Next Article in Journal
Nanoparticles Enhance In Vitro Micropropagation and Secondary Metabolite Accumulation in Origanum petraeum
Previous Article in Journal
Single-Layer Full-Color Waveguide Display Based on a Broadband Efficient Meta-Grating
Previous Article in Special Issue
Tailoring Co Distribution in PtCo Alloys for Enhanced Oxygen Reduction Reaction Activity and Durability in Fuel Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Strategies for Enhancing BiVO4 Photoanodes for PEC Water Splitting: A State-of-the-Art Review

Department of Chemical Engineering, Konkuk University, Seoul 05029, Republic of Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2025, 15(19), 1494; https://doi.org/10.3390/nano15191494
Submission received: 22 August 2025 / Revised: 17 September 2025 / Accepted: 24 September 2025 / Published: 30 September 2025

Abstract

Bismuth vanadate (BiVO4) has attracted significant attention as a photoanode material for photoelectrochemical (PEC) water splitting due to its suitable bandgap (~2.4 eV), strong visible light absorption, chemical stability, and cost-effectiveness. Despite these advantages, its practical application remains constrained by intrinsic limitations, including poor charge carrier mobility, short diffusion length, and sluggish oxygen evolution reaction (OER) kinetics. This review critically summarizes recent advancements aimed at enhancing BiVO4 PEC performance, encompassing synthesis strategies, defect engineering, heterojunction formation, cocatalyst integration, light-harvesting optimization, and stability improvements. Key fabrication methods—such as solution-based, vapor-phase, and electrochemical approaches—along with targeted modifications, including metal/nonmetal doping, surface passivation, and incorporation of electron transport layers, are discussed. Emphasis is placed on strategies to improve light absorption, charge separation efficiency (ηsep), and charge transfer efficiency (ηtrans) through bandgap engineering, optical structure design, and catalytic interface optimization. Approaches to enhance stability via protective overlayers and electrolyte tuning are also reviewed, alongside emerging applications of BiVO4 in tandem PEC systems and selective solar-driven production of value-added chemicals, such as H2O2. Finally, critical challenges, including the scale-up of electrode fabrication and the elucidation of fundamental reaction mechanisms, are highlighted, providing perspectives for bridging the gap between laboratory performance and practical implementation.

1. Introduction

The escalating concerns over global warming and the depletion of fossil fuels have intensified the demand for clean and renewable energy sources [1,2,3,4]. In this context, hydrogen has emerged as a promising energy carrier due to its high energy density and environmentally benign characteristics [5,6,7,8]. In particular, photoelectrochemical (PEC) water splitting utilizing solar energy has attracted considerable interest as a sustainable approach for hydrogen production. PEC water splitting provides an efficient and cost-effective route to directly convert solar energy into chemical energy, exhibiting higher energy conversion efficiency compared to conventional electrolysis methods [9,10,11].
In PEC systems, the selection of photoelectrode materials is critical for achieving high efficiency and stability. An ideal photoelectrode should possess strong visible light absorption, efficient charge separation and transport, appropriate band alignment, and chemical stability in aqueous environments [4,12,13,14]. Among various semiconductors, bismuth vanadate (BiVO4) has drawn substantial attention as a material that meets these requirements [15,16,17].
BiVO4 exhibits several advantageous properties for PEC hydrogen production, including a suitable bandgap (~2.4 eV), strong visible light absorption, cost-effectiveness, and environmental benignity [18,19,20,21]. Its band structure is particularly favorable for water oxidation, making BiVO4 a widely studied n-type photoelectrode [22,23,24]. Moreover, BiVO4 demonstrates high chemical stability, resulting in minimal performance degradation under long-term operation [25,26,27].
Despite these promising characteristics, BiVO4 faces intrinsic limitations that hinder its practical application. Low charge carrier mobility, short minority carrier diffusion length, and sluggish oxygen evolution reaction (OER) kinetics are major challenges [28,29,30], leading to a notable gap between theoretical and actual PEC performance. To overcome these limitations, various strategies have been explored, including nanostructure engineering, heteroatom doping, heterojunction formation, cocatalyst integration, and surface passivation [31,32,33].
Nanostructure engineering increases the surface area and shortens charge transfer distances, thereby enhancing the photocurrent [34,35]. Heteroatom doping improves electrical conductivity and tunes the band structure [36,37]. Heterojunction formation promotes charge separation and extends light absorption [38,39]. Cocatalyst integration lowers the overpotential for OER and accelerates reaction rates [40,41], while surface passivation mitigates surface defects and suppresses charge recombination [42,43]. These strategies are often combined to achieve significant improvements in BiVO4 PEC performance.
This review aims to provide a comprehensive overview of recent research trends in BiVO4-based PEC hydrogen production. We discussed the fundamental properties of BiVO4, synthesis methods, performance enhancement strategies, and PEC system configurations.

2. Principle of Operation for Photoanode in PEC Water Splitting

Electrochemical (EC) water splitting is initiated by photon absorption in the semiconductor material of the photoelectrode. For this process to occur, incident photons must possess sufficient energy to excite electrons from the valence band (VB) to the conduction band (CB), i.e., when the photon energy () is equal to or greater than the semiconductor’s bandgap energy (Eg), generating electron–hole pairs that drive subsequent chemical reactions:
S C + h v E g e C B + h V B + : L i g h t   a b s o r p t i o n
The photogenerated charge carriers participate in spatially separated redox reactions. Electrons in the CB (eCB) with potentials more negative than 0.0 VRHE migrate to the photoelectrode surface and reduce protons to hydrogen, typically facilitated by a hydrogen evolution cocatalyst (HEC):
4 H + + 4 e C B 2 H 2   : H E R
Simultaneously, holes in the VB (hVB+) with potentials more positive than 1.23 VRHE can oxidize water to oxygen, typically facilitated by an oxygen evolution cocatalyst (OEC):
2 H 2 O + 4 h V B + O 2 + 4 H + : O E R
Overall, water splitting proceeds as
2 H 2 O 2 H 2 + O 2 ,   G ° = 238   k J   m o l 1
To theoretically split water, a semiconductor must satisfy two criteria: (i) a bandgap energy (Eg) exceeding 1.23 eV, and (ii) a CB edge more negative than 0.0 VRHE and a VB edge more positive than 1.23 VRHE [44]. In practice, overpotentials associated with the hydrogen and oxygen evolution reactions—approximately 0.05 V for HER and 0.25 V for OER—require a larger effective bandgap (>1.5 eV) to drive the reactions efficiently.
Hydrogen production efficiency is often limited by electron–hole recombination, which reduces the availability of charge carriers for water splitting reactions. PEC cells employ a dual-electrode configuration, producing O2 at the anode and H2 at the cathode [45]. This spatial separation of oxidation and reduction reactions enhances energy conversion efficiency, improves operational stability, and reduces recombination losses compared to conventional photocatalytic systems using dispersed semiconductor powders. By maintaining distinct reaction sites, PEC cells optimize light absorption and charge carrier dynamics, enabling more reliable hydrogen production.
Figure 1 illustrates that when the semiconductor absorbs light, electron–hole pairs are generated in the CB and VB, respectively. Due to the band bending at the semiconductor–electrolyte interface, holes migrate toward the surface, where they initiate the OER, oxidizing water. Concurrently, the photogenerated electrons are transported through a transparent conducting substrate, such as fluorine-doped tin oxide (FTO), and collected into an external circuit. An applied potential (Eapp) is often used to provide the necessary energetic boost for the electrons, enabling them to travel to the metallic cathode. At the cathode, these electrons drive the hydrogen evolution reaction (HER), leading to hydrogen production. This synchronized process of water oxidation and reduction effectively converts light energy into chemical energy.
Figure 1. (a) Schematic of PEC cell in two-electrode configuration. Graphic visualization of PEC energy diagrams for PEC water splitting using (b) a photoanode; (c) a photocathode. Reprinted with permission from Ref. [46], Copyright © 2022, ECS Advances.
Figure 1. (a) Schematic of PEC cell in two-electrode configuration. Graphic visualization of PEC energy diagrams for PEC water splitting using (b) a photoanode; (c) a photocathode. Reprinted with permission from Ref. [46], Copyright © 2022, ECS Advances.
Nanomaterials 15 01494 g001
This bias voltage is generally supplied by a photovoltaic (PV) cell positioned behind the photoanode, which captures any transmitted photons and converts them into electrical power. In this tandem configuration, the PV device enhances the overall efficiency of the PEC system. The primary goal of this design is to achieve effective and sustainable hydrogen generation via water splitting.
To better evaluate the efficiency and intrinsic limitations of charge transfer in photoelectrodes, a well-established experimental methodology allows for the quantification of key performance parameters, including the photocurrent density (Jph), absorbed photon flux (Jabs), and the rates of surface recombination (Jsr) and bulk recombination (Jbr) [17]. Photogenerated charge carriers on a photoanode face two competing fates: they can either contribute to the desired photocurrent or undergo recombination, which constitutes a major energy loss pathway. Surface recombination occurs when charges recombine at the interface before reacting with the electrolyte, whereas bulk recombination refers to the loss of carriers within the semiconductor material. The balance between these two pathways is a critical factor determining the overall performance and efficiency of the photoanode in solar-driven water splitting.
When highly reactive species, such as hydrogen peroxide (H2O2) or sodium sulfite (Na2SO3), are introduced as sacrificial reagents, the surface recombination rate is significantly mitigated. Due to their strong tendency to undergo oxidation, these species provide kinetically favorable pathways for hole transfer to the electrolyte. Under such conditions, photogenerated holes are efficiently consumed in oxidation reactions, preventing their recombination with electrons at the photoanode surface.
By analyzing these recombination pathways, researchers can gain insights into the key factors limiting charge carrier dynamics, which in turn determine the overall efficiency of a photoelectrode. During the water oxidation reaction, the net photocurrent—denoted as JH2O—reflects the portion of absorbed photon flux (Jabs) that successfully contributes to oxygen evolution after accounting for losses due to surface and bulk recombination:
J H 2 O = J a b s ( J s r + J b r )
The establishment of a semiconductor–electrolyte junction, commonly referred to as a semiconductor–liquid (SCL) junction, is a crucial process in PEC cells featuring photoelectrodes without a buried junction. The SCL junction facilitates the separation of photogenerated electron–hole pairs at the interface, directing electrons toward the external circuit and holes toward the electrolyte. Unlike buried junctions formed by stacking two semiconductors, the SCL junction operates directly at the semiconductor–electrolyte boundary, presenting unique challenges and opportunities for optimizing PEC efficiency and stability.

3. Synthesis Methods of BiVO4 Photoelectrodes

The synthesis method plays a critical role in achieving high-efficiency photoelectrodes [16], as it directly determines the nature and concentration of defects within the material. These defects, in turn, influence the overall conductivity, its type, and the effectiveness of any applied modifications. Table 1 provides a summary of common synthetic techniques and the resulting morphologies of BiVO4. The fabrication of BiVO4 on substrates—primarily fluorine-doped tin oxide (FTO)—typically begins with the deposition of bismuth (Bi) and vanadium (V) precursors, which are subsequently converted in situ into a BiVO4 film at temperatures ranging from 400 to 500 °C, either in air or under a controlled atmosphere.
Most BiVO4 films produced by this method exhibit a monoclinic phase with relatively poor crystallinity, though some show characteristics of a tetragonal phase, as indicated by the weak splitting of the XRD peak around 2θ ≈ 19° [55]. Despite the low crystallinity, BiVO4 possesses “defect tolerance” [2], which allows it to maintain efficient charge separation. As a result, crystallinity is not the primary factor determining the performance of BiVO4 in PEC applications.
Figure 2 presents representative exemplars for selected synthesis routes discussed in Section 3.1,Section 3.2,Section 3.3,Section 3.4,Section 3.5; methods that are not depicted in Figure 2 are referenced textually
Mapping of panels to sections: Figure 2a→Section 3.1 (Hydrothermal), Figure 2b→Section 3.2 (Electrodeposition), Figure 2c→Section 3.3 (BiOI conversion), Figure 2d→Section 3.4 (Sol–gel). Methods in Section 3.5 (e.g., aerosol/ALD) are illustrated in Figure 2e,f.

3.1. Preparation of Nanoporous BiVO4 Films

A representative hydrothermal morphology (nanoporous BiVO4) is shown in Figure 2a, a variety of solution-based approaches—such as soaking, dip coating, and impregnation—have been extensively employed for the preparation of BiVO4 photoanodes [56,57,58,59]. In the soaking method, the substrate is immersed in a precursor solution for an extended duration, allowing surface chemical reactions to occur; although straightforward, this technique often requires long processing times [56]. Dip-coating provides greater control by immersing the substrate in the solution followed by withdrawal at a regulated speed, enabling tunable film thickness through adjustment of the withdrawal rate [57]. Impregnation involves introducing the precursor solution into porous supports, which are subsequently dried and calcined, making this approach particularly effective for producing catalyst-supported BiVO4 materials with high specific surface area [58,59].

