in CoPH/BVO and CoPH/ALO/BVO. To further ascertain the structure characteristics of the BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes, Raman spectra were examined within the range of 100–1100 cm−1 (Figure 2b). For the BVO, the peak at 826 cm−1 belongs to the typical V-O bond, while the peaks at 367 cm−1 and 326 cm−1 are attributed to the distinct VO43−. Moreover, the peaks detected at 124 and 211 cm−1 correspond to the external mode of the BVO [34]. Notably, the Raman spectra of the ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes are identical to the pristine BVO photoanode, which may be attributed to the low content of ALO and CoPH. The optical properties of these photoanodes were elucidated by UV-Vis diffused reflectance spectra. As shown in Figure 2c, the ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes exhibit higher absorption intensity in contrast to the pristine BVO. In addition, all photoanodes possess a similar absorption edge at 506 nm, indicating the negligible effect with the incorporation of ALO and CoPH on light absorption capacity, which is consistent with previous study [35]. Moreover, the UV-Vis spectrum of CoPH displays absorption edges located at 472, 490, 532, 574, and 682 nm, as shown in Figure S3. Both absorption bands can be attributed to the d-d transition of high-spin Co2+ in the twisted octahedron [36]. The morphologies of the BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes were investigated by SEM and HRTEM. As illustrated in Figure 2d, the BVO photoanode exhibits a cross-linked worm-like structure. As shown in Figure 2e, negligible morphology change can be observed for the ALO/BVO photoanode, revealing the low content of ALO. For the CoPH/ALO/BVO photoanode, it can be observed that circular block-like CoPH particles are embedded into the photoanode in Figure 2f. To further explore the morphological changes in CoPH with varying concentration, SEM images of CoPH/BVO photoanodes are presented in Figure S4. These photoanodes were prepared using different concentrations of CoPH (0.1 M, 0.2 M, 0.4 M). It can be obtained that the size of CoPH increases with incremental CoPH concentrations, ranging from 0.1 M to 0.4 M. HRTEM was employed to explore the detailed information of the CoPH/ALO/BVO photoanode as presented in Figure 2g. The lattice spacings of 0.30, 0.35, and 0.32 nm correspond to the BVO (112), ALO (012), and CoPH (021) facets, respectively. EDX mapping images (Figure 2h) of the CoPH/ALO/BVO photoanode reveal the existence of O, Al, P, Bi, V, and Co elements, which is consistent with EDX analysis (Figure S5). The elemental compositions were listed in Table S1. These results suggest the uniform distribution of O, Al, P, Bi, V, and Co elements, indicating the successful preparation of the CoPH/ALO/BVO photoanode.
X-ray photoelectron spectroscopy (XPS) measurement was carried out to determine the chemical states and surface compositions of BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes. As depicted in Figure 3a, the Bi 4f spectra exhibit distinct peaks at 159.0 and 164.3 eV correspond to Bi3+ 4f7/2 and Bi3+ 4f5/2, suggesting the presence of Bi3+ cation. In Figure 3b, the peaks at 516.5 and 524.3 eV can be attributed to V 2p3/2 and V 2p1/2 of V5+ species. Compared to the BVO photoanode, the ALO/BVO and CoPH/BVO photoanodes demonstrate that Bi 4f and V 2p peaks shift to higher binding energies (159.0 and 164.3 eV assigned to Bi3+ 4f7/2 and Bi3+ 4f5/2, and 516.6 and 524.1 eV attributed to V5+ 2p3/2 and V5+ 2p1/2 for ALO/BVO; 159.2 and 164.4 eV assigned to Bi3+ 4f7/2 and Bi3+ 4f5/2, and 516.7 and 524.3 eV attributed to V5+ 2p3/2 and V5+ 2p1/2 for CoPH/BVO). This finding suggests that there is a mutually attractive chemical interaction facilitating the photoinduced electron transfer between BVO and CoPH/ALO [32,37], thus enhancing PEC performance. This finding can be confirmed by the following EIS test and the LSV measurement. As shown in Figure 3c, the two divided O 1s peaks at 529.6 and 531.4 eV can be assigned to the lattice oxygen of Bi-O bond (purple color) and -OH groups (green color), respectively [38,39]. For the CoPH/BVO and CoPH/ALO/BVO photoanodes, O 1s peak at 532.9 eV assigned to P = O (cyan color) in HPO42− appears, suggesting the successful incorporation of CoPH.
Furthermore, the Al 2p XPS spectra of the ALO/BVO and CoPH/ALO/BVO photoanodes are displayed in Figure 3d. The peak located at 74.7 eV and assigned to Al3+ ions in ALO indicates the successful loading of ALO [40]. In Figure 3e, the P 2p XPS spectra of the CoPH/BVO and CoPH/ALO/BVO photoanodes possess two obvious peaks at 133.1 and 134.1 eV, assigned to P 2p3/2 and P 2p1/2, respectively. This analysis confirms the existence of (HPO4)2− in CoPH/BVO and CoPH/ALO/BVO photoanodes, which is likely to enhance the surface reaction kinetics [26]. The Co 2p XPS spectra of the CoPH/BVO and CoPH/ALO/BVO photoanodes are illustrated in Figure 3f. The peaks at 781.6 and 797.6 eV correspond to Co 2p3/2 and Co 2p1/2 (purple color) of Co2+. In addition, two oscillatory satellite peaks (labeled “Sat”, green color) can be observed at 803.5 and 786.4 eV, suggesting the presence of the divalent oxidation state of cobalt in CoPH/BVO and CoPH/ALO/BVO photoanodes [26,41]. This result is consistent with the XRD measurement and confirms the successful incorporation of CoPH in CoPH/BVO and CoPH/ALO/BVO photoanodes [42].
Initially, various Co-based OECs including CoBi, CoPi, Co(OH)2, and CoPH are employed to a modified BVO photoanode according to previous reports [43,44,45]. Figure 4a demonstrates that the CoPH/BVO photoanode achieves the optimal photocurrent density of 4.7 mA cm−2 at 1.23 VRHE, which is higher than that of the BVO (1.5 mA cm−2), CoPi/BVO (3.6 mA cm−2), CoBi/BVO (3.3 mA cm−2), and Co(OH)2/BVO (3.3 mA cm−2) photoanodes. Based on this analysis, the CoPH/BVO photoanode was further optimized with various CoPH concentrations (0.1, 0.2, and 0.4 M) as shown in Figure S6. The maximum photocurrent density was achieved with CoPH concentration at 0.2 M. Then, the as-prepared CoPH/BVO photoanode was selected for further investigation. To further assess the PEC performance of pristine BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes, LSV measurement was performed under illumination (AM 1.5 G, 100 mW cm−2). As shown in Figure 4b, the photocurrent densities for BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO at 1.23 VRHE are 1.5, 2.1, 4.7, and 4.9 mA cm−2, respectively. This result indicates the introduction of CoPH and CoPH/ALO can greatly enhance the PEC performance, which may ascribe to the accelerated water oxidation kinetics by CoPH. Table S2 shows the comparisons of PEC performance between previous BiVO4-based photoanodes and the CoPH/ALO/BVO photoanode. It can be obtained that CoPH/ALO/BVO exhibits an outstanding PEC performance among these photoanodes. In addition, the stability of the photoanode as an important parameter was evaluated with J-t plots at 1.23 VRHE. As illustrated in Figure 4c, it can be obtained that a rapid decay occurs for the BVO photoanode owing to the severe photocorrosion. ALO/BVO and CoPH/ALO/BVO photoanodes exhibit a great improvement on the durable stability (60 min) compared to the BVO photoanode, which may ascribe to the passivation effect induced by the ALO layer.