3.2. Spin Coating, Drop-Casting, Spray Pyrolysis, and Electrospray Deposition

An electrodeposited BiOI precursor prior to conversion is illustrated in Figure 2; and thin-film deposition techniques—including spin coating, drop-casting, spray pyrolysis, and electrospray deposition (ESD)—have also been applied in BiVO4 fabrication [20,60,61,62]. In spin coating, a small quantity of precursor solution is dispensed onto a rotating substrate, where centrifugal forces ensure the formation of a uniform thin film [60]. Drop-casting involves placing droplets of the precursor solution onto the substrate and allowing them to dry under ambient conditions; however, this method can result in film thickness variations [20]. Spray pyrolysis directs an aerosolized precursor onto a heated substrate, initiating pyrolysis and forming a continuous thin film [61]. Electrospray deposition (ESD) employs a high-voltage electric field to generate a fine mist of the precursor solution, enabling precise deposition of nanoparticles or uniform films onto the substrate surface [62].
Figure 2. Representative morphologies for selected BiVO4 synthesis routes. (a) Automatic dip coating machine. Reprinted with permission from ref. [48], Copyright © MDPI, 2018. Schematic representation for the synthesis by (b) hydrothermal, reprinted with permission from ref. [63], Copyright © Royal Society of Chemistry, 2023; (c) spin coating, reprinted with permission from ref. [64], Copyright © MDPI, 2021; (d) mechanism of the in situ generation of NiFeOx catalyst on BiVO4, reprinted with permission from ref. [65], Copyright © 2020, Wiley-VCH GmbH; (e) ion exchange technique, reprinted with permission from ref. [66], Copyright © 2016, American Chemical Society; (f) electrophoretic deposition of BiVO4, reprinted with permission from ref. [67], Copyright © 2021, Research square.
Figure 2. Representative morphologies for selected BiVO4 synthesis routes. (a) Automatic dip coating machine. Reprinted with permission from ref. [48], Copyright © MDPI, 2018. Schematic representation for the synthesis by (b) hydrothermal, reprinted with permission from ref. [63], Copyright © Royal Society of Chemistry, 2023; (c) spin coating, reprinted with permission from ref. [64], Copyright © MDPI, 2021; (d) mechanism of the in situ generation of NiFeOx catalyst on BiVO4, reprinted with permission from ref. [65], Copyright © 2020, Wiley-VCH GmbH; (e) ion exchange technique, reprinted with permission from ref. [66], Copyright © 2016, American Chemical Society; (f) electrophoretic deposition of BiVO4, reprinted with permission from ref. [67], Copyright © 2021, Research square.
Nanomaterials 15 01494 g002

3.3. Sol–Gel Method, Wet-Chemical Process, Hydrothermal, Solvothermal, SILAR

Cross-sectional features after BiOI→BiVO4 conversion are shown in Figure 2c; enlarged views and phase verification are compiled; solution-based synthesis routes—including hydrothermal, solvothermal, wet-chemical, successive ionic layer adsorption and reaction (SILAR), and sol–gel processes—are widely utilized for fabricating nanostructured BiVO4 [68,69,70,71,72]. Hydrothermal synthesis, conducted in an aqueous medium under elevated temperature and pressure, enables the growth of highly crystalline BiVO4 [68]. Solvothermal synthesis, analogous to hydrothermal synthesis but carried out in organic solvents, provides additional versatility in tailoring particle size and morphology [69]. Wet-chemical methods encompass various liquid-phase synthesis techniques, offering flexibility in adjusting material composition and properties [70]. The SILAR technique, a layer-by-layer deposition process, is effective for thin-film growth under ambient conditions, making it a cost-efficient route for large-scale fabrication [71]. The sol–gel method involves generating a sol from metal precursors in solution, which subsequently transforms into a gel and undergoes heat treatment to yield solid BiVO4, facilitating uniform nanostructure formation at low temperatures and scalability [72].

3.4. Chemical Vapor Deposition (CVD), Reactive Co-Sputtering, and Atomic Layer Deposition (ALD)

A dense, sol–gel-derived BiVO4 layer is depicted in Figure 2d; roughness tuning vs. annealing temperature is summarized; post-synthesis modification techniques—such as annealing, plasma treatment, and surface grafting—are commonly employed to enhance the structural, electronic, and interfacial properties of BiVO4 thin films [73,74,75,76]. Annealing at elevated temperatures improves crystallinity, reduces structural defects, and enhances charge transport [73]. Plasma treatment introduces reactive species to the film surface, enabling controlled modification of surface chemistry and improving catalytic activity, wettability, or interfacial charge transfer [74,75]. Surface grafting involves covalently attaching functional molecules to the BiVO4 surface, allowing tailored chemical functionalities that enhance PEC performance, stability, or compatibility with cocatalysts [76].

3.5. In Situ Coordination, Ion Exchange Technique, and Grafting

As shown in Figure 2e,f, Chemical modification strategies—including in situ coordination, ion exchange, and surface grafting—allow precise tailoring of BiVO4 composition and structure during or after synthesis [77,78,79,80,81,82]. In situ coordination facilitates controlled formation of coordination complexes during synthesis, influencing crystal structure, morphology, and optoelectronic properties [77]. Ion exchange techniques substitute selected ions within the BiVO4 lattice, enabling precise doping or compositional tuning to enhance optical absorption and catalytic activity [78]. Grafting, as a post-synthesis surface functionalization, allows covalent attachment of molecular species or functional groups to the BiVO4 surface, improving surface reactivity, stability, or compatibility with cocatalysts [79].
Electrochemical fabrication methods—including electrodeposition and photoelectrophoretic deposition (PEPD)—have also been successfully applied for BiVO4 photoanodes [80,81,82]. Electrodeposition uses an externally applied current to deposit BiVO4 from a precursor solution onto a conductive substrate, providing precise control over film thickness, morphology, and stoichiometry by tuning deposition parameters [80,81]. PEPD employs a photoinduced electric field to drive nanoparticle deposition onto the substrate, forming uniform, size-controlled nanoparticle-based BiVO4 films with homogeneous distribution [82].

4. Key Factors Influencing the PEC Performance of the BIVO4 Photoanode

The photocurrent density (JPEC) is a fundamental parameter for assessing the PEC performance of a photoelectrode. For a BiVO4 photoanode, JPEC can be calculated according to the following equation [31]:
J P E C = J m a x × L H E × η t r a n s × η s e p
A r r h e n i u s   e q u a t i o n : k = A e x p ( E A O P R T )
In Equation (6), Jmax represents the theoretical maximum photocurrent density attainable by BiVO4, which has been reported to be approximately 7.5 mA/cm2. According to this relationship, the PEC efficiency of BiVO4-based photoanodes can be significantly improved by enhancing the light-harvesting efficiency (LHE), charge separation efficiency ηsep, and charge transport efficiency ηtrans). Optimization of one or more of these factors—either individually or synergistically—is crucial for maximizing the overall performance of the photoanode in solar water splitting applications [83].
The rapid advancement of BiVO4-based photoanodes has been driven by various modification strategies, including elemental doping, post-treatment methods, incorporation of oxygen evolution catalysts (OECs), and integration of electron transport layers (ETLs) or hole transport layers (HTLs), in addition to nanostructuring techniques, as summarized in Table 2. The effectiveness of these strategies is multifaceted; for instance, ETLs not only reduce charge recombination at the interface between BiVO4 and ETL, as well as within the bulk of the photoelectrode, but they also significantly enhance surface charge transfer. This improvement arises from the influence of the underlying ETL on the deposition and structural characteristics of the overlying BiVO4 layer.
As shown in Figure 3, Park et al. developed a heterostructured photoanode by decorating anodic WO3 “nanocoral” arrays with ultrafine BiVO4 nanoparticles (~10 nm) via spin coating. This linked configuration established an efficient electron transfer pathway while preserving the nanocoral morphology. Under simulated solar illumination, the BiVO4/WO3 nanocoral electrode exhibited a photocurrent density of approximately 2.4 times that of bare WO3. More notably, incident photon-to-current conversion efficiency (IPCE) at 410 nm increased by a factor of 8.3, indicating substantially enhanced light absorption and charge carrier utilization. This study demonstrates that nanoscale heterojunction engineering combined with optimized morphology control can synergistically improve both optical harvesting and interfacial charge dynamics in PEC systems [106].
Figure 3. PEC performances of the WO3 nanocorals and BiVO4/WO3 (BW) photoanodes of 5-BW, 10-BW, and 30-BW: (a) photocurrent densities by linear sweep voltammetry (LSV), (b) Nyquist plots and equivalent circuits, (c) PEC H2 production diagram as a function of time, and (d) IPCE spectra. Reprinted with permission from Ref. [106], Copyright © 2024, Royal Society of Chemistry.
Figure 3. PEC performances of the WO3 nanocorals and BiVO4/WO3 (BW) photoanodes of 5-BW, 10-BW, and 30-BW: (a) photocurrent densities by linear sweep voltammetry (LSV), (b) Nyquist plots and equivalent circuits, (c) PEC H2 production diagram as a function of time, and (d) IPCE spectra. Reprinted with permission from Ref. [106], Copyright © 2024, Royal Society of Chemistry.
Nanomaterials 15 01494 g003

4.1. Introducing Extrinsic/Intrinsic Defects Through Doping

Doping plays a critical role in enhancing the performance of BiVO4 photoelectrodes by modulating their electronic and chemical properties. A key strategy for improving PEC efficiency involves either the introduction of intrinsic defects, such as oxygen vacancies, or the incorporation of foreign metal ions into the crystal lattice. Recent studies have provided valuable insights into different approaches for defect engineering, highlighting how such modifications can influence charge separation and transport processes.
Kong et al. demonstrated that Ni doping in BiVO4, achieved via electrodeposition, significantly enhances PEC performance through defect engineering. Ni-doped BiVO4 samples were categorized according to Ni content. The nanoparticle size decreased with increasing Ni content: 500 nm for pristine BiVO4, 440 nm for 3-Ni-BiVO4, 310 nm for 5-Ni-BiVO4, and 200 nm for 10-Ni-BiVO4. Controlled Ni incorporation promotes the formation of oxygen vacancies and surface V4+ species, which simultaneously improve charge separation and catalytic activity at the electrode–electrolyte interface. These synergistic effects effectively suppress recombination and accelerate charge transport. The morphology of Ni-doped BiVO4 are shown in Figure 4a [107]. As a result, the optimized 5-Ni-BiVO4 exhibited a photocurrent density of 2.39 mA/cm2 at 1.23 VRHE, approximately 2.5 times higher than pristine BiVO4 (0.94 mA/cm2), along with an IPCE of 45% (400–450 nm) and an ABPE of 0.55%, as shown is Figure 4b. In contrast, excessive Ni in 10-Ni-BiVO4 introduced additional recombination centers, highlighting the importance of finely tuning oxygen vacancy concentrations to achieve optimal PEC efficiency [107].
Figure 4. (a) SEM images of Ni-doped BiVO4. (b) JV curves of pure BiVO4 and Ni-doped BiVO4. Reprinted with permission from Ref. [107], Copyright © 2024, Springer Nature. (c) Schematic illustration of photoanode fabrication process (d) HR-TEM images of Ba:BiVO4/HfO2/NiPt. (e) JV curves of pure BiVO4 and Ni-doped BiVO4. Reprinted with permission from Ref. [108], Copyright © 2024, Wiley-VCH GmbH.
Figure 4. (a) SEM images of Ni-doped BiVO4. (b) JV curves of pure BiVO4 and Ni-doped BiVO4. Reprinted with permission from Ref. [107], Copyright © 2024, Springer Nature. (c) Schematic illustration of photoanode fabrication process (d) HR-TEM images of Ba:BiVO4/HfO2/NiPt. (e) JV curves of pure BiVO4 and Ni-doped BiVO4. Reprinted with permission from Ref. [108], Copyright © 2024, Wiley-VCH GmbH.
Nanomaterials 15 01494 g004
Shin et al. employed a synergistic combination of Ba doping, HfO2 passivation, and NiPt alloy cocatalyst decoration to enhance charge transport and interfacial kinetics. Ba doping into the BiVO4 lattice introduced beneficial defect states, improving carrier separation. A thin HfO2 layer, deposited via atomic layer deposition (ALD), served as a surface passivation layer, suppressing surface recombination and stabilizing the electrode, as shown in Figure 4c. High-resolution transmission electron microscopy (HR-TEM) confirmed successful dopant incorporation into the BiVO4 lattice and the formation of well-defined interfaces. Lattice fringes with an interplanar spacing of 0.31 nm corresponded to the (121) plane of highly crystallized BiVO4, whereas the top-side fringes with a spacing of 0.19 nm were attributed to the Ni (111) plane, indicating intimate contact between the cocatalyst and the photoanode surface. This multi-step surface modification strategy led to a substantial improvement in PEC performance, achieving a photocurrent density of 3.1 mA/cm2 at 1.23 VRHE under AM 1.5G illumination. These results highlight the effectiveness of combining electronic structure tuning with interfacial engineering to overcome key limitations of BiVO4, in particular, poor surface kinetics and charge recombination. The HR-TEM images and photocurrent curves validating the synergistic effect of Ba-doping, HfO2 passivation, and NiPt decoration are presented in Figure 4d,e [108].
In conclusion, these studies collectively demonstrate that defect engineering through doping, when coupled with meticulous interfacial engineering, constitutes a powerful approach to significantly enhance the PEC performance of BiVO4 photoanodes.

4.2. Heterojunction with ETL (Electron Transport Layer)

Despite its favorable optical bandgap and stability under neutral pH conditions, BiVO4 photoanodes suffer from poor charge separation and inefficient electron transport, which limit their practical PEC performance. Traditional strategies, such as embedding WO3 as a buried electron transport layer, can enhance electron extraction but necessitate thick absorbers, conflicting with the inherently low charge mobility in BiVO4.
Yu et al. addressed these limitations by applying an ultrathin Co-phthalocyanine (CoPc) surface layer to the BiVO4 photoanode, effectively tuning its surface hydrophilicity. XRD confirmed the formation of monoclinic BiVO4 on FTO without observable CoPc peaks, likely due to its low loading and high dispersion, while Raman spectra revealed characteristic C–C and C–N vibrational bands of CoPc, verifying its successful incorporation and stronger interfacial coupling in solvothermally treated samples. The successful incorporation of CoPc and the improved interfacial coupling are supported by the XRD and Raman results shown in Figure 5a,b, where characteristic peaks and morphological changes are clearly observed. This interface modification improved interaction with the electrolyte and suppressed surface recombination, resulting in a photocurrent density of 4.0 mA/cm2 at 1.23 VRHE—approximately 3.1 times higher than that of pristine BiVO4—along with enhanced IPCE and operational stability. The ultrathin interface layer overcomes the drawbacks of thick buried ETLs, providing an elegant solution for performance improvement without compromising charge mobility [109].
Figure 5. (a) XRD patterns; (b) Raman spectra and SEM images of the BVO@CoPc-S photoanodes. Reprinted with permission from Ref. [109], Copyright © 2025, Royal Society of Chemistry; (c) SEM image of Mo:BiVO4, (d) XRD of Mo:BiVO4 and CPF-TCB/Mo:BiVO4; (e) schematic illustration of the electropolymerization of CPF-TCB on Mo:BiVO4. Reprinted with permission from Ref. [110], Copyright © 2024, Royal Society of Chemistry.
Figure 5. (a) XRD patterns; (b) Raman spectra and SEM images of the BVO@CoPc-S photoanodes. Reprinted with permission from Ref. [109], Copyright © 2025, Royal Society of Chemistry; (c) SEM image of Mo:BiVO4, (d) XRD of Mo:BiVO4 and CPF-TCB/Mo:BiVO4; (e) schematic illustration of the electropolymerization of CPF-TCB on Mo:BiVO4. Reprinted with permission from Ref. [110], Copyright © 2024, Royal Society of Chemistry.
Nanomaterials 15 01494 g005
Furthermore, Yang et al. introduced an organic–inorganic heterojunction by conformally coating a conjugated polycarbazole framework (CPF-TCB) onto nanoporous Mo-doped BiVO4. The morphology, XRD patterns, and schematic representation of CPF-TCB deposition on Mo:BiVO4 are presented in Figure 5c–e, highlighting the conformal coverage and structural features of the hybrid interface. This hole transport layer (HTL) forms a type-II band alignment with BiVO4, facilitating efficient hole extraction and suppressing recombination. After loading a NiFeCoOx cocatalyst, the photoanode achieved a record-high water oxidation photocurrent density of 6.66 mA/cm2 at 1.23 VRHE. When integrated into unassisted tandem PEC devices, it delivered solar-to-hydrogen conversion efficiencies of up to 9.02%, setting benchmarks for organic–inorganic hybrid photoelectrodes under operational conditions [110].
Overall, the integration of an external ETL via the carbon shell provides a powerful and scalable strategy to overcome the intrinsic charge transport limitations of oxide semiconductors, offering promising avenues for the development of high efficiency and durable PEC energy conversion systems.

4.3. Hole-Transport Layers (HTLs) for BiVO4 Photoanodes

Beyond electron transport layers, hole transport layers (HTLs) play a critical role in enhancing the surface charge extraction of BiVO4. By creating favorable band offsets for hole transfer, passivating surface trap states that promote interfacial recombination, and offering abundant well-dispersed oxygen evolution sites when coupled with cocatalysts, HTLs directly improve the charge transfer efficiency (ηtrans) under water oxidation conditions [111,112,113,114].
Cui et al. demonstrated that inserting a partially oxidized 2D bismuthene layer between BiVO4 and NiFeOOH enhanced interfacial band bending, passivated VO-related traps, and extended hole lifetime. The interfacial hole transport mechanism and performance enhancement enabled by such interlayers are schematically illustrated in Figure 6a,b. This strategy yielded a 5.8-fold increase in photocurrent compared to bare BiVO4, achieving 3.4 ± 0.2 mA/cm2 at +0.8 VRHE and 4.7 ± 0.2 mA/cm2 at +1.23 VRHE, with stable operation under illumination [111].
Figure 6. (a) Schematic illustration of interfacial hole transport in bare BiVO4 photoanode and (b) interfacial hole transport pathway in NiFeV/B-BiVO4 photoanode. Reprinted with permission from Ref. [113], Copyright © 2021, Royal Society of Chemistry. (c) Photocurrent density stability of photoanodes at 0.8 VRHE. Reprinted with permission from Ref. [111], Copyright © 2022, Wiley-VCH GmbH. (d) LSV curves of BiVO4, B-BiVO4, NiFeV/BiVO4, and NiFeV/B-BiVO4 photoanodes under AM 1.5G illumination in 1.0 M potassium borate buffer (pH 9.3, scan rate: 10 mV/s). Reprinted with permission from Ref. [113], Copyright © 2021, Royal Society of Chemistry.
Figure 6. (a) Schematic illustration of interfacial hole transport in bare BiVO4 photoanode and (b) interfacial hole transport pathway in NiFeV/B-BiVO4 photoanode. Reprinted with permission from Ref. [113], Copyright © 2021, Royal Society of Chemistry. (c) Photocurrent density stability of photoanodes at 0.8 VRHE. Reprinted with permission from Ref. [111], Copyright © 2022, Wiley-VCH GmbH. (d) LSV curves of BiVO4, B-BiVO4, NiFeV/BiVO4, and NiFeV/B-BiVO4 photoanodes under AM 1.5G illumination in 1.0 M potassium borate buffer (pH 9.3, scan rate: 10 mV/s). Reprinted with permission from Ref. [113], Copyright © 2021, Royal Society of Chemistry.
Nanomaterials 15 01494 g006
In another approach, Li et al. employed a serial HTL architecture by sequentially depositing Fe2O3 (facilitating bulk-to-surface hole transfer) and NiOOH/FeOOH (promoting surface-to-electrolyte transfer) on BiVO4. The photocurrent stability and LSV curves of these stepwise HTLs are shown in Figure 6c,d, confirming the substantial improvements in charge transfer efficiency. The resulting photoanode delivered 2.24 mA/cm2 at 1.23 VRHE (≈2.95× over pristine BiVO4). At the same potential, ηbulk and ηsurface were enhanced by 1.63× and 2.62×, respectively, demonstrating stepwise improvements in charge transfer. When coupled with a commercial Si photovoltaic for self-biasing, the device sustained 2.60 mA/cm2, corresponding to a solar-to-hydrogen (STH) efficiency of 3.2%, underscoring the practicality of sequential HTLs [112].
Expanding this concept, Meng et al. reported that NiFeV-LDH deposited on borate-treated BiVO4 (B-BiVO4) served as an interfacial hole transport/storage layer. The optimized NiFeV/B-BiVO4 photoanode achieved 4.6 mA/cm2 at 1.23 VRHE with an early onset potential of 0.2 VRHE, while retaining ≥80% of the initial photocurrent after 24 h and maintaining ~95% faradaic efficiency. Mechanistic studies revealed a substantial increase in Z_surface (~90% vs. 31% for bare BiVO4) and reduced charge transfer resistance (Rct), confirming accelerated interfacial hole transfer upon LDH integration [113].
Collectively, these studies highlight that rational HTL design—from 2D bismuthene insertion to serial inorganic HTLs and LDH-based transport/storage layers—offers complementary pathways to lower interfacial barriers, enhance band bending, and provide robust scaffolds for cocatalyst integration, thereby substantially boosting both activity and stability of BiVO4 photoanodes.

4.4. Cocatalyst and Surface Modification Strategies for BiVO4 Photoanodes

A wide range of cocatalyst deposition and surface modification strategies have been developed to improve the PEC performance of BiVO4 photoanodes. These approaches primarily aim to accelerate OER kinetics, passivate surface trap states, and suppress interfacial charge recombination, thereby enhancing both photocurrent density and operational stability. Table 3 summarizes representative strategies such as dual-layer deposition, ligand-assisted passivation, and facet-selective cocatalyst loading, highlighting their impact on photocurrent density, onset potential, Faradaic efficiency, and durability.
Among these strategies, Dong et al. reported a dual modification approach of coupling CdS nanoparticles with NiFe-LDH nanosheets on BiVO4. XRD confirmed the monoclinic scheelite phase of BiVO4, while the absence of distinct CdS and NiFe-LDH peaks was attributed to the low CdS loading and amorphous nature of NiFe-LDH. XPS verified the coexistence of CdS and NiFe-LDH through Ni2+ and Fe3+ peaks, and SEM/HRTEM revealed a defect-rich amorphous nanosheet structure providing abundant active sites. The CdS nanoparticles broadened the light absorption range and facilitated charge transfer, while the NiFe-LDH overlayer functioned as a cocatalyst to accelerate OER kinetics and passivate surface traps. This synergistic interface engineering significantly improved charge separation and transport properties, resulting in a photocurrent density of 3.1 mA cm−2 at 1.23 V_RHE—about 5.8 times higher than bare BiVO4—and excellent stability under continuous illumination. The enhanced photocurrent density, crystallographic features, and XPS confirmation of the cocatalyst loading are shown in Figure 7a–d [49].
Figure 7. (a) LSV curves and (b) XRD patterns of NiFe-LDH/CdS/BiVO4, NiFe-LDH/BiVO4, CdS/BiVO4, and pristine BiVO4 photoanodes measured in 0.5 M Na2SO4 electrolyte (pH 6.1) at 1.23 VRHE. (c,d) XPS spectra of Ni 2p and Fe 2p, respectively. Reprinted with permission from Ref. [49], Copyright © 2024, MDPI. (e) LSV and OER polarization curves of the photoanodes measured without illumination Reprinted with permission from Ref. [124], Copyright © 2024, Royal Society of Chemistry. (f) LSV curves recorded under intermittent AM 1.5 G illumination in 0.5 M borate buffer (pH 9.0) with a cathodic scan rate of 10 mV/s. Reprinted with permission from Ref. [125], Copyright © 2024, Royal Society of Chemistry.
Figure 7. (a) LSV curves and (b) XRD patterns of NiFe-LDH/CdS/BiVO4, NiFe-LDH/BiVO4, CdS/BiVO4, and pristine BiVO4 photoanodes measured in 0.5 M Na2SO4 electrolyte (pH 6.1) at 1.23 VRHE. (c,d) XPS spectra of Ni 2p and Fe 2p, respectively. Reprinted with permission from Ref. [49], Copyright © 2024, MDPI. (e) LSV and OER polarization curves of the photoanodes measured without illumination Reprinted with permission from Ref. [124], Copyright © 2024, Royal Society of Chemistry. (f) LSV curves recorded under intermittent AM 1.5 G illumination in 0.5 M borate buffer (pH 9.0) with a cathodic scan rate of 10 mV/s. Reprinted with permission from Ref. [125], Copyright © 2024, Royal Society of Chemistry.
Nanomaterials 15 01494 g007
Beyond inorganic systems, organic framework materials such as MOFs and Covalent organic frameworks (COFs) offer a molecularly tunable platform with intrinsically high surface area and porosity, enabling precise control over catalytic site density and efficient mass transport. Guo et al. reported the in situ growth of COF–Azo and COF–Ben on BiVO4, yielding hybrids with reduced charge transfer resistance, lower onset potential, and higher photocurrent densities compared to pristine BiVO4. The improved OER activity of the COF–Azo modified electrode is evidenced by the polarization and LSV curves in Figure 7e. EIS confirmed enhanced interfacial kinetics, and the intimate COF–BiVO4 heterojunction facilitated efficient hole extraction and suppressed recombination, leading to markedly improved OER activity [124]. Similarly, Feng et al. demonstrated that combining Mo doping with cobalt polyoxometalate (CoPOM) deposition effectively enhanced bulk charge transport, lowered OER overpotential, and delivered significantly higher photocurrent densities with improved operational stability. The photocurrent enhancement and stability of the CoPOM-modified BiVO4 are illustrated in Figure 7f [125]. Collectively, these results demonstrate that both inorganic cocatalysts/overlayers and MOF/COF-based organic frameworks provide complementary strategies to overcome sluggish surface kinetics and interfacial recombination in BiVO4 photoanodes, ultimately enabling high efficiency and durable PEC water splitting devices.

4.5. Light Absorption Efficiency

BiVO4 possesses a bandgap of approximately 2.4 eV, enabling the absorption of photons with wavelengths ranging from 300 to 520 nm [126]. Under AM 1.5G solar illumination, the theoretical maximum photocurrent density (Jmax) is 7.5 mA/cm2 [20]. In practice, however, optical losses due to light refraction and reflection reduce effective light absorption, yielding an actual theoretical photocurrent density of only 4.45 mA/cm2 [127]. This discrepancy highlights that light absorption efficiency is a critical factor limiting the PEC performance of BiVO4-based photoanodes.
Enhancing light absorption is, therefore, essential for improving PEC activity. Effective strategies include constructing heterostructures and integrating optical elements into the BiVO4 architecture, both of which can strengthen photon–matter interactions [16,128]. These approaches increase photon utilization and, consequently, improve the overall performance of the photoanode.

4.5.1. Band Edge Engineering

Reducing the bandgap of semiconductor photocatalysts is an established strategy to broaden their light absorption spectrum, potentially extending it into the near-infrared region. Among various approaches, defect engineering via doping is particularly effective for BiVO4, as it introduces defect levels within the band structure, modifies the electronic configuration, shifts the band edges, and enhances light absorption. Careful control of dopants and defect states allows the bandgap to be narrowed without compromising charge transport, resulting in improved PEC performance.
Fan Feng et al. systematically investigated the synergistic effects of Mo doping and surface modification with a cobalt polyoxometalate (CoPOM) on BiVO4 photoanodes. UV–Vis absorption spectra showed similar absorption onsets of around 515 nm for all samples, and Tauc plot analysis indicated bandgap energies of ~2.56 eV, confirming that Mo doping and CoPOM deposition do not alter the fundamental absorption edge or introduce significant parasitic light absorption. This suggests that the observed performance enhancement originates primarily from improved charge transport and interfacial catalytic activity rather than changes in light-harvesting properties. Photocurrent–potential curves showed that Mo doping alone increased the photocurrent from ~0.65 to ~2.53 mA/cm2 at 1.23 VRHE, while the combination of Mo–BiVO4/CoPOM further enhanced it to ~4.32 mA/cm2. The absorption spectra, Tauc plots, and density of states analysis supporting these results are shown in Figure 8a–c. As a result, this represents a ~6.6-fold improvement over pristine BiVO4 and ~1.7-fold compared to Mo-doped BiVO4 alone. The enhancement is attributed to improved conductivity through Mo doping and accelerated water oxidation kinetics via CoPOM catalysis. Electrochemical impedance spectroscopy (EIS) confirmed a significant reduction in charge transfer resistance, while the optical appearance of the electrodes validated enhanced visible light response [125].
Similarly, Elnagar et al. demonstrated that Mo-doped BiVO4 modified with a CoPOM surface layer exhibited a slight bandgap narrowing and extended light absorption into the longer-wavelength visible region [129]. The UV–Vis absorption spectra and Tauc plots confirming this bandgap narrowing are presented in Figure 8d,e. This, combined with surface modification, resulted in synergistic improvement in PEC performance.
Furthermore, Zhou et al. reported that co-doping BiVO4 with Mo and Ni effectively narrows the bandgap and enhances light-harvesting capability. UV–Vis spectra showed a red-shift in the absorption edge, corresponding to a bandgap reduction of approximately 0.15–0.20 eV compared to pristine BiVO4. The co-doped photoanode exhibited a photocurrent density of ~3.6 mA/cm2 at 1.23 VRHE, over 40% higher than undoped BiVO4. EIS analysis confirmed improved charge transfer resistance, and the color change in the electrodes reflected the enhanced visible light response. These results indicate that band-edge tuning via Mo–Ni co-doping effectively extends light absorption and boosts PEC activity [130].
These findings collectively confirm that band-edge engineering through doping and defect creation remains a promising and actively evolving strategy to enhance the PEC performance of BiVO4 photoanodes in solar water splitting.