Furthermore, the surface charge separation efficiencies (ηsurface) for these photoanodes were measured with Na2SO3 as the hole scavenger. As exhibited in Figure 4d, the CoPH/ALO/BVO photoanode demonstrates a charge separation efficiency of 85.9%, which is much higher than that of BVO (23.8%), CoPH/BVO (78.4%), and ALO/BVO (45.9%) photoanodes. This result suggests the modification of CoPH/ALO can greatly promote surface charge separation. Moreover, the applied bias photon-to-current efficiencies (ABPEs) of BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes were calculated and presented in Figure 4e. The CoPH/ALO/BVO photoanode could achieve an ABPE of 1.47% at 0.76 VRHE, which is 4.6, 3.1, and 1.1 times that of BVO (0.32%, 0.86 VRHE), ALO/BVO (0.48%, 0.81 VRHE), and CoPH/BVO (1.38%, 0.75 VRHE) photoanodes, respectively. This finding confirms that the modification of CoPH/ALO on the BVO photoanode surface promotes the transfer of photogenerated holes to the photoanode/electrolyte and improves the energy conversion, which improves the PEC performance. This result is also consistent with the LSV measurement and long-time stability test. In addition, the incident photon-to-current conversion efficiency (IPCE) of these photoanodes was calculated as shown in Figure 4f. The IPCE of a CoPH/ALO/BVO photoanode can be up to 81.7% at 400 nm, which is much higher than 22.9%, 45.5%, and 65.7% for BVO, ALO/BVO, and CoPH/BVO photoanodes. These measurements express that the introduction of the CoPH cocatalyst and the ALO passivation layer can improve the surface charge transfer efficiency of BVO, resulting in superior photocurrent density, and the result that is also confirmed by LSV measurement.
In order to investigate the surface water oxidation kinetics and interfacial charge transfer of BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes, electrochemical impedance spectroscopy (EIS) measurement was performed [44]. The EIS measurements provided us with Nyquist plots for the BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes, as depicted in Figure 5a. These plots were analyzed using an equivalent mode. In this model, Rs and Rct represent the solution resistance and charge transfer resistance of the electrolyte, respectively. According to the fitted values of the BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes, CoPH/ALO/BVO demonstrates a lower Rct (92.0 Ω) than those of BVO (566.7 Ω), ALO/BVO (364.2 Ω), and CoPH/BVO (98.5 Ω), indicating that the introduced CoPH/ALO could facilitate the interfacial charge transfer and accelerate the water oxidation kinetics. This analysis is consistent with the order of LSV measurement. Our findings, as demonstrated by these EIS results, provide a deeper understanding of the mechanisms at play in these prepared photoanodes. This result is vital for the future design and optimization of photoanodes for PEC applications, paving the way for more efficient and sustainable solar fuel generation technologies. The Mott–Schottky (M-S) plots of the BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes are displayed in Figure 5b. All photoanodes exhibit positive slopes, indicating the n-type feature of the BVO semiconductor [46]. The charge carrier densities for the BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes were estimated to be 2.1 × 1018, 2.3 × 1018, 2.7 × 1018, and 3.2 × 1018 cm−3, respectively. The increased carrier density of CoPH/ALO/BVO can be attributed to the facilitated interfacial charge transfer by the introduction of CoPH/ALO, which is consistent with the EIS results. Additionally, the findings also confirm the positive effect that the introduction of the CoPH cocatalyst and the ALO passivation layer has on the BVO photoelectrode, significantly improving the PEC performance. The charge carrier densities of the BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes were estimated to be 2.1 × 1018, 2.3 × 1018, 2.7 × 1018, and 3.2 × 1018 cm−3, respectively. It can be observed that CoPH could obviously increase the carrier density. Notably, the Nd of the ALO/CoPH/BVO photoanode is 1.5 times that of the pristine BVO, indicating the enhanced electronic conductivity with the introduction of the ALO passivation layer and CoPH OEC. This enhancement could result in the accelerated electron–hole separation and promoted electron transport property.
Based on the above PEC analyses, we have proposed a mechanism for PEC water splitting as illustrated in Figure 6. Under light illumination, the electron–hole pairs are generated and separated at the BVO layer. The holes are transferred to the surface of the photoanode for the water oxidation, and the electrons are flowed to the Pt electrode for the hydrogen evolution. The observed PEC enhancement can be primarily ascribed to several critical factors. First of all, CoPH combines the advantages of cobalt oxide and cobalt phosphide to increase the conductivity, thus improving the photocurrent density [26]. This enhancement can be confirmed by LSV tests. Secondly, EIS and M-S measurements reveal the decreased transfer resistance and the increased carrier density with CoPH modification. This suggests that the CoPH incorporation accelerates the water oxidation kinetics and facilitates the interfacial charge transfer, further improving the PEC performance. However, CoPH/BVO still suffers from severe photocorrosion, leading to a rapid decay in photocurrent density. With the introduction of an ALO passivation layer, the stability can be obviously improved. It can ascribe to the reduced photocorrosion with the passivated surface defect state [28]. The presence of a negative charge on the ALO promotes the transfer of photogenerated holes, thereby inhibiting the surface charge recombination [47]. This improvement is evident from the ηsurface and EIS tests. In conclusion, the BVO photoanode achieves optimal PEC performance under the synergistic effect of the CoPH cocatalyst and the ALO passivation layer.