4.5.2. Constructing Heterojunction Structure

Constructing heterojunctions is an effective strategy to enhance the light-harvesting efficiency (LHE) of BiVO4, suppress charge recombination, and accelerate OER kinetics. By coupling BiVO4 with suitable semiconductors or cocatalysts, researchers have developed advanced heterostructures that significantly improve PEC water splitting performance.
Seo et al. demonstrated the effectiveness of a p–n heterojunction between BiVO4 and PbS quantum dots (QDs), combined with a ZnS overlayer, for enhanced PEC hydrogen production [54]. Nanoporous BiVO4 films were prepared via electrodeposition, followed by PbS QD sensitization using the successive ionic layer adsorption and reaction (SILAR) method. The PbS QDs extended light absorption into the near-infrared region and improved charge separation through the p–n junction interface. Subsequently, a conformal ZnS overlayer was deposited to passivate surface defects on the QDs, reducing non-radiative recombination.
As a result, the photocurrent density at 1.23 VRHE increased from 2.92 mA/cm2 for bare BiVO4 to 4.88 mA/cm2 after PbS QD sensitization, and further to 5.19 mA/cm2 with the ZnS overlayer. The JV and stability curves of these heterostructured electrodes are shown in Figure 9a,b, confirming the improved photocurrent and durability. EIS confirmed a substantial reduction in charge transfer resistance (Rct), indicating enhanced surface kinetics. Stability tests over 2 h revealed retention rates of 91.24% for BiVO4/PbS and 96.20% for BiVO4/PbS/ZnS, emphasizing the role of passivation layers in maintaining operational durability. This study highlights that combining heterojunction engineering with surface passivation can simultaneously maximize efficiency and stability in BiVO4-based PEC systems.
Similarly, Zhang et al. reported highly efficient BiVO4 single-crystal nanosheets modified via phosphorus doping and selective Ag decoration. The synthetic route and structural modification strategy are schematically illustrated in Figure 9c. This dual modification generated oxygen vacancies and improved charge migration pathways, leading to a ~2.8-fold increase in methylene blue degradation compared to pristine BiVO4 under visible light irradiation [131].
Li et al. fabricated nanostructured BiVO4/Co-Pi heterostructures using a vanadium-infused interaction method. The incorporation of the Co-Pi cocatalyst facilitated efficient hole extraction and prolonged carrier lifetime, enabling a photocurrent density of 2.9 mA/cm2 at 1.23 VRHE under simulated solar illumination [47]. These results underscore the effectiveness of heterojunction construction and cocatalyst integration for enhancing the PEC performance of BiVO4 photoanodes.
In another example, Gil-Rostra et al. developed ITO/WO3/BiVO4/CoPi multishell nanotube arrays using a soft-templating technique [132]. The intimate WO3/BiVO4 interface generated an internal electric field that enhanced charge separation, while the CoPi overlayer improved surface OER kinetics. In Figure 9d, LSV curves confirmed a substantial increase in photocurrent upon CoPi deposition, and the band diagram illustrates the overall charge transport pathway: photoholes generated in BiVO4 migrate to the surface to drive OER, whereas photoexcited electrons are efficiently funneled through the WO3 shell and ITO layer to the external circuit. This heterostructured architecture enables efficient light absorption within BiVO4, while the nanoscale tubular design shortens the electron diffusion path, thereby minimizing ohmic losses and maximizing charge collection. The corresponding band diagram describing the fate of photogenerated carriers is presented in Figure 9e. The resulting hierarchical electrode exhibited outstanding stability and PEC performance under one-sun illumination.
These studies collectively demonstrate that combining heterojunction formation with surface modification strategies can simultaneously maximize optical absorption, charge separation, and operational stability of BiVO4-based photoanodes, paving the way toward efficient solar fuel generation.

4.5.3. Optics-Based Elements

Tuning the morphological features of photocatalysts provides an effective strategy to modulate light scattering and reflection, thereby enhancing overall light absorption and improving catalytic performance [133]. Three-dimensional (3D) inverse opal architectures, in particular, have emerged as a powerful design due to their highly ordered macro–mesoporous networks and substantially increased specific surface areas [134].
A recent study reported the fabrication of a 3D core–shell inverse opal photoanode composed of FTO/TiO2/BiVO4, using a combination of atomic layer deposition and electrodeposition techniques [135]. This configuration creates a uniform opal framework that facilitates rapid electron transport and improved light management, enabling the photoanode to achieve a photocurrent density of 4.11 mA/cm2 at 1.23 VRHE, significantly outperforming comparable planar structures.
For example, Marinković et al. synthesized zircon-type BiVO4 nanoparticles (2–8 nm) via a facile colloidal route. The nanosized BiVO4 exhibited a tetragonal zircon structure with enhanced optical absorption and a significantly enlarged surface-to-volume ratio. These features led to superior photocatalytic degradation of methyl orange compared with commercial TiO2, underscoring the potential of ultrafine BiVO4 nanoparticles for efficient light harvesting and environmental remediation The photoluminescence spectra and photodegradation performance confirming these results are presented in Figure 10a,b [136].
Similarly, Zhao et al. demonstrated that controlling the crystal orientation of BiVO4 thin films has a critical impact on PEC behavior. BiVO4 films preferentially oriented along the [010] direction exhibited markedly improved carrier separation and mobility compared to [119]-oriented or randomly oriented samples. The synthesis process and characterization of these catalytic films are schematically illustrated in Figure 10c. As a result, the [010]-oriented BiVO4 delivered a photocurrent density of 0.2 mA/cm2 at 1.21 VRHE, outperforming [119]-oriented (0.056 mA/cm2) and random films (0.11 mA/cm2) [137]. Despite these significant improvements in light absorption, optimizing light utilization remains a key challenge in advancing the performance of BiVO4 photoanodes.

4.6. Photogenerated Charge Separation Efficiency (ηsep)

The primary bottleneck in the PEC efficiency of BiVO4 photoanodes is the rapid recombination of photogenerated electron–hole pairs. This limitation is further exacerbated by the low carrier mobility of BiVO4, which is approximately 10−2 cm2/V·s, resulting in substantial charge recombination before the charges can reach the electrode surface [138]. Enhancing charge migration is, therefore, essential to unlock the full PEC potential of BiVO4.
To mitigate recombination, extensive strategies have been developed to improve carrier transport and separation. These include elemental doping to modulate carrier density and defect states, construction of heterojunctions to establish favorable band alignments, and interface engineering with cocatalysts to accelerate surface charge transfer. Additionally, defect engineering, such as introducing oxygen vacancies, can further promote charge separation by creating internal electric fields or shallow trap states that reduce recombination.
Representative examples are summarized in Table 4, which highlights the PEC performance and ηsep of selected BiVO4-based photoanodes. For instance, the introduction of a WO3/S:Bi2O3/(Ga,W):BiVO4/Co–Pi interface structure significantly improved photocurrent density through optimized charge transfer pathways [135], while oxygen vacancy-engineered BiVO4 (Ov-BiVO4) achieved an outstanding ηsep of 94% with a photocurrent density of 6.29 mA/cm2 [139]. These results collectively demonstrate that tailored modifications in BiVO4 can directly translate to higher charge separation efficiency and overall PEC activity.
Taken together, these findings underscore that ηsep serves as a crucial performance metric in evaluating BiVO4-based photoanodes. Continued progress in tuning bulk properties, interface chemistry, and defect states has already shown remarkable improvements, but further breakthroughs will likely depend on synergistically integrating multiple strategies.
Therefore, improving the charge separation efficiency of BiVO4 requires a comprehensive consideration of multiple strategies, including the utilization of internal electric fields and the modulation of conductivity and carrier density. The following section will examine these representative approaches in detail, with a focus on their underlying mechanisms and demonstrated performance.

4.6.1. Utilizing in Internal Electric Field

A practical approach to overcome the intrinsically low charge separation efficiency (ηsep) of BiVO4 is the construction of heterojunctions that induce a built-in internal electric field (IEF). This IEF bends the energy bands at the interface, directing electrons and holes in opposite directions and thereby extending carrier lifetimes. As a result, the overall PEC water splitting performance is significantly enhanced.
For example, Bhat et al. constructed a triple planar SnO2/WO3/BiVO4 heterojunction. The staggered (type-II) band alignment, along with the SnO2 hole-blocking layer, facilitated interfacial charge separation and electron transport, yielding a photocurrent density of ~2.01 mA/cm2 at 1.23 VRHE under front illumination, with remarkably high IPCE values of ~90% (front) and ~80% (back) at 400 nm. The photocurrent response and impedance characteristics of this triple planar heterojunction are presented in Figure 11a,b. The authors attributed this enhancement to efficient interfacial separation and reduced recombination under front illumination, consistent with the effect of a built-in field across the multilayer junction [147].
Similarly, Liang et al. fabricated surface-dispersed WO3/BiVO4 heterojunction arrays (BiVO4-nanoparticle@WO3-nanoflake). Decorating WO3 nanoflakes uniformly with ~20–50 nm BiVO4 nanoparticles increased the effective junction area and mitigated interfacial hole accumulation, thereby enhancing charge separation. The schematic formation process and corresponding LSV curves of these heterojunction arrays are shown in Figure 11c,d. The resulting composite achieved a photocurrent density of 3.53 mA/cm2 for PEC water oxidation—approximately twice that of WO3 alone—while maintaining good stability, indicating a robust IEF over the nanojunction network [148].
In another optics-plus interface example, Cao et al. engineered an FTO/WO3/BiVO4 heterojunction and applied a mild NaOH post-treatment to modify the BiVO4 surface chemistry. The heterojunction provided an interfacial driving force for carrier separation, while the surface treatment accelerated water oxidation kinetics, together yielding ~1.75 mA/cm2 at 1.23 VRHE—more than double that of WO3,as shown in Figure 11e [149].
Collectively, these studies demonstrate that establishing heterointerfaces (SnO2/WO3/BiVO4; WO3/BiVO4) and tuning surface chemistry can create or strengthen internal electric fields, which (i) promote directional carrier migration, (ii) suppress interfacial recombination, and (iii) enhance photocurrent and IPCE without the need for precious metal cocatalysts.

4.6.2. Enhancing Electrical Conductivity and Carrier Concentration

Enhancing the electrical conductivity of BiVO4 has emerged as an important strategy to improve its intrinsic material properties [150]. Doping, in particular, is widely recognized as an effective means of tuning the electronic structure and thereby increasing conductivity [151]. Both metal and nonmetal dopants have been successfully incorporated into BiVO4 for this purpose [152]. Among them, tungsten (W) and molybdenum (Mo) are especially effective: substituting V5+ ions with W6+ or Mo6+ ions significantly enhance the electronic conductivity of BiVO4 [153].
Enhancing the electrical conductivity of BiVO4 is a critical strategy to facilitate efficient charge transport and suppress recombination losses during PEC water splitting. Improved conductivity not only enables faster carrier movement but also contributes directly to higher photocurrent generation. Wu et al. demonstrated that Mo doping in BiVO4 significantly lowers the interfacial charge transfer resistance by 2–3 orders of magnitude under illumination. This finding indicates that interfacial kinetics, rather than bulk resistance, dominate the PEC performance of pristine BiVO4. As a result, the enhanced conductivity and accelerated charge transfer led to a remarkable increase in photocurrent density, reaching values ~2.7 times higher than those of undoped BiVO4 at 1.23 VRHE. Figure 12a–c present the LSV curves and Mott–Schottky plots, confirming the enhanced donor density and reduced resistance [154].
Similarly, Chaudhari et al. reported that introducing a conductive carbon network into BiVO4 through MOF-derived Bi–V–O/carbon composites (BVC) effectively reduced charge transport resistance and enlarged the electrochemically active surface area, leading to enhanced PEC water oxidation activity and additional applicability in energy storage systems. The corresponding LSV, stability, and EIS results supporting these improvements are shown in Figure 12d–f [155].
Taken together, these results underscore that strategies improving the electrical conductivity of BiVO4—whether through cation doping or conductive composite engineering—consistently enhance PEC performance by lowering resistance, increasing carrier density, and enabling more efficient charge extraction.

4.7. Charge Transfer Efficiency (ηtrans)

Enhancing the charge transfer efficiency ηtrans at the photoanode/electrolyte interface is crucial for realizing efficient PEC performance in BiVO4. Although photogenerated holes in BiVO4 can reach the electrode surface to drive the OER, a substantial fraction is often lost through recombination with electrons during transport or at the interface. To assess this limitation, ηtrans is generally determined by comparing the photocurrent densities measured in the presence and absence of a hole scavenger such as Na2SO3, which suppresses surface recombination and reflects the ideal case of complete hole utilization. The efficiency can be expressed as
η t r a n s f e r % = J H 2 O J N a 2 S O 3 × 100
where JH2O represents the photocurrent density obtained in water oxidation, and JNa2SO3 corresponds to the photocurrent density measured with Na2SO3 present.
This evaluation emphasizes that the sluggish surface reaction kinetics of BiVO4 remain a major bottleneck for water oxidation. Accordingly, considerable effort has been devoted to loading suitable cocatalysts onto BiVO4, which not only suppress surface recombination but also lower the reaction overpotential. These strategies have been consistently reported to enhance ηtrans, resulting in improved photocurrent densities and overall PEC water splitting efficiency.

Co-Catalysts Based on Metal (Oxy) Hydroxides

Incorporating metal (oxy)hydroxide (M-OOH) cocatalysts, such as FeOOH and NiFeOOH, onto BiVO4-based photoanodes has emerged as one of the most effective approaches to enhance charge transfer efficiency (ηtrans). These cocatalysts supply abundant active sites, suppress surface recombination, and accelerate water oxidation kinetics [156].
For instance, Li et al. fabricated a micro–nanostructured FeOOH/BiVO4/WO3 photoanode via a combination of hydrothermal, electrodeposition, and impregnation methods. The resulting heterostructure exhibited a photocurrent density of ~2.04 mA/cm2 at 1.23 VRHE, nearly doubling that of the FeOOH-free counterpart (~1.09 mA/cm2), along with a positive shift in onset potential from 0.80 to 0.60 VRHE. The SEM morphology, surface photovoltage, and dark current characteristics supporting these improvements are shown in Figure 13a–c [157].
Similarly, Creasey et al. constructed a WO3/BiVO4/NiFeOOH photoanode using a scalable aerosol-assisted CVD process. The incorporation of NiFeOOH not only stabilized the heterojunction but also suppressed BiVO4 dissolution, enabling a stable photocurrent of 1.75 mA/cm2 at 1.23 VRHE that was sustained for 24 h under one-sun illumination, as shown in Figure 12d,e [158].
Another study by Li et al. reported BiVO4/CoPi electrodes prepared through in situ electrodeposition. The CoPi layer increased photocurrent by ~1.8× relative to bare BiVO4, achieving 1.39 mA/cm2 at 1.23 VRHE, while simultaneously enhancing applied-bias photon-to-current efficiency (ABPE) and reducing onset potential—highlighting its role in effective passivation and charge separation. The JV curves and ABPE values of the BiVO4/CoPi photoanodes are illustrated in Figure 13f–h [159].
Collectively, these findings confirm that M-OOH cocatalyst coatings on BiVO4 or BiVO4-based heterostructures significantly promote charge transfer, reduce onset potential, enhance stability, and thereby improve overall PEC performance.

4.8. Stability

The operational stability of BiVO4 photoanodes is a key factor for their practical application, as maintaining a consistent photocurrent over extended periods is essential. Photocorrosion, caused by undesirable redox reactions between photogenerated electrons and holes within the BiVO4 lattice, can significantly reduce charge extraction efficiency. This degradation is further exacerbated by light-induced leaching of V5+ ions, which distorts the lattice and causes long-term structural damage. Recent studies have highlighted two primary strategies to mitigate these effects: the incorporation of protective surface layers and the optimization of electrolyte composition, both of which are discussed in the following sections.

4.8.1. Protection Layer Incorporation

Photocorrosion in BiVO4 photoanodes primarily arises from photogenerated holes reacting with surface atoms or via light-assisted dissolution processes, resulting in structural degradation and reduced stability. A practical approach to mitigate these effects is the application of a protective layer, which physically blocks corrosive interactions while still allowing hole transfer to support OER.
For example, Zhang et al. systematically examined the operando photostability of BiVO4 in near-neutral electrolytes using a scanning flow cell, directly tracking dissolution during light-driven OER. They demonstrated that the dissolution rate is strongly dependent on the electrolyte composition (borate > phosphate > citrate), and that citrate provides kinetic protection via hole scavenging. Distinct dissolution potentials for Bi and V were identified, and time-resolved measurements revealed that photocurrent and dissolution can evolve independently, highlighting that surface chemistry control is essential for achieving long-term stability. These findings underscore the need for barrier or passivation layers to decouple corrosion from charge extraction on BiVO4 surfaces. Figure 14a–f provides direct visual confirmation of these dynamics, coupling photocurrent/thickness profiles with dissolution rate measurements and post-mortem morphology analyses under borate and phosphate conditions [160].
Similarly, Wang et al. reviewed interface regulation strategies for BiVO4, emphasizing artificial protection layers—such as ultrathin ALD oxides (TiO2, Al2O3) and organic/organic–inorganic interlayers—that suppress photocorrosion while maintaining hole transfer. The review highlighted the design principles for these coatings, which include (i) blocking electrolyte attack, (ii) passivating surface traps, and (iii) forming favorable band alignment for OER catalysis, thereby lowering interfacial resistance, shifting onset potentials, and improving operational durability under AM 1.5G illumination. Representative protection strategies and schematic illustrations of stabilized BiVO4 electrodes are shown in Figure 14g–i. [161].
Collectively, these examples demonstrate that protective layers—whether polymeric or ALD-oxide—are central to stabilizing BiVO4 photoanodes, mitigating operando dissolution, and sustaining interfacial hole transport necessary for efficient oxygen evolution, ultimately enabling durable and high-performance PEC devices.

4.8.2. Dual-Protection Strategies for Durable BiVO4 Photoanodes

Photocorrosion and dissolution of BiVO4—especially the leaching of V5+ ions—represent critical challenges to achieving long-term PEC stability. A practical approach to mitigate these effects involves tuning the electrolyte composition to suppress dissolution and stabilize the surface chemistry.
Lei et al. demonstrated that conformal NiOx overlayers deposited via ALD, combined with the deliberate addition of NaVO3 to the electrolyte, provided a synergistic stabilization effect. The NiOx coating suppressed surface charge recombination and facilitated hole transfer by stabilizing Bi–O bonds, while dissolved vanadate species helped re-establish chemical equilibrium and compensate for V loss. This dual-protection strategy enabled NiOx/BiVO4 photoanodes to achieve an applied-bias photon-to-current efficiency (ABPE) of 2.05% with a fill factor of 47.1%, and, remarkably, operational durability exceeding 2100 h under continuous PEC operation. The long-term stability and proposed anticorrosion mechanism of the NiOx-protected BiVO4 electrodes are illustrated in Figure 15a,b [162].
More recently, Lee et al. introduced a complementary approach by selectively engineering surface oxygen vacancies (VO) via a surface chemical reduction (SCR) process, followed by deposition of an ultrathin TiO2 layer (≈5 nm) via ALD. The VO-rich surface promoted strong oxygen-end networking with TiO2, yielding a nearly pinhole-free protective layer that minimized interfacial recombination while maintaining efficient charge tunneling. When further coupled with a cobalt phosphate (CoPi) oxygen evolution catalyst, the optimized SCR-BiVO4/TiO2/CoPi photoanode exhibited a stable photocurrent density of 3.9 mA/cm2 at 1.23 VRHE, with charge transfer efficiency of up to 97% and sustained photostability for over 50 h while generating stoichiometric H2 and O2. The PEC performances in different electrolytes and the schematic of charge kinetics in this optimized multilayer system are shown in Figure 15c–e. [163].
These findings confirm that electrolyte composition is a critical factor in BiVO4 stability. By adding protective species (e.g., phosphate buffers, vanadium ions) or coupling with passivation layers (e.g., NiOx), the dissolution rate can be significantly suppressed while preserving efficient hole transfer, making electrolyte engineering a powerful tool for enabling durable PEC water splitting devices.

4.8.3. Synergistic Surface–Electrolyte Protection Strategies

The identity of the electrolyte—including buffer anions/cations and pH—has been reported to govern not only interfacial oxygen evolution kinetics but also the chemical pathways of photocorrosion in BiVO4. In phosphate electrolytes, competitive reactions between water oxidation and lattice dissolution proceed via phosphate–bismuth surface interactions, whose rates accelerate under illumination and elevated potentials, as illustrated in Figure 16a. These findings emphasize that stability should be assessed alongside kinetics rather than inferred from photocurrent alone, as demonstrated by the contrasting stability profiles in Figure 16b [164].
In borate buffers, a frequently observed “activation” effect has been attributed to trace Fe impurities, which deposit as an ultrathin oxyhydroxide layer that passivates surface traps and lowers interfacial resistance. Under optimized conditions, bare BiVO4 photoanodes have been reported to reach ≈4.5 mA/cm2 at 1.23 VRHE and surface Fe activation layers further boost photocurrent and enhance photostability by promoting Fe-layer formation on the surface. Comparable enhancements across Li+ Na+, and K+ borates support the view that Fe-mediated passivation, rather than alkali identity, governs this activation behavior. The photocurrent response before and after Fe impurity removal, as well as the effect of Fe2+ addition, are presented in Figure 16c [165]. Nevertheless, quantitative correlations between Fe concentration and performance gain remain limited and merit systematic study.
Beyond conventional phosphate and borate systems, bicarbonate electrolytes—including mixed or non-aqueous variants—have been shown to mitigate vanadium leaching, especially when supplemented with saturated V5+. This approach reduces morphological degradation and preserves optical density over extended operation, underscoring electrolyte engineering as a viable lever for durability improvement. The long-term stability of BiVO4 photoanodes in aqueous and mixed MeCN/NaHCO3 electrolytes is shown in Figure 16d [166].
Electrochemical conditioning protocols, such as controlled pre-bias or photocharging sequences, have also been demonstrated to enhance operational stability by tuning surface states and optimizing interfacial charge transfer pathways. These effects are synergistic with electrolyte composition, suggesting that combined strategies may yield more durable BiVO4 photoanodes [167].
In summary, electrolyte design directly shapes both surface kinetics and corrosion chemistry in BiVO4 photoanodes. Phosphate media can accelerate degradation without stabilization measures [164]; borate buffers “activate” BiVO4 via Fe-assisted passivation [165]; bicarbonates, particularly with V5+ supplementation, suppress vanadium loss and slow morphological decay [166]; and electrochemical conditioning provides an orthogonal route to stabilize interfacial dynamics [167]. Future studies are encouraged to report kinetic and stability metrics in parallel, clearly specify buffer composition and pH ranges, and combine electrolyte engineering (e.g., Fe-controlled borate or V5+-supplemented bicarbonate) with conditioning protocols to enable reproducible and robust long-term PEC operation.

5. Emerging Applications of BiVO4 Photoelectrodes in Solar Water Splitting and Beyond

While materials strategies (doping, heterojunctions, HTLs/overlayers, and electrolyte engineering) have substantially improved BiVO4 at the electrode level, practical relevance is established only in overall water splitting devices, where a BiVO4 photoanode is paired with a photocathode or photovoltaic element to operate unassisted (0 V external bias) or at low bias. This section surveys recent device architectures (PEC–PEC tandems, PEC–PV hybrids, monolithic vs. wired configurations) and applies a consistent set of performance benchmarks under AM 1.5G: (i) JPEC at 1.23 VRHE to normalize anodic OER kinetics, (ii) STH (%) as a device-level energy conversion metric, (iii) electrolyte/pH and applied bias for fair comparison, and (iv) durability (h) with supporting evidence (e.g., gas-tight collection and faradaic efficiency) when available. Building on these data, we discuss such as energy-level matching in tandems, reduction in ohmic/contact losses, membrane and electrode spacing design with hydrodynamic considerations, as well as pre-conditioning protocols and corrosion-mitigating electrolytes (e.g., impurity-controlled borate or vanadate-buffered media). The accompanying comparison table includes only reports with complete entries to provide a quantitative basis for assessing genuine progress and guiding the integration of BiVO4 into scalable modules.
Table 5 summarizes BiVO4-based photoanodes for overall water splitting under unbiased conditions, highlighting that the integration of oxygen evolution catalysts, Mo doping, surface nano-structuring, and organic hole transport layers can substantially enhance both JPEC and STH efficiencies. Among the reported systems, Mo:BiVO4 combined with a polycarbazole HTL and NiFeCoOx OEC exhibits one of the highest performances (JPEC ≈ 6.6 mA/cm2, STH ≈ 9%), emphasizing the importance of optimizing charge separation and catalytic activity in achieving high-efficiency photoanodes. [110].

5.1. BiVO4-Based Tandem System for Comprehensive Water Splitting

Achieving efficient overall water splitting without external bias requires that the top photoanode generates sufficient photovoltage while maintaining strong light absorption and operational stability. BiVO4, with a bandgap of ~2.4 eV and robust visible light absorption, remains a promising candidate. To overcome its intrinsic limitation in photovoltage, researchers have explored tandem PEC configurations that combine BiVO4 with additional light-harvesting materials.
A notable example is a BiVO4 nanocone/Fe(Ni)OOH photoanode coupled with a perovskite solar cell. Under AM 1.5G illumination, this system achieved unassisted overall water splitting with a STH efficiency of up to 6.2%, leveraging complementary light absorption and efficient interfacial charge transfer. The standalone PEC BiVO4-based component exhibited a photocurrent density of 5.82 ± 0.36 mA/cm2 at 1.23 VRHE The optical absorption mechanism of the nanocone substrate and the corresponding JV curves confirming this photocurrent enhancement are shown in Figure 17a,b [33].
Another recent advancement involves a monolithically integrated tandem system combining a BiVO4 photoanode with a Cu2O photocathode, both protected by conformal layers. Sitaaraman et al. fabricated a Mo-BiVO4/TiO2/FeOOH photoanode paired with a Cu2O/TiO2/MoS2 photocathode, forming an unassisted tandem PEC cell. The TiO2 layers serve as protective barriers and facilitate charge extraction, whereas FeOOH and MoS2 act as cocatalysts to accelerate the OER and HER. Under AM 1.5G illumination, the individual electrodes delivered ~0.81 mA/cm2 (photoanode) and –1.88 mA/cm2 (photocathode), and the assembled tandem achieved an unassisted current density of approximately 65.3 µA/cm2, with enhanced stability compared to unprotected devices. The LSV response, unassisted stability test, and energy band diagram of these tandem PEC devices are illustrated in Figure 17c–e [53].
The solar-to-hydrogen efficiency for tandem devices is calculated as
η S T H % = J s c m A c m 2 × 1.23 V × η F P T o t a l m W c m 2 A M 1.5 G
Here, Jsc represents the short-circuit photocurrent density corresponding to the rate of hydrogen generation expressed as current density, ηF is the Faradaic efficiency for hydrogen production (typically near 100%), and PTotal denotes the total solar irradiance intensity, set at 100 mW/cm2 under AM 1.5G conditions. By overlaying the current–voltage (JV) curve of the photoanode with those of the photocathode or photovoltaic cells, the operating point of the tandem device can be determined, defined by the short-circuit photocurrent density at 0 V applied bias. However, the actual device performance often falls short of the predicted operating point due to overpotentials between the two electrodes. This discrepancy arises because the JV curves of individual half-cells (including photovoltaics), referenced to the reversible hydrogen electrode (RHE) scale, do not account for resistive losses such as ohmic resistance and pH gradients present between the components.

5.2. PEC Cells for the Generation of Value-Added Chemicals

While PEC systems are primarily designed for overall water splitting, there is increasing interest in their application for the selective synthesis of value-added chemicals, such as hydrogen peroxide (H2O2), via partial oxidation or reduction under solar illumination. H2O2 can be selectively produced when the PEC system is appropriately engineered, exploiting its redox potentials of 1.77 VRHE (via 2e water oxidation) and 0.68 VRHE (via oxygen reduction).
The fundamental half-reactions for PEC-based H2O2 generation are summarized as follows:
Oxygen   evolution   reaction : 2 H 2 O + 4 h + 4 H + + O 2 ,   E ° = 1.23   V RHE
H 2 O 2 production   by   water   oxidation : 2 H 2 O + 2 h + 2 H + + H 2 O 2 ,     E ° = 1.77   V RHE
Hydrogen   evolution   reaction : 2 H + + 2 e H 2 ,   E ° = 0.0   V RHE
H 2 O 2 production   by   oxygen   reduction : O 2 + 2 H + + 2 e H 2 O 2 ,   E ° = 0.68   V RHE
These reactions highlight that by controlling the PEC environment and electrode materials, the pathways for partial oxidation or reduction can be preferentially promoted, enabling efficient and selective H2O2 production.
Recent advances have highlighted BiVO4-based systems as promising platforms for PEC hydrogen peroxide production. Wan et al. demonstrated that phosphate-modified BiVO4 (PBVO) photoanodes undergo dynamic anion exchange with bicarbonate (HCO3) at the semiconductor–electrolyte interface, which accelerates charge transfer and enhances the selectivity of the two-electron water oxidation pathway. This modification achieved an average Faradaic efficiency of 82.6% (maximum 92.1%) with a H2O2 production rate of 66.5 µmol/h, while the formation of H2PO4 intermediates suppressed H2O2 overoxidation. The LSV curves and charge balance plots confirming the selective H2O2 production on PBVO-2 photoanodes are shown in Figure 18a,b [173]. Under AM 1.5G illumination in a H-type PEC cell, PBVO-2 photoanodes exhibited ~1.6-fold higher photocurrent density than pristine BiVO4, maintained a solar-to-hydrogen efficiency of 1.27%, and enabled continuous H2O2 accumulation up to 2.34 × 10−3 M.
Complementarily, Shi et al. reported a heterojunction photocathode based on p-BiVO4/SnO2/NiNC, which delivered a H2O2 generation rate of 65.46 µmol/h with a Faradaic efficiency of 76.1%. LSV revealed significantly enhanced photocurrent density at 0.4 VRHE compared to pristine p-BVO and other oxide-modified photocathodes (TiOx, NiOx, ZnO), confirming the charge transport benefit of the SnO2 interlayer. The IPCE reached a maximum of 8.5%, and It measurements demonstrated excellent stability with photocurrent maintained above 0.15 mA/cm2 and only 8.9% loss after 20 h of operation under O2-saturated conditions. Integration with a Mo-doped BiVO4 photoanode in a tandem cell achieved nearly quantitative H2O2 selectivity. The chopped JV curves, transient current responses, and Faradaic efficiency data supporting this high selectivity are presented in Figure 18c–e [174].
Collectively, these studies demonstrate that surface anion modulation and rational heterojunction design enable efficient and selective PEC conversion of water to H2O2, establishing BiVO4-based systems as a viable platform for solar-driven production of value-added oxidants.