3. Experimental Section

3.1. Preparation of Photoanodes

3.1.1. Synthesis of BVO Photoanode

In a standard procedure, 3.32 g of KI were dissolved in 50 mL H2O, and the PH was adjusted to 1.7 using HNO3 while stirring in an ice bath. Subsequently, 970 mg of Bi(NO3)3·5H2O was added to the solution, and it was stirred for 1 h. Afterward, the solution was then mixed with 20 mL of absolute ethanol containing 0.23 M p-Benzoquinone and stirred for 1 h to create an electrolyte. For the electrodeposition, a typical three-electrode cell was utilized. The working electrode was conducted from FTO (1 cm × 2 cm), the reference electrode was Ag/AgCl (4 M KCl), and the counter electrode was Pt foil (1 cm × 1 cm). The deposition area on the FTO was confined to 1 cm × 1 cm. Electrodeposition was performed at −0.1 V vs. Ag/AgCl to produce BiOI, with a charge of 0.3 C cm−2 passing through the system. Subsequently, a 75 μL DMSO solution containing 0.2 M VO(acac)2 was applied to the BiOI surface, followed by calcinating at 450 °C for 2 h to convert BiOI to BVO. Following cooling to room temperature, the BVO material was immersed in a 1 M NaOH solution for 20 min under gentle agitation, then rinsed with DI water and dried.

3.1.2. Synthesis of ALO/BVO Photoanode

Al2O3 layer was synthesized by immersion–calcination method. The BVO photoanode was immersed in a 10 mM AlCl3 aqueous solution for 2 min and subsequently calcinated in air at 200 °C for 0.5 h.

3.1.3. Synthesis of CoPH/BVO Photoanode

CoPH/BVO was synthesized by hydrothermal method. 3 mM Co(NO3)2·6H2O and 2 mM K2HPO4 were mixed in 15 mL distilled water and stirred for 15 min. The reaction mixture and BVO photoanode were transferred to a 25 mL Teflon-lined hydrothermal vessel and then placed in an oven at 180 °C for 12 h. After the completion of the reaction, it was washed twice with distilled water and subsequently dried in an oven at 60 °C for 20 min.

3.1.4. Synthesis of CoPH/ALO/BVO Photoanode

The CoPH/ALO/BVO photoanode was prepared according to the above steps (synthesis of CoPH/BVO photoanode), except for changing BVO to ALO/BVO into Teflon-lined hydrothermal vessel.

3.1.5. Synthesis of CoBi/BVO Photoanode

First, CoBi/BVO photoanode was synthesized by a facile electrodeposition method in a standard three-electrode cell with 1 M potassium borate electrolyte and 0.5 mM Co(NO3)2·6H2O. Platinum wire and Ag/AgCl electrodes were employed as the counter and reference electrodes, respectively. Subsequently, under AM 1.5 G illumination, CoBi was electrodeposited on BVO photoanode at 1.2 VRHE for 5 s.

3.1.6. Synthesis of Co(OH)2/BVO Photoanode

First, BVO photoanode was immersed in a 5 mM CoCl2 6H2O solution for 10 h. After that, it was removed and dried in an oven at 60 °C for 30 min, and marked as Co(OH)2/BVO.