6. Challenges and Perspectives

Despite significant advances in BiVO4 electrodes over the past four decades, these materials still fall short of the requirements for practical applications, highlighting several challenges and opportunities.
Photocurrent Density Limitations: While recent reports have achieved photocurrent densities up to 6.0 mA/cm2 at 1.23 VRHE, this performance remains below the theoretical limit of 7.5 mA/cm2. Improving PEC water splitting efficiency, therefore, requires enhancements in light absorption, charge transfer, and carrier separation. A deeper understanding of factors influencing these processes could guide the design of more efficient BiVO4 electrodes. In particular, the role of crystal orientation on PEC performance remains underexplored, representing a promising avenue for future research. Additionally, synthesizing BiVO4 electrodes with multicomponent and multifunctional catalysts could provide additional active sites for hydrogen and oxygen evolution while promoting more effective charge separation.
Challenges in Understanding PEC Mechanisms: Understanding the fundamental PEC reaction mechanisms remains challenging. Most studies have prioritized performance improvements, often at the expense of mechanistic insight. Limited access to advance in situ techniques has hindered in-depth exploration of water oxidation on BiVO4. Approaches such as infrared spectroscopy, X-ray spectroscopy, and transient absorption spectroscopy could enable observation of reaction dynamics and intermediates at the molecular or atomic scale. Developing these methodologies will be essential for advancing the mechanistic understanding of PEC water oxidation.
Scaling Up BiVO4-Based Photoanodes: Current research largely involves small-area BiVO4 photoanodes (~1.0 cm2), whereas practical deployment demands large-area materials with high uniformity and crystallinity. Fabrication via electroplated BiOI conversion faces challenges in maintaining uniformity across large substrates due to increased resistance, nonuniform films, and nonlinear diffusion of reactants and products. Developing scalable and cost-effective synthesis strategies for large-area, high-quality photoelectrodes will be critical for commercial applications.
Design of Tandem PEC Systems: Tandem configurations using dual- or three-electrode setups are common in practical PEC systems. A feasible approach for real-world devices involves compartmentalization with membranes, along with careful optimization of electrode spacing, membrane design, and pressure balance. Tandem PEC modules often integrate multijunction or perovskite photovoltaic cells, requiring precise matching of energy levels between the photoanode and photocathode. Further engineering efforts, including fluid dynamics optimization, are needed to maximize device performance under operational conditions.
Advanced/Operando Spectroscopy and Diagnostics: To convert performance gains into durable operation, buried-interface processes must be observed under realistic bias, illumination, and electrolyte conditions. High-energy-resolution X-ray absorption (HERFD-XANES/XAS) at the Bi L and V K edges can track oxidation-state shifts, ligand fields, and local coordination during OER, revealing whether interlayers/overlayers stabilize Bi–O and V–O polyhedra or merely delay dissolution. Ambient-pressure/operando XPS resolves surface terminations, adsorbates (e.g., phosphate/borate/vanadate), and HTL-derived functional groups that modulate band bending and recombination velocities. Operando GIWAXS/XRD identifies phase transformations (e.g., BiPO4 signatures, amorphization), while Raman/IR spectroelectrochemistry reports reaction intermediates and lattice strain. Charge-dynamics probes—IMPS/EIS to deconvolute η_sep vs. η_trans, transient absorption (fs–µs) and time-resolved photoluminescence to quantify interfacial transfer constants, and UV–Vis spectroelectrochemistry to monitor polaron/charge accumulation—link kinetic rate constants to optical populations. Finally, pairing photocurrent with operando dissolution (e.g., ICP-MS in a scanning flow cell or EQCM) closes the gap between apparent activity and true stability by quantifying Bi/V loss in real time.
Standardization and Reporting of Stability: To ensure comparability across studies, we recommend (i) co-reporting JV (or ABPE) with operando dissolution traces on the same electrode; (ii) specifying light intensity (AM 1.5G), temperature, pH, buffer composition and impurities (e.g., trace Fe), bias protocols (hold vs. scan; scan rate), and operation time; (iii) adopting common endurance metrics (e.g., current retention and dissolution yield after 10–24 h at a defined potential); and (iv) providing raw spectra and analysis scripts for XAS/XPS/Raman together with IMPS/EIS fitting models. Such practices reduce ambiguity between kinetic- and corrosion-driven losses and accelerate translation to device design.
Diagnostics for Scale-Up: As devices move beyond ~1 cm2, spatial non-uniformities dominate. Integrating spatially resolved tools—scanning photoelectrochemical microscopy (SPECM), light-beam-induced current mapping, micro-XRD/µ-XRF, or X-ray/optical tomography—can identify local recombination hot spots, catalyst delamination, and electrolyte transport limitations, informing fluid dynamics optimization and contact/grid design for large-area modules.
Given these multifaceted challenges, interdisciplinary collaboration is increasingly important. Combining expertise from materials science, physical chemistry, surface science, and advanced characterization techniques will be essential to develop more stable and efficient BiVO4-based photoelectrodes for PEC water splitting.