3.1.7. Synthesis of CoPi/BVO Photoanode

CoPi/BVO photoanode was synthesized with electrodeposition method. The electrolytes were prepared by dissolving 0.5 mM Co(NO3)2·6H2O in 0.1 M potassium phosphate (KPi) buffer solution. Electrodeposition was carried out at 1.5 VRHE for 30 s. After that, it was removed and dried in an oven at 60 °C for 30 min, and marked as CoPi/BVO.

3.2. Characterization

The crystallinity and structure were carried out by X-ray diffraction analysis (XRD, Rigaku Smart Lab, Tokyo, Japan). The elemental composition and chemical state of the materials were explored by X-ray photoelectron spectroscopy (XPS, Thermo Fisher Scientific K-Alpha+, Waltham, MA, USA). The C 1 s peak with a binding energy of 284.8 eV was used to correct the binding energies of other elements. Surface morphology and elemental distribution were further investigated by scanning electron microscopy (SEM, Zeiss Sigma300, Jena, Germany) equipped with an energy dispersive X-ray spectrometer (EDS). The samples were evaporated with Pt for 60 s before SEM measurements. Optical properties were analyzed by ultraviolet–visible spectroscopy (UV-Vis, PerkinElmer lambda 950, Hongkong, China).

3.3. Photoelectrochemical Measurement

All photoelectrochemical measurements were tested in a standard three-electrode cell with an electrochemical workstation (CHI760E, Shanghai Chenhua Instrument Co., Ltd., Shanghai, China). The photoanode was used as working electrode, Ag/AgCl (4 M KCl) was used as reference electrode, Pt foil (1 cm × 1 cm) was used as counter electrode, and 1 M KBi buffer (pH = 9) was used as electrolyte. A 300 W xenon lamp (PLS-SEX300, Beijing Perfect Light Technology Co., Ltd., Beijing, China) equipped with an AM 1.5 G filter was used as the light source, and the incident light intensity was calibrated to 100 mW/cm2. Linear sweep voltammogram (LSV) was measured at a scan rate of 10 mV s−1. The frequency range for photoelectrochemical impedance spectroscopy (PEIS) was 100 kHz to 0.1 Hz. The Nyquist plot was fitted to the corresponding equivalent circuit by Z-view 2 software.
The carrier density (Nd) is calculated using the following equation:
Nd = (2/e0εε0) [d(1/C2)/dV]−1
In the given equation, C indicates the capacitance of the space charge region, e0 indicates the electron charge (1.602 × 10−19 C), ε indicates the permittivity of BVO (68), ε0 is the vacuum permittivity (8.854 × 10−14 F/cm), and V is the potential applied to the electrode.
The applied bias photon current efficiency (ABPE) is computed using the equation provided:
ABPE = ((1.23 − Vb) × J/Plight) × 100%
where Vb denotes the applied potential versus RHE, while J is the observed photocurrent density (mA cm−2). Plight refers to the optical density (100 mW cm−2).
The incident photon-to-current conversion efficiency (IPCE) is calculated according to the formula:
IPCE = (1240 × J)/(λ × Plight) × 100%
In the formula, J represents the photocurrent density (mA cm−2) measured at a certain wavelength, λ is the wavelength of the incident light (nm), and Plight refers to the intensity of the power at a specific wavelength (mW cm−2).
The surface charge separation efficiency (ηsurface) is calculated based on the following formula:
ηsurface = JH2O/JNa2SO3
where JH2O refers to the photocurrent density of PEC water oxidation, while JNa2SO3 is the photocurrent density achieved when using Na2SO3 as the sacrificial agent.

4. Conclusions

In summary, we have successfully prepared a CoPH/ALO/BVO photoanode, demonstrating a significant enhancement in PEC performance. The introduction of the CoPH cocatalyst and the ALO passivation layer onto the BVO photoanode has resulted in a remarkable improvement in efficiency and stability in PEC water splitting. The CoPH/ALO/BVO photoanode achieves an impressive photocurrent density of 4.9 mA cm−2 at 1.23 VRHE, representing a profound 3.3-fold enhancement over the pristine BVO. This notable enhancement in PEC performance is primarily due to the combination of the passivation effect induced by the catalytical activity provided by CoPH and the ALO layer. This work offers a new idea to design the high-quality photoanode with the synergetic strategy combining the OEC and passivation effect.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29030683/s1, Figure S1: Optical images of these photoanodes; Figure S2: XRD patterns of CoPH; Figure S3: UV-Vis spectrum of CoPH; Figure S4: SEM images of CoPH/BVO photoanodes with different concentrations of CoPH; Figure S5: EDX pattern of CoPH/ALO/BVO photoanode; Table S1: Elemental composition of ALO/CoPH/BVO; Figure S6: LSV curves of BVO photoanodes with different concentrations of CoPH; Table S2: The PEC performance comparison between previous reports and this work.