Author Contributions

Conceptualization, J.-Y.K.; methodology, B.D.N. and I.-H.C.; formal analysis, B.D.N., I.-H.C. and J.-Y.K.; investigation, B.D.N. and I.-H.C.; writing—original draft preparation, J.-Y.K., B.D.N. and I.-H.C.; writing—review and editing, J.-Y.K.; supervision, J.-Y.K.; project administration, J.-Y.K.; funding acquisition, J.-Y.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported in part by a Human Resources Development Program of the Korea Institute of Energy Technology Evaluation and Planning (KETEP) grant funded by the Ministry of Trade, Industry and Energy, Republic of Korea (No. RS-2023-00237035). This work was also supported by the Korea Institute of Energy Technology Evaluation and Planning (KETEP) and the Ministry of Trade, Industry and Energy (MOTIE) of the Republic of Korea (RS-2025-02422969). This research was also supported by the Nano and Material Technology Development Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science and ICT (RS-2024-00452380).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Abid, M.Z.; Rafiq, K.; Rauf, A.; Althomali, R.H.; Jin, R.; Hussain, E. Interface engineering of BiVO4/Zn3V2O8 heterocatalysts for escalating the synergism: Impact of Cu electron mediator for overall water splitting. Renew. Energy 2024, 225, 120223. [Google Scholar] [CrossRef]
  2. Sivula, K.; van de Krol, R. Semiconducting materials for photoelectrochemical energy conversion. Nat. Rev. Mater. 2016, 1, 15010. [Google Scholar] [CrossRef]
  3. 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, 1908–1971. [Google Scholar] [CrossRef]
  4. Takanabe, K. Photocatalytic water splitting: Quantitative approaches toward photocatalysis by design. ACS Catal. 2017, 7, 8006–8022. [Google Scholar] [CrossRef]
  5. Wang, Z.; Guo, Y.; Liu, M.; Liu, X.; Zhang, H.; Jiang, W.; Wang, P.; Zheng, Z.; Liu, Y.; Cheng, H.; et al. Boosting H2 Production from a BiVO4 Photoelectrochemical Biomass Fuel Cell by the Construction of a Bridge for Charge and Energy Transfer. Adv. Mater. 2022, 34, e2201594. [Google Scholar] [CrossRef]
  6. Staffell, I.; Scamman, D.; Velazquez Abad, A.; Balcombe, P.; Dodds, P.E.; Ekins, P.; Shah, N.; Ward, K.R. The role of hydrogen and fuel cells in the global energy system. Energy Environ. Sci. 2019, 12, 463–491. [Google Scholar] [CrossRef]
  7. Nikolaidis, P.; Poullikkas, A. A comparative overview of hydrogen production processes. Renew. Sustain. Energy Rev. 2016, 67, 597–611. [Google Scholar] [CrossRef]
  8. Dincer, I.; Acar, C. Review and evaluation of hydrogen production methods for better sustainability. Int. J. Hydrogen Energy 2015, 40, 11094–11111. [Google Scholar] [CrossRef]
  9. Walter, M.G.; Warren, E.L.; McKone, J.R.; Boettcher, S.W.; Mi, Q.; Santori, E.A.; Lewis, N.S. Solar water splitting cells. Chem. Rev. 2010, 110, 6446–6473. [Google Scholar] [CrossRef]
  10. Chen, S.; Takata, T.; Domen, K. Particulate photocatalysts for overall water splitting. Nat. Rev. Mater. 2017, 2, 17050. [Google Scholar] [CrossRef]
  11. Ardo, S.; Fernandez Rivas, D.; Modestino, M.A.; Schulze Greiving, V.; Abdi, F.F.; Alarcon Llado, E.; Artero, V.; Ayers, K.; Battaglia, C.; Becker, J.-P.; et al. Pathways to electrochemical solar-hydrogen technologies. Energy Environ. Sci. 2018, 11, 2768–2783. [Google Scholar] [CrossRef]
  12. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef] [PubMed]
  13. Li, J.; Wu, N. Semiconductor-based photocatalysts and photoelectrochemical cells for solar fuel generation: A review. Catal. Sci. Technol. 2015, 5, 1360–1384. [Google Scholar] [CrossRef]
  14. Yang, Y.; Niu, S.; Han, D.; Liu, T.; Wang, G.; Li, Y. Progress in developing metal oxide nanomaterials for photoelectrochemical water splitting. Adv. Energy Mater. 2017, 7, 1700555. [Google Scholar] [CrossRef]
  15. Li, C.; Fan, W.; Chen, S.; Zhang, F. Effective charge carrier utilization of BiVO4 for solar overall water splitting. Chemistry 2022, 28, e202201812. [Google Scholar] [CrossRef]
  16. Park, Y.; McDonald, K.J.; Choi, K.S. Progress in bismuth vanadate photoanodes for use in solar water oxidation. Chem. Soc. Rev. 2013, 42, 2321–2337. [Google Scholar] [CrossRef]
  17. Tan, H.L.; Amal, R.; Ng, Y.H. Alternative strategies in improving the photocatalytic and photoelectrochemical activities of visible light-driven BiVO4: A review. J. Mater. Chem. A 2017, 5, 16498–16521. [Google Scholar] [CrossRef]
  18. Kudo, A.; Ueda, K.; Kato, H.; Mikami, I. Photocatalytic O2 evolution under visible light irradiation over BiVO4 in an aqueous AgNO3 solution. Catal. Lett. 1998, 53, 229–230. [Google Scholar] [CrossRef]
  19. Cooper, J.K.; Gul, S.; Toma, F.M. Electronic structure of monoclinic BiVO4. Chem. Mater. 2014, 26, 5365–5373. [Google Scholar] [CrossRef]
  20. 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]
  21. Jiang, C.; Moniz, S.J.A.; Wang, A.; Zhang, T.; Tang, J. Photoelectrochemical Devices for Solar Water Splitting—Materials and Challenges. Chem. Soc. Rev. 2017, 46, 4645–4660. [Google Scholar] [CrossRef]
  22. Cooper, J.K.; Scott, S.B.; Ling, Y.; Yang, J.; Hao, S.; Li, Y.; Toma, F.M.; Stutzmann, M.; Lakshmi, K.V.; Sharp, I.D. Role of Hydrogen in Defining the N-Type Character of BiVO4 Photoanodes. Chem. Mater. 2016, 28, 5761–5771. [Google Scholar] [CrossRef]
  23. Lyu, Y.; Zhou, Y.; Zheng, J. Unbiased photoelectrochemical tandem configuration for water splitting. J. Am. Chem. Soc. 2013, 135, 18661–18664. [Google Scholar] [CrossRef]
  24. Kuang, Y.; Jia, Q.; Nishiyama, H.; Yamada, T.; Kudo, A.; Domen, K. A front-illuminated nanostructured transparent BiVO4 photoanode for >2% efficient water splitting. Adv. Energy Mater. 2016, 6, 1501645. [Google Scholar] [CrossRef]
  25. Kho, Y.K.; Teoh, W.Y.; Iwase, A.; Mädler, L.; Kudo, A.; Amal, R. Flame preparation of visible-light responsive BiVO4 photocatalysts. J. Phys. Chem. C 2011, 115, 8591–8596. [Google Scholar]
  26. Zhang, J.; Yu, J.; Zhang, Y.; Li, Q.; Gong, J.R. Efficient and stable photo-oxidation of water by a bismuth vanadate photoanode. J. Am. Chem. Soc. 2011, 133, 19306–19309. [Google Scholar]
  27. Pihosh, Y.; Turkevych, I.; Mawatari, K.; Uemura, J.; Kazoe, Y.; Kosar, S.; Makita, K.; Sugaya, T.; Matsui, T.; Fujita, D.; et al. Photocatalytic generation of hydrogen by core-shell WO3/BiVO4 nanorods with ultimate water splitting efficiency. Sci. Rep. 2015, 5, 11141. [Google Scholar] [CrossRef]
  28. Abdi, F.F.; Savenije, T.J.; May, M.M.; Dam, B.; Van De Krol, R. The origin of slow carrier transport in BiVO4 thin film photoanodes. J. Phys. Chem. Lett. 2013, 4, 2752–2757. [Google Scholar] [CrossRef]
  29. 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]
  30. Nellist, M.R.; Qiu, J.; Laskowski, F.A.; Toma, F.M.; Boettcher, S.W. Potential-sensing electrochemical AFM shows CoPi as a hole collector and Oxygen Evolution Catalyst on BiVO4 Water-Splitting photoanodes. ACS Energy Lett. 2018, 3, 2286–2291. [Google Scholar] [CrossRef]
  31. Kim, J.H.; Lee, J.S. Elaborately modified BiVO4 photoanodes for solar water splitting. Adv. Mater. 2019, 31, e1806938. [Google Scholar] [CrossRef]
  32. Wang, S.; Chen, P.; Bai, Y.; Yun, J.-H.; Liu, G.; Wang, L. New BiVO4 dual photoanodes with enriched oxygen vacancies for efficient solar-driven water splitting. Adv. Mater. 2018, 30, 1800486. [Google Scholar] [CrossRef] [PubMed]
  33. Qiu, Y.; Liu, W.; Chen, W.; Chen, W.; Zhou, G.; Hsu, P.-C.; Zhang, R.; Liang, Z.; Fan, S.; Zhang, Y.; et al. Efficient solar-driven water splitting by nanocone BiVO4–perovskite tandem cells. Sci. Adv. 2016, 2, e1501764. [Google Scholar] [CrossRef] [PubMed]
  34. Kim, J.H.; Jo, Y.; Kim, J.H.; Jang, J.W.; Kang, H.J.; Lee, Y.H.; Kim, D.S.; Jun, Y.; Lee, J.S. Wireless solar water splitting device with robust cobalt-catalyzed, dual-doped BiVO4 photoanode and perovskite solar cell in tandem: A dual absorber artificial leaf. ACS Nano 2015, 9, 11820–11829. [Google Scholar] [CrossRef] [PubMed]
  35. Luo, W.; Li, Z.; Zhang, M.; Feng, J.; Liu, J.; Zhao, Z.; Wang, S.; Yu, T.; Zou, Z. Quantitative analysis and visualized evidence for high charge separation efficiency in a solid/liquid bulk heterojunction. Adv. Energy Mater. 2014, 4, 1301785. [Google Scholar]
  36. Thomas, N.; Singh, V.; Ahmed, N.; Trinh, D.; Kuss, S.; Halasyamani, P.S.; Parkinson, B.A. Factors in the Metal Doping of BiVO4 for Improved Photoelectrochemical Water Oxidation. J. Phys. Chem. C 2011, 115, 1–9. [Google Scholar]
  37. Berglund, S.P.; Rettie, A.J.E.; Hoang, S.; Mullins, C.B. Incorporation of Mo and W into nanostructured BiVO4 films for efficient photoelectrochemical water oxidation. Phys. Chem. Chem. Phys. 2012, 14, 7065–7075. [Google Scholar] [CrossRef]
  38. Hong, S.J.; Lee, S.; Jang, J.S.; Lee, J.S. Heterojunction BiVO4/WO3 electrodes for enhanced photoactivity of water oxidation. Energy Environ. Sci. 2011, 4, 1781–1787. [Google Scholar] [CrossRef]
  39. Zhang, L.; Zhang, X.; Zhang, Y.; Liu, J.; Wang, Y.; Li, J.; Li, Y.; Wang, X. BiVO4–graphene catalyst and its high photocatalytic performance under visible light irradiation. Mater. Chem. Phys. 2012, 132, 614–619. [Google Scholar]
  40. Zhong, D.K.; Choi, S.; Gamelin, D.R. Near-Complete Suppression of Surface Recombination in Solar Photoelectrolysis by “Co-Pi” Catalyst-Modified W:BiVO4. J. Am. Chem. Soc. 2011, 133, 18370–18377. [Google Scholar] [CrossRef]
  41. Pihosh, Y.; Turkevych, I.; Mawatari, K.; Asai, T.; Hisatomi, T.; Uemura, J.; Tosa, M.; Shimamura, K.; Kubota, J.; Domen, K.; et al. Nanostructured WO3/BiVO4 photoanodes for efficient photoelectrochemical water splitting. Small 2014, 10, 3692–3699. [Google Scholar] [CrossRef]
  42. Trześniewski, B.J.; Digdaya, I.A.; Nagaki, T.; Ravishankar, S.; Herraiz-Cardona, I.; Vermaas, D.A.; Longo, A.; Gimenez, S.; Smith, W.A. Near-complete suppression of surface losses and total internal quantum efficiency in BiVO4 photoanodes. Energy Environ. Sci. 2017, 10, 1517–1529. [Google Scholar] [CrossRef]
  43. Baek, J.H.; Kim, B.J.; Han, G.S.; Hwang, S.W.; Kim, D.R.; Cho, I.S.; Jung, H.S. BiVO4/WO3/SnO2 Double-Heterojunction Photoanode with Enhanced Charge Separation and Visible Transparency for Bias-Free Solar Water Splitting with a Perovskite Solar Cell. ACS Appl. Mater. Interfaces 2017, 9, 1479–1487. [Google Scholar] [CrossRef] [PubMed]
  44. Saraswat, S.K.; Rodene, D.D.; Gupta, R.B. Recent Advancements in Semiconductor Materials for Photoelectrochemical Water Splitting for Hydrogen Production Using Visible Light. Renew. Sustain. Energy Rev. 2018, 89, 228–248. [Google Scholar] [CrossRef]
  45. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  46. Sharma, C.; Pooja, D.; Thakur, A.; Negi, Y.S. Review—Combining Experimental and Engineering Aspects of Catalyst Design for Photoelectrochemical Water Splitting. ECS Adv. 2022, 1, 030501. [Google Scholar] [CrossRef]
  47. Salih, A.K.; Khan, A.Z.; Drmosh, Q.; Kandiel, T.; Qamar, M.; Jahangir, T.; Ton-That, C.; Yamani, Z. Nanostructured BiVO4 Photoanodes Fabricated by Vanadium-Infused Interaction for Efficient Solar Water Splitting. ACS Appl. Nano Mater. 2024, 7, 14115–14122. [Google Scholar] [CrossRef]
  48. Scherino, L.; Giaquinto, M.; Micco, A.; Aliberti, A.; Bobeico, E.; La Ferrara, V.; Ruvo, M.; Ricciardi, A.; Cusano, A. A Time-Efficient Dip Coating Technique for the Deposition of Microgels onto the Optical Fiber Tip. Fibers 2018, 6, 72. [Google Scholar] [CrossRef]
  49. Dong, G.; Chen, T.; Kou, F.; Xie, F.; Xiao, C.; Liang, J.; Lou, C.; Zhuang, J.; Du, S. Promoting the Photoelectrochemical Properties of BiVO4 Photoanode via Dual Modification with CdS Nanoparticles and NiFe-LDH Nanosheets. Nanomaterials 2024, 14, 1100. [Google Scholar] [CrossRef]
  50. Wang, Z.-Q.; Wang, H. Fabrication of Cocatalyst NiO-Modified BiVO4 Composites for Enhanced Photoelectrochemical Performances. Front. Chem. 2022, 10, 864143. [Google Scholar] [CrossRef]
  51. Xie, Z.; Liu, X.; Jia, P.; Zou, Y.; Jiang, J. Facile Construction of a BiVO4/CoV-LDHs/Ag Photoanode for Enhanced Photo-Electrocatalytic Glycerol Oxidation and Hydrogen Evolution. Energy Adv. 2023, 2, 1366–1374. [Google Scholar] [CrossRef]
  52. Ho-Kimura, S.; Soontornchaiyakul, W.; Yamaguchi, Y.; Kudo, A. Preparation of Nanoparticle Porous-Structured BiVO4 Photoanodes by a New Two-Step Electrochemical Deposition Method for Water Splitting. Catalysts 2021, 11, 136. [Google Scholar] [CrossRef]
  53. Sitaaraman, S.R.; Nirmala Grace, A.; Sellappan, R. Photoelectrochemical Performance of a Spin Coated TiO2 Protected BiVO4-Cu2O Thin Film Tandem Cell for Unassisted Solar Water Splitting. RSC Adv. 2022, 12, 31380–31391. [Google Scholar] [CrossRef] [PubMed]
  54. Seo, J.-W.; Ha, S.-B.; Song, I.-C.; Kim, J.-Y. PbS Quantum Dots-Decorated BiVO4 Photoanodes for Highly Efficient Photoelectrochemical Hydrogen Production. Nanomaterials 2023, 13, 799. [Google Scholar] [CrossRef] [PubMed]
  55. Kudo, A.; Omori, K.; Kato, H. A Novel Aqueous Process for Preparation of Crystal Form-Controlled and Highly Crystalline BiVO4 Powder from Layered Vanadates at Room Temperature and Its Photocatalytic and Photophysical Properties. J. Am. Chem. Soc. 1999, 121, 11459–11467. [Google Scholar] [CrossRef]
  56. Kwan, T.H.; Wu, X.; Yao, Q. Parameter Sizing and Stability Analysis of a Highway Fuel Cell Electric Bus Power System Using a Multi-Objective Optimization Approach. Int. J. Hydrogen Energy 2018, 43, 20976–20992. [Google Scholar] [CrossRef]
  57. Xia, T.; Chen, M.; Xiao, L.; Fan, W.; Mao, B.; Xu, D.; Guan, P.; Zhu, J.; Shi, W. Phosphorus doped BiVO4 photoanodes prepared by dip coating for improved photoelectrochemical water splitting. J. Taiwan Inst. Chem. Eng. 2018, 82, 582–589. [Google Scholar] [CrossRef]
  58. Sun, Y.; Zheng, Y.; Wang, R.; Fan, J.; Liu, Y. Direct-Current Piezoelectric Nanogenerator Based on Two-Layer Zinc Oxide Nanorod Arrays with Equal c-Axis Orientation for Energy Harvesting. Chem. Eng. J. 2021, 426, 131262. [Google Scholar] [CrossRef]
  59. Rohloff, M.; Anke, B.; Wiedemann, D.; Ulpe, A.C.; Kasian, O.; Zhang, S.; Scheu, C.; Bredow, T.; Lerch, M.; Fischer, A. Synthesis and Doping Strategies to Improve the Photoelectrochemical Water Oxidation Activity of BiVO4 Photoanodes. Z. Phys. Chem. 2020, 234, 655–682. [Google Scholar] [CrossRef]
  60. Moniz, S.J.A.; Shevlin, S.A.; Martin, D.J.; Guo, Z.-X.; Tang, J. Visible-Light Driven Heterojunction Photocatalysts for Water Splitting—A Critical Review. Energy Environ. Sci. 2015, 8, 731–759. [Google Scholar] [CrossRef]
  61. Liu, R.; Qiang, L.-S.; Yang, W.-D.; Liu, H.-Y. The Effect of Calcination Conditions on the Morphology, the Architecture and the Photo-Electrical Properties of TiO2 Nanotube Arrays. Mater. Res. Bull. 2013, 48, 1458–1467. [Google Scholar] [CrossRef]