Author Contributions

Methodology, Z.S. and Y.W.; Investigation, Z.L. (Zhen Li) and Y.Y.; Data curation, J.C. and Y.L.; Writing—original draft, Z.L. (Zhen Li) and Z.L. (Zhenhong Lu); Writing—review & editing, Z.S. and X.H.; Visualization, C.S.; Project administration, Y.W.; Funding acquisition, Z.S. and X.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work is financially supported by the Natural Science Foundation of Guangxi Province (2021GXNSFBA075025 and 2023GXNSFBA026182), the Basic Ability Enhancement Program for Young and Middle-aged Teachers of Guangxi (2021KY0355), and the Doctoral Foundation of Guangxi University of Science and Technology (18Z12, 19Z23, and 20Z41).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Acknowledgments

The authors would like to thank Yu Yang from Shiyanjia Lab (www.shiyanjia.com, accessed on 17 November 2021) for the SEM characterization.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Le, H.V.; Nguyen, M.D.; Pham, Y.T.H.; Nguyen, D.N.; Le, L.T.; Han, H.; Tran, P.D. Decoration of AgOx hole collector to boost photocatalytic water oxidation activity of BiVO4 photoanode. Mater. Today Energy 2021, 21, 9. [Google Scholar] [CrossRef]
  2. Majumder, S.; Gu, M.; Kim, K.H. Facile fabrication of BiVO4/Bi2S3/NiCoO2 for significant photoelectrochemical water splitting. Appl. Surf. Sci. 2022, 574, 14. [Google Scholar] [CrossRef]
  3. Zahraei, A.A.; Yaghobi, R.; Golmohammad, M. In situ synthesize Bi nanostructures on copper foam toward efficient electrocatalytic hydrogen evolution reaction. Chem. Pap. 2023, 77, 859–866. [Google Scholar] [CrossRef]
  4. Zhou, T.S.; Chen, S.; Wang, J.C.; Zhang, Y.; Li, J.H.; Bai, J.; Zhou, B.X. Dramatically enhanced solar-driven water splitting of BiVO4 photoanode via strengthening hole transfer and light harvesting by co-modification of CQDs and ultrathin β-FeOOH layers. Chem. Eng. J. 2021, 403, 13. [Google Scholar] [CrossRef]
  5. Wang, J.M.; Kuo, M.T.; Zeng, P.; Xu, L.; Chen, S.T.; Peng, T.Y. Few-layer BiVO4 nanosheets decorated with SrTiO3: Rh nanoparticles for highly efficient visible-light-driven overall water splitting. Appl. Catal. B-Environ. 2020, 279, 12. [Google Scholar] [CrossRef]
  6. Xu, D.B.; Xia, T.; Xu, H.M.; Fan, W.Q.; Shi, W.D. Synthesis of ternary spinel MCo2O4 (M = Mn, Zn)/BiVO4 photoelectrodes for photolectrochemical water splitting. Chem. Eng. J. 2020, 392, 9. [Google Scholar] [CrossRef]
  7. Huang, J.W.; Liu, T.T.; Wang, R.F.; Zhang, M.Y.; Wang, L.; She, H.D.; Wang, Q.Z. Facile loading of cobalt oxide on bismuth vanadate: Proved construction of p-n junction for efficient photoelectrochemical water oxidation. J. Colloid Interface Sci. 2020, 570, 89–98. [Google Scholar] [CrossRef]
  8. Zhang, B.B.; Chou, L.J.; Bi, Y.P. Tuning surface electronegativity of BiVO4 photoanodes toward high-performance water splitting. Appl. Catal. B-Environ. 2020, 262, 6. [Google Scholar] [CrossRef]
  9. Zhang, B.B.; Huang, X.J.; Zhang, Y.; Lu, G.X.; Chou, L.J.; Bi, Y.P. Unveiling the activity and stability origin of BiVO(4) photoanodes with FeNi oxyhydroxides for oxygen evolution. Angew. Chem.-Int. Ed. 2020, 59, 18990–18995. [Google Scholar] [CrossRef]
  10. Dabodiya, T.S.; Selvarasu, P.; Murugan, A.V. Tetragonal to monoclinic crystalline phases change of BiVO4 via microwave-hydrothermal reaction: In correlation with visible-light-driven photocatalytic performance. Inorg. Chem. 2019, 58, 5096–5110. [Google Scholar] [CrossRef]
  11. Parmar, K.P.S.; Kang, H.J.; Bist, A.; Dua, P.; Jang, J.S.; Lee, J.S. Photocatalytic and photoelectrochemical water oxidation over metal-doped monoclinic BiVO4 photoanodes. ChemSusChem 2012, 5, 1926–1934. [Google Scholar] [CrossRef] [PubMed]
  12. Zhao, W.; Tu, X.Y.; Wang, X.M.; Dai, B.L.; Zhang, L.L.; Xu, J.M.; Feng, Y.; Sheng, N.; Zhu, F.X. Novel p-n heterojunction photocatalyst fabricated by flower-like BiVO4 and Ag2S nanoparticles: Simple synthesis and excellent photocatalytic performance. Chem. Eng. J. 2019, 361, 1173–1181. [Google Scholar] [CrossRef]
  13. Zhong, M.; Hisatomi, T.; Kuang, Y.B.; Zhao, J.; Liu, M.; Iwase, A.; Jia, Q.X.; Nishiyama, H.; Minegishi, T.; Nakabayashi, M.; et al. Surface modification of CoOx loaded BiVO4 photoanodes with ultrathin p-type NiO layers for improved solar water oxidation. J. Am. Chem. Soc. 2015, 137, 5053–5060. [Google Scholar] [CrossRef] [PubMed]
  14. Kim, T.W.; Choi, K.S. Nanoporous BiVO4 photoanodes with dual-layer oxygen evolution catalysts for solar water splitting. Science 2014, 343, 990–994. [Google Scholar] [CrossRef] [PubMed]
  15. Ye, K.H.; Wang, Z.L.; Gu, J.W.; Xiao, S.; Yuan, Y.F.; Zhu, Y.; Zhang, Y.M.; Mai, W.J.; Yang, S.H. Carbon quantum dots as a visible light sensitizer to significantly increase the solar water splitting performance of bismuth vanadate photoanodes. Energy Environ. Sci. 2017, 10, 772–779. [Google Scholar] [CrossRef]
  16. Zhang, S.C.; Liu, Z.F.; Ruan, M.N.; Guo, Z.G.; Lei, E.; Zhao, W.; Zhao, D.; Wu, X.F.; Chen, D.M. Enhanced piezoelectric-effect-assisted photoelectrochemical performance in ZnO modified with dual cocatalysts. Appl. Catal. B-Environ. 2020, 262, 10. [Google Scholar] [CrossRef]
  17. Lu, W.X.; Wang, B.; Chen, W.J.; Xie, J.L.; Huang, Z.Q.; Jin, W.; Song, J.L. Nanosheet-like Co3(OH)2(HPO4)2 as a highly efficient and stable electrocatalyst for oxygen evolution reaction. ACS Sustain. Chem. Eng. 2019, 7, 3083–3091. [Google Scholar] [CrossRef]
  18. Khiarak, B.N.; Imanparast, S.; Yengejeh, M.M.; Zahraei, A.A.; Yaghobi, R.; Golmohammad, M. Efficient water oxidation catalyzed by a graphene oxide/copper electrode, supported on carbon cloth. Russ. J. Electrochem. 2021, 57, 1196–1206. [Google Scholar] [CrossRef]
  19. Jin, H.Y.; Mao, S.J.; Zhan, G.P.; Xu, F.; Bao, X.B.; Wang, Y. Fe incorporated α-Co(OH)2 nanosheets with remarkably improved activity towards the oxygen evolution reaction. J. Mater. Chem. A 2017, 5, 1078–1084. [Google Scholar] [CrossRef]
  20. Ning, X.M.; Yin, D.; Fan, Y.P.; Zhang, Q.; Du, P.Y.; Zhang, D.X.; Chen, J.; Lu, X.Q. Plasmon-Enhanced Charge Separation and Surface Reactions Based on Ag-Loaded Transition-Metal Hydroxide for Photoelectrochemical Water Oxidation. Adv. Energy Mater. 2021, 11, 8. [Google Scholar] [CrossRef]
  21. Bu, Q.J.; Li, S.; Zhang, K.; Lin, Y.H.; Wang, D.J.; Zou, X.X.; Xie, T.F. Hole transfer channel of ferrihydrite designed between Ti-Fe2O3 and CoPi as an efficient and durable photoanode. ACS Sustain. Chem. Eng. 2019, 7, 10971–10978. [Google Scholar] [CrossRef]
  22. Hernández, S.; Gerardi, G.; Bejtka, K.; Fina, A.; Russo, N. Evaluation of the charge transfer kinetics of spin-coated BiVO4 thin films for sun-driven water photoelectrolysis. Appl. Catal. B-Environ. 2016, 190, 66–74. [Google Scholar] [CrossRef]
  23. Reddy, D.A.; Kim, Y.; Shim, H.S.; Reddy, K.A.J.; Gopannagari, M.; Kumar, D.P.; Song, J.K.; Kim, T.K. Significant Improvements on BiVO4@CoPi Photoanode Solar Water Splitting Performance by Extending Visible-Light Harvesting Capacity and Charge Carrier Transportation. ACS Appl. Energy Mater. 2020, 3, 4474–4483. [Google Scholar] [CrossRef]
  24. Kumar, R.; Inta, H.R.; Koppisetti, H.; Ganguli, S.; Ghosh, S.; Mahalingam, V. Electrochemical reconstruction of Zn0.3Co2.7(PO4)2•4H2O for enhanced water oxidation performance. ACS Appl. Energy Mater. 2020, 3, 12088–12098. [Google Scholar] [CrossRef]
  25. Zhang, S.C.; Liu, Z.F.; Chen, D.; Yan, W.G. An efficient hole transfer pathway on hematite integrated by ultrathin Al2O3 interlayer and novel CuCoOx cocatalyst for efficient photoelectrochemical water oxidation. Appl. Catal. B-Environ. 2020, 277, 9. [Google Scholar] [CrossRef]
  26. Lu, P.C.; Chen, Y.F.; Zhou, R.Y.; Guo, C.C.; Liu, X.R.; Yang, F.F.; Zhu, Y.R. Novel bouquet-like cobalt phosphate as an ultrahigh-rate and durable battery-type cathode material for hybrid supercapacitors. Sci. China-Mater. 2022, 65, 1503–1511. [Google Scholar] [CrossRef]
  27. Kafizas, A.; Xing, X.T.; Selim, S.; Mesa, C.A.; Ma, Y.M.; Burgess, C.; McLachlan, M.A.; Durrant, J.R. Ultra-thin Al2O3 coatings on BiVO4 photoanodes: Impact on performance and charge carrier dynamics. Catal. Today 2019, 321, 59–66. [Google Scholar] [CrossRef]
  28. Chang, G.L.; Wang, D.G.; Zhang, Y.Y.; Aldalbahi, A.; Wang, L.H.; Li, Q.; Wang, K. ALD-coated ultrathin Al2O3 film on BiVO4 nanoparticles for efficient PEC water splitting. Nucl. Sci. Tech. 2016, 27, 6. [Google Scholar] [CrossRef]
  29. Ma, Z.Z.; Song, K.; Wang, L.; Gao, F.M.; Tang, B.; Hou, H.L.; Yang, W.Y. WO3/BiVO4 type-II heterojunction arrays decorated with oxygen-deficient ZnO passivation layer: A highly efficient and stable photoanode. ACS Appl. Mater. Interfaces 2019, 11, 889–897. [Google Scholar] [CrossRef]
  30. Malara, F.; Fabbri, F.; Marelli, M.; Naldoni, A. Controlling the surface energetics and kinetics of hematite photoanodes through few stomic layers of NiOx. ACS Catal. 2016, 6, 3619–3628. [Google Scholar] [CrossRef]
  31. Balaji, N.; Park, C.; Raja, J.; Ju, M.; Venkatesan, M.R.; Lee, H.; Yi, J. Low surface recombination velocity on P-Type Cz-Si surface by sol-gel deposition of Al2O3 films for solar cell applications. J. Nanosci. Nanotechnol. 2015, 15, 5123–5128. [Google Scholar] [CrossRef] [PubMed]
  32. Ma, C.; Din, S.T.U.; Seo, W.C.; Lee, J.; Kim, Y.; Jung, H.; Yang, W. BiVO4 ternary photocatalyst co-modified with N-doped graphene nanodots and Ag nanoparticles for improved photocatalytic oxidation: A significant enhancement in photoinduced carrier separation and broad-spectrum light absorption. Sep. Purif. Technol. 2021, 264, 118423. [Google Scholar] [CrossRef]
  33. Xu, C.J.; Sun, W.J.; Dong, Y.J.; Dong, C.Z.; Hu, Q.Y.; Ma, B.C.; Ding, Y. A graphene oxide-molecular Cu porphyrin-integrated BiVO4 photoanode for improved photoelectrochemical water oxidation performance. J. Mater. Chem. A 2020, 8, 4062–4072. [Google Scholar] [CrossRef]
  34. Fan, K.; Chen, H.; He, B.W.; Yu, J.G. Cobalt polyoxometalate on N-doped carbon layer to boost photoelectrochemical water oxidation of BiVO4. Chem. Eng. J. 2020, 392, 123744. [Google Scholar] [CrossRef]
  35. Zhang, H.; Yu, Y.; Zhang, L.L.; Dong, S.J. Fuel-free Bio-photoelectrochemical cells based on a water/oxygen circulation system with a Ni:FeOOH/BiVO4 photoanode. Angew. Chem. Int. Ed. 2018, 57, 1547–1551. [Google Scholar] [CrossRef]
  36. Elkabouss, K.; Kacimi, M.; Ziyad, M.; Ammar, S.; Bozon-Verduraz, F. Cobalt-exchanged hydroxyapatite catalysts: Magnetic studies, spectroscopic investigations, performance in 2-butanol and ethane oxidative dehydrogenations. J. Catal. 2004, 226, 16–24. [Google Scholar] [CrossRef]
  37. Gogoi, D.; Koyani, R.; Golder, A.K.; Peela, N.R. Enhanced photocatalytic hydrogen evolution using green carbon quantum dots modified 1-D CdS nanowires under visible light irradiation. Sol. Energy 2020, 208, 966–977. [Google Scholar] [CrossRef]
  38. Ai, L.H.; Niu, Z.G.; Jiang, J. Mechanistic insight into oxygen evolution electrocatalysis of surface phosphate modified cobalt phosphide nanorod bundles and their superior performance for overall water splitting. Electrochim. Acta 2017, 242, 355–363. [Google Scholar] [CrossRef]
  39. Kuang, P.Y.; Zhang, L.Y.; Cheng, B.; Yu, J.G. Enhanced charge transfer kinetics of Fe2O3/CdS composite nanorod arrays using cobalt-phosphate as cocatalyst. Appl. Catal. B 2017, 218, 570–580. [Google Scholar] [CrossRef]
  40. Gao, J.; Sun, X.W.; Wang, Y.F.; Li, Y.; Li, X.C.; Chen, C.Z.; Ni, J.B. Ultrathin Al2O3 passivation layer-wrapped Ag@TiO2 nanorods by atomic layer deposition for enhanced photoelectrochemical performance. Appl. Surf. Sci. 2020, 499, 143971. [Google Scholar] [CrossRef]
  41. Huang, J.; Xiong, Y.S.; Peng, Z.Y.; Chen, L.F.; Wang, L.; Xu, Y.Z.; Tan, L.C.; Yuan, K.; Chen, Y.W. A general electrodeposition strategy for fabricating ultrathin nickel cobalt phosphate nanosheets with ultrahigh capacity and rate performance. ACS Nano 2020, 14, 14201–14211. [Google Scholar] [CrossRef] [PubMed]
  42. Ning, X.M.; Lu, B.Z.; Zhang, Z.; Du, P.Y.; Ren, H.X.; Shan, D.L.; Chen, J.; Gao, Y.J.; Lu, X.Q. An efficient strategy for boosting photogenerated charge separation by using porphyrins as interfacial charge mediators. Angew. Chem.-Int. Ed. 2019, 58, 16800–16805. [Google Scholar] [CrossRef] [PubMed]
  43. Li, X.T.; Jia, M.L.; Lu, Y.T.; Li, N.; Zheng, Y.Z.; Tao, X.; Huang, M.L. Co(OH)2/BiVO4 photoanode in tandem with a carbon-based perovskite solar cell for solar-driven overall water splitting. Electrochim. Acta 2020, 330, 9. [Google Scholar] [CrossRef]
  44. Wang, S.C.; Chen, P.; Yun, J.H.; Hu, Y.X.; Wang, L.Z. An electrochemically treated BiVO4 photoanode for efficient photoelectrochemical water splitting. Angew. Chem.-Int. Ed. 2017, 56, 8500–8504. [Google Scholar] [CrossRef] [PubMed]
  45. Zachäus, C.; Abdi, F.F.; Peter, L.M.; van de Krol, R. Photocurrent of BiVO4 is limited by surface recombination, not surface catalysis. Chem. Sci. 2017, 8, 3712–3719. [Google Scholar] [CrossRef]
  46. Li, Y.; Mei, Q.; Liu, Z.J.; Hu, X.S.; Zhou, Z.H.; Huang, J.W.; Bai, B.; Liu, H.; Ding, F.; Wang, Q.Z. Fluorine-doped iron oxyhydroxide cocatalyst: Promotion on the WO3 photoanode conducted photoelectrochemical water splitting. Appl. Catal. B-Environ. 2022, 304, 11. [Google Scholar] [CrossRef]
  47. Zhao, Y.L.; Xie, H.C.; Shi, W.W.; Wang, H.; Shao, C.Y.; Li, C. Unravelling the essential difference between TiOx and AlOx interface layers on Ta3N5 photoanode for photoelectrochemical water oxidation. J. Energy Chem. 2022, 64, 33–37. [Google Scholar] [CrossRef]
Figure 1. Schematic of the synthetic procedure for the CoPH/ALO/BVO photoanode.
Figure 1. Schematic of the synthetic procedure for the CoPH/ALO/BVO photoanode.
Molecules 29 00683 g001
Figure 2. (a) XRD patterns, (b) Raman spectra, and (c) UV-Vis diffused reflectance spectra of BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes. SEM images of (d) BVO, (e) ALO/BVO, and (f) CoPH/ALO/BVO photoanodes. (g) HRTEM images and (h) EDX mapping of CoPH/ALO/BVO photoanode.
Figure 2. (a) XRD patterns, (b) Raman spectra, and (c) UV-Vis diffused reflectance spectra of BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes. SEM images of (d) BVO, (e) ALO/BVO, and (f) CoPH/ALO/BVO photoanodes. (g) HRTEM images and (h) EDX mapping of CoPH/ALO/BVO photoanode.
Molecules 29 00683 g002
Figure 3. XPS spectra of BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes. (a) Bi 4f, (b) V 2p, (c) O 1s, (d) P 2p, (e) Al 2p, and (f) Co 2p.
Figure 3. XPS spectra of BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes. (a) Bi 4f, (b) V 2p, (c) O 1s, (d) P 2p, (e) Al 2p, and (f) Co 2p.
Molecules 29 00683 g003
Figure 4. (a,b) LSV curves of different photoanodes. (c) Long-time stability test, (d) ηsurface, (e) ABPE, and (f) IPCE curves of BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes.
Figure 4. (a,b) LSV curves of different photoanodes. (c) Long-time stability test, (d) ηsurface, (e) ABPE, and (f) IPCE curves of BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes.
Molecules 29 00683 g004
Figure 5. (a) EIS Nyquist plots, and (b) M-S plots of BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes.
Figure 5. (a) EIS Nyquist plots, and (b) M-S plots of BVO, ALO/BVO, CoPH/BVO, and CoPH/ALO/BVO photoanodes.
Molecules 29 00683 g005
Figure 6. A proposed mechanism of the CoPH/ALO/BVO photoanode.
Figure 6. A proposed mechanism of the CoPH/ALO/BVO photoanode.
Molecules 29 00683 g006
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

Sun, Z.; Li, Z.; Chen, J.; Yang, Y.; Su, C.; Lv, Y.; Lu, Z.; He, X.; Wang, Y. Synergistic Effect of Co3(HPO4)2(OH)2 Cocatalyst and Al2O3 Passivation Layer on BiVO4 Photoanode for Enhanced Photoelectrochemical Water Oxidation. Molecules 2024, 29, 683. https://doi.org/10.3390/molecules29030683

AMA Style

Sun Z, Li Z, Chen J, Yang Y, Su C, Lv Y, Lu Z, He X, Wang Y. Synergistic Effect of Co3(HPO4)2(OH)2 Cocatalyst and Al2O3 Passivation Layer on BiVO4 Photoanode for Enhanced Photoelectrochemical Water Oxidation. Molecules. 2024; 29(3):683. https://doi.org/10.3390/molecules29030683

Chicago/Turabian Style

Sun, Zijun, Zhen Li, Jinlin Chen, Yuying Yang, Chunrong Su, Yumin Lv, Zhenhong Lu, Xiong He, and Yongqing Wang. 2024. "Synergistic Effect of Co3(HPO4)2(OH)2 Cocatalyst and Al2O3 Passivation Layer on BiVO4 Photoanode for Enhanced Photoelectrochemical Water Oxidation" Molecules 29, no. 3: 683. https://doi.org/10.3390/molecules29030683

Article Metrics

Back to TopTop