  62. Lai, W.; Di, L.; Zhao, C.; Tian, Y.; Duan, Y.; Pan, Y.; Ye, D.; Jiang, L.; Guo, Y.; He, G.; et al. Electrospray Deposition for Electronic Thin Films on 3D Freeform Surfaces: From Mechanisms to Applications. Adv Mater. Technol. 2024, 9, 2400192. [Google Scholar] [CrossRef]
  63. Ilyas, A.; Rafiq, K.; Abid, M.Z.; Rauf, A.; Hussain, E. Growth of Villi-Microstructured Bismuth Vanadate (Vm-BiVO4) for Photocatalytic Degradation of Crystal Violet Dye. RSC Adv. 2023, 13, 2379–2391. [Google Scholar] [CrossRef] [PubMed]
  64. Montañés, L.; Mesa, C.A.; Gutiérrez-Blanco, A.; Robles, C.; Julián-López, B.; Giménez, S. Facile Surfactant-Assisted Synthesis of BiVO4 Nanoparticulate Films for Solar Water Splitting. Catalysts 2021, 11, 1244. [Google Scholar] [CrossRef]
  65. Thalluri, S.M.; Bai, L.; Lv, C.; Huang, Z.; Hu, X.; Liu, L. Strategies for Semiconductor/Electrocatalyst Coupling toward Solar-Driven Water Splitting. Adv. Sci. 2020, 7, 1902102. [Google Scholar] [CrossRef]
  66. Liu, C.; Su, J.; Zhou, J.; Guo, L. A Multistep Ion Exchange Approach for Fabrication of Porous BiVO4 Nanorod Arrays on Transparent Conductive Substrate. ACS Sustain. Chem. Eng. 2016, 4, 4492–4497. [Google Scholar] [CrossRef]
  67. Shan, L.; Yuan, M.; Ma, C.; Lu, C.; Fang, Z.; Lin, S.; Li, D.; Wu, Z.; Dong, L.; Qian, G.; et al. Z-Scheme CdS/BiVO4 Photocatalyst for Highly Efficient Photodegradation of Organic Pollutants and DFT Calculation. Available online: https://www.researchgate.net/publication/350549778_Z-scheme_CdSBiVO4_Photocatalyst_for_Highly_Efficient_Photodegradation_of_Organic_Pollutants_and_DFT_Calculation (accessed on 1 April 2021).
  68. De Laat, J.; Le, T.G. Effects of Chloride Ions on the Iron(III)-Catalyzed Decomposition of Hydrogen Peroxide and on the Efficiency of the Fenton-like Oxidation Process. Appl. Catal. B Environ. 2006, 66, 137–146. [Google Scholar]
  69. Laleh, M.; Kargar, F. Suppression of Chromium Depletion and Sensitization in Austenitic Stainless Steel by Surface Mechanical Attrition Treatment. Mater. Lett. 2011, 65, 1935–1937. [Google Scholar] [CrossRef]
  70. Na, S.; Seo, S.; Lee, H. Recent Developments of Advanced Ti3+-Self-Doped TiO2 for Efficient Visible-Light-Driven Photocatalysis. Catalysts 2020, 10, 679. [Google Scholar]
  71. Haji Yassin, S.N.; Sim, A.S.L.; Jennings, J.R. Photoelectrochemical Evaluation of SILAR-Deposited Nanoporous BiVO4 Photoanodes for Solar-Driven Water Splitting. Nano Mater. Sci. 2020, 2, 227–234. [Google Scholar] [CrossRef]
  72. Huang, X.H.; Xia, X.H.; Yuan, Y.F.; Zhou, F. Porous ZnO Nanosheets Grown on Copper Substrates as Anodes for Lithium Ion Batteries. Electrochim. Acta 2011, 56, 4960–4965. [Google Scholar] [CrossRef]
  73. Yang, J.; Wang, D.; Han, H.; Li, C. Roles of Cocatalysts in Photocatalysis and Photoelectrocatalysis. Acc. Chem. Res. 2013, 46, 1900–1909. [Google Scholar] [CrossRef] [PubMed]
  74. Zhang, Y.; Yang, Z. The Mechanism of the High Resistance to Sulfur Poisoning of the Rhenium Doped Nickel/Yttria-Stabilized Zirconia. Appl. Surf. Sci. 2018, 447, 561–568. [Google Scholar] [CrossRef]
  75. Liang, Y.; Messinger, J. Improving BiVO4 photoanodes for solar water splitting through surface passivation. Phys. Chem. Chem. Phys 2014, 16, 12014–12020. [Google Scholar] [CrossRef]
  76. Lin, J.; Li, X.; Wang, Z.; Liu, R.; Pan, H.; Zhao, Y.; Kong, L.; Li, Y.; Zhu, S.; Wang, L. Organic ligand nanoarchitectonics for BiVO4 photoanodes surface passivation and cocatalyst grafting. Nano Res 2024, 17, 3667–3674. [Google Scholar] [CrossRef]
  77. Qayum, A.; Guo, M.; Wei, J.; Dong, S.; Jiao, X.; Chen, D.; Wang, T. An in situ combustion method for scale-up fabrication of BiVO4 photoanodes with enhanced long-term photostability for unassisted solar water splitting. J. Mater. Chem. A 2020, 8, 10989–10997. [Google Scholar] [CrossRef]
  78. Song, J.; Cha, J.; Lee, M.G.; Jeong, H.W.; Seo, S.; Yoo, J.A.; Kim, T.L.; Lee, J.; No, H.; Kim, D.H.; et al. Template-engineered epitaxial BiVO4 photoanodes for efficient solar water splitting. J. Mater. Chem. A 2017, 5, 18831–18838. [Google Scholar] [CrossRef]
  79. Shi, J.; Zhao, X.; Li, C. Surface passivation engineering for photoelectrochemical water splitting. Catalysts 2023, 13, 217. [Google Scholar] [CrossRef]
  80. Tokunaga, S.; Kato, H.; Kudo, A. Selective Preparation of Monoclinic and Tetragonal BiVO4 with Scheelite Structure and Their Photocatalytic Properties. Chem. Mater. 2001, 13, 4624–4628. [Google Scholar] [CrossRef]
  81. Chang, T.-F.M.; Sone, M.; Shibata, A.; Ishiyama, C.; Higo, Y. Bright Nickel Film Deposited by Supercritical Carbon Dioxide Emulsion Using Additive-Free Watts Bath. Electrochim. Acta 2010, 55, 6469–6475. [Google Scholar] [CrossRef]
  82. Ojo, A.A.; Dharmadasa, I.M. Electroplating of semiconductor materials for applications in large area electronics: A review. Coatings 2018, 8, 262. [Google Scholar] [CrossRef]
  83. Marschall, R. Semiconductor Composites: Strategies for Enhancing Charge Carrier Separation to Improve Photocatalytic Activity. Adv. Funct. Mater. 2014, 24, 2421–2440. [Google Scholar] [CrossRef]
  84. Murugan, C.; Mary, A.S.; Velmurugan, R.; Subramanian, B.; Murugan, P.; Pandikumar, A. Investigating the interfacial charge transfer between electrodeposited BiVO4 and pulsed laser-deposited Co3O4 p-n junction photoanode in photoelectrocatalytic water splitting. Chem. Eng. J. 2024, 483, 149104. [Google Scholar] [CrossRef]
  85. Nguyen, C.M.; Nguyen, T.N.; Nguyen, H.N.; Nguyen, S.N.; Nguyen, M.T. Photoelectrochemical water splitting with BiVO4 photoanodes synthesized by different methods. Reac. Kinet. Mech. Catal. 2025, 138, 2479–2493. [Google Scholar] [CrossRef]
  86. Obeso, J.L.; Martínez-Ahumada, E.; López-Olvera, A.; Ortiz-Landeros, J.; Lara-García, H.A.; Balmaseda, J.; López-Morales, S.; Sánchez-González, E.; Solis-Ibarra, D.; Leyva, C.; et al. MOF-303 as an effective adsorbent to clean CH4: SO2 capture and detection. ACS Appl. Energy Mater. 2022, 6, 9084–9091. [Google Scholar] [CrossRef]
  87. Zhang, Q.; Zhang, X.; Qiao, S.; Lei, D.; Wang, Q.; Shi, X.; Huang, C.; Lu, W.; Yang, S.; Tian, Y.; et al. Synthesis of the Ni2P–Co Mott–Schottky junction as an electrocatalyst to boost sulfur conversion kinetics and application in separator modification in Li–S batteries. ACS Appl. Mater. Interfaces 2023, 15, 5253–5264. [Google Scholar] [CrossRef]
  88. Tayebi, M.; Lee, B. The effects of W/Mo-co-doped BiVO4 photoanodes for improving photoelectrochemical water splitting performance. Catal. Today 2021, 361, 183–190. [Google Scholar] [CrossRef]
  89. Huang, M.; Bian, J.; Xiong, W.; Huang, C.; Zhang, R. Low-dimensional Mo:BiVO4 photoanodes for enhanced photoelectrochemical activity. J. Mater. Chem. A 2018, 6, 3602–3609. [Google Scholar] [CrossRef]
  90. Pham, T.-D.; Lee, B.-K.; Nguyen, V.N.; Dao, V.-D.; Nguyen, T.D.C.; Mai, H.T.; Dao, V.-D.; Nguyen, M.P.; Nguyen, V.N. Fabrication of titanium-doped BiVO4 as a novel visible light-driven photocatalyst for degradation of residual tetracycline pollutant. Ceram. Int. 2021, 47, 34253–34259. [Google Scholar]
  91. Li, X.; Tang, Q.; Wang, S.; Liu, Y.; Chen, G.; Zhao, W.; Han, M.; Liu, G. Fe-N-C nanoparticles as a highly efficient electrocatalyst for oxygen reduction reaction in alkaline electrolyte. Int. J. Hydrogen Energy 2020, 45, 1686–1695. [Google Scholar]
  92. Mohd Raub, A.A.; Bahru, R.; Mohd Nashruddin, S.N.A.; Yunas, J. Advances of Nanostructured Metal Oxide as Photoanode in Photoelectrochemical (PEC) Water Splitting Application. Heliyon 2024, 10, e39079. [Google Scholar]
  93. Fang, G.; Liu, Z.; Han, C. Enhancing the PEC Water Splitting Performance of BiVO4 Co-Modifying with NiFeOOH and Co-Pi Double Layer Cocatalysts. Appl. Surf. Sci. 2020, 515, 146095. [Google Scholar]
  94. Polo, A.; Nomellini, C.; Marra, G.; Selli, E.; Dozzi, M.V. WO3/BiVO4 heterojunction photoanodes: Optimized photoelectrochemical performance in relation to both oxides layer thickness. Catal. Today 2025, 446, 115137. [Google Scholar] [CrossRef]
  95. Bai, S.; Cao, S.; Yan, X.; Luo, R.; Wang, X. Two-step electrodeposition to fabricate the p–n heterojunction of a Cu2O/BiVO4 photoanode for the enhancement of photoelectrochemical water splitting. Dalton Trans. 2018, 47, 6763–6771. [Google Scholar] [CrossRef] [PubMed]
  96. Liu, C.; Yu, L.; Xue, K.; Luo, H.; Zhang, Y. Phase engineering of 1 T-MoS2 on BiVO4 photoanode with p-n junctions establishing high-speed charges transport channels for efficient photoelectrochemical water splitting. Chem. Eng. J. 2023, 470, 144965. [Google Scholar]
  97. Luo, W.; Tang, T.; Ma, T.; Xu, H.; Cui, Y.; Li, X.; Zhou, H.; Fan, B.; He, Y. Engineering a Tandem S-Scheme TiO2@SrTiO3@BiVO4 Photoanode for Superior PEC Water Oxidation. Electrochim. Acta 2025, 539, 147105. [Google Scholar]
  98. Dong, L.; Zhang, J.; Wang, J.; Dong, S.; Song, M.; Zhao, C. In situ unravelling surface plasmon resonance subject-object role of BiVO4@Ag in photocatalytic water splitting. J. Colloid Interface Sci. 2025, 701, 138656. [Google Scholar] [CrossRef]
  99. Hu, Y.; Ma, Y.; Liu, C. Sunlight Selective Photodeposition of CoOx(OH)y and NiOx(OH)y on Truncated Bipyramidal BiVO4 for Highly Efficient Photocatalysis. ACS Appl. Mater. Interfaces 2017, 9, 20436–20442. [Google Scholar]
  100. Sun, Z.J.; Ma, J.; Ma, Y.; Wang, J.G.; Hou, Y.J.; Chen, G.D.; Dissanayake, M.A.W.C.L.N.K.; Hettiarachchi, Y.M.A.G.; Liu, J. Surface Reconstruction and Passivation of BiVO4 Photoanodes Depending on the “Structure Breaker” Cs+. JACS Au 2024, 4, 196–205. [Google Scholar]
  101. Kim, D.S.; Lee, K.W.; Jung, S.Y.; Choi, J.H.; Lee, H.H.; Yang, W.S.; Cho, Y.S.; Yang, J.W.; Kang, J.; Boettcher, S.W.; et al. Efficient BiVO4/CoFeOxHy photoanodes using controlled annealing and conformal linear-sweep electrocatalyst photodeposition. Appl. Surf. Sci. 2025, 689, 162470. [Google Scholar] [CrossRef]
  102. Zhou, B.; Xu, S.; Wu, L.; Li, M.; Chong, Y.; Qiu, Y.; Chen, G.; Zhao, Y.; Feng, C.; Ye, D.; et al. Strain-Engineering of Mesoporous Cs3Bi2Br9/BiVO4 S-Scheme Heterojunction for Efficient CO2 Photoreduction. Small 2023, 19, 2302058. [Google Scholar] [CrossRef]
  103. Arunachalam, M.; Yun, G.; Lee, H.S.; Ahn, K.-S.; Heo, J.; Kang, S.H. Effects of Al2O3 coating on BiVO4 and Mo-doped BiVO4 film for solar water oxidation. J. Electrochem. Sci. Technol 2019, 10, 424–432. [Google Scholar] [CrossRef]
  104. Zhang, Z.; Song, Y.; Xiang, Y.; Zhu, Z. Vacancy defect engineered BiVO4 with low-index surfaces for photocatalytic application: A first principles study. RSC Adv. 2022, 12, 31317–31325. [Google Scholar] [CrossRef]
  105. Ponchio, C.; Nareejun, W. Laser-Assisted FTO/WO3/BiVO4 Photoanode Fabrication for Enhanced Photoelectrocatalytic Performance and Durability for Organic Dye Degradation. J. Appl. Res. Sci. Technol. 2025, 24, 260565. [Google Scholar]
  106. Park, E.; Yoo, J.; Lee, K. Enhanced Photoelectrochemical Hydrogen Production via Linked BiVO4 Nanoparticles on Anodic WO3 Nanocoral Structures. Sustain. Energy Fuels 2024, 8, 1793. [Google Scholar] [CrossRef]
  107. Kong, D.; Qi, J.; Liu, D.; Zhang, X.; Pan, L.; Zou, J. Ni-Doped BiVO4 with V4+ Species and Oxygen Vacancies for Efficient Photoelectrochemical Water Splitting. Trans. Tianjin Univ. 2019, 25, 340–347. [Google Scholar] [CrossRef]
  108. Arunachalam, M.; Lee, K.; Zhu, K.; Kang, S.H. Effective Corrosion-Resistant Single-Atom Alloy Catalyst on HfO2 -Passivated BiVO4 Photoanode for Durable (≈800 h) Solar Water Oxidation. Adv. Energy Mater. 2024, 14, 2402607. [Google Scholar] [CrossRef]
  109. Yu, S.; Su, C.; Xiao, Z.; Kuang, Y.; Gong, X.; He, X.; Liu, J.; Jin, Q.; Sun, Z. Tuning Surface Hydrophilicity of a BiVO4 Photoanode through Interface Engineering for Efficient PEC Water Splitting. RSC Adv. 2025, 15, 815–823. [Google Scholar] [CrossRef] [PubMed]
  110. Jang, J.W.; Ji, S.G.; Jeong, C.; Kim, J.; Kwon, H.R.; Lee, T.H.; Lee, S.A.; Cheon, W.S.; Lee, S.; Lee, H.; et al. High-efficiency unbiased water splitting with photoanodes harnessing polycarbazole hole transport layers. Energy Environ. Sci. 2024, 17, 2541–2553. [Google Scholar] [CrossRef]
  111. Cui, J.; Daboczi, M.; Regue, M.; Chin, Y.; Pagano, K.; Zhang, J.; Isaacs, M.A.; Kerherve, G.; Mornto, A.; West, J.; et al. 2D Bismuthene as a Functional Interlayer between BiVO4 and NiFeOOH for Enhanced Oxygen-Evolution Photoanodes. Adv. Funct. Mater. 2022, 32, 2270250. [Google Scholar] [CrossRef]
  112. Li, L.; Li, J.; Bai, J.; Zeng, Q.; Xia, L.; Zhang, Y.; Chen, S.; Xu, Q.; Zhou, B. Serial Hole Transfer Layers for a BiVO4 Photoanode with Enhanced Photoelectrochemical Water Splitting. Nanoscale 2018, 10, 18378–18386. [Google Scholar] [CrossRef]
  113. Meng, Q.; Zhang, B.; Yang, H.; Liu, C.; Li, Y.; Kravchenko, A.; Sheng, X.; Fan, L.; Li, F.; Sun, L. Remarkable Synergy of Borate and Interfacial Hole Transporter on BiVO4 Photoanodes for Photoelectrochemical Water Oxidation. Mater. Adv. 2021, 2, 4323–4332. [Google Scholar] [CrossRef]
  114. Yi, Q.; Wang, H.; Lee, J. BiVO4 -Based Photoelectrochemical Water Splitting. ChemElectroChem 2025, 12, e202400600. [Google Scholar] [CrossRef]
  115. Li, T.-B.; Wang, S.-C.; Li, Y.; Zhao, J.; Su, R.; Wang, J.-G.; Chen, G.; Han, M.-J.; Liu, C.-L.; Liu, J.; et al. Cation Doping to Enable Sub-Bandgap Hydrogen Evolution on BiVO4 Photoanodes. ACS Appl. Mater. Interfaces 2022, 14, 35882–35892. [Google Scholar]
  116. Li, J.; Xu, D.; Li, T.; Cao, H.; Yu, K.; Li, H.; Chen, G.; Han, M.; Liu, C.; Wang, J.; et al. Band structure and light absorption engineering of Cs-doped BiVO4 photoanodes for efficient photoelectrochemical water splitting. J. Mater. Chem. A 2021, 9, 674–683. [Google Scholar]
  117. Lin, T.; Zhang, Y.; Luo, J.; Yu, K.; Luo, H.; Zhu, H.; Xue, K.; Li, T.; Wang, X.; Zhang, H.; et al. Band-Bending-Driven Enhanced Charge Transfer in a Step-by-Step BiVO4/WO3 Photoanode for Efficient Photoelectrochemical Water Splitting. Adv. Sci. 2023, 10, 2206729. [Google Scholar]
  118. Zhang, Z.; Liu, H.; Cheng, Z.; Zhou, Y.; Cai, Y.; Xie, L.; Liu, Y.; Cao, H.; Yu, K.; Zhang, H.; et al. BiVO4 photocatalysts co-doped with Cu2+ and Mo6+ via an optimized process for efficient CO2 reduction. Nano Res. 2023, 16, 1943–1953. [Google Scholar]
  119. Shao, Z.; Zhou, Y.; Liu, H.; Cai, Y.; Zhang, Z.; Lu, Z.; Yu, K.; Cao, H.; Hou, Y.; Liu, J. In-Plane Strain Engineering of Carbon Nitride/BiVO4 S-Scheme Heterojunction for Efficient Photocatalytic CO2 Reduction. ACS Mater. Lett. 2024, 6, 180–187. [Google Scholar]
  120. Ren, H.; Zhou, Y.; Liu, H.; Cai, Y.; Zhang, Z.; Lu, Z.; Wang, J.; Yu, K.; Liu, Y.; Cao, H.; et al. Facet engineering of CdS/BiVO4 S-scheme heterojunction for highly efficient photocatalytic CO2 reduction. Nano Energy 2024, 117, 109010. [Google Scholar]
  121. Han, W.; Li, J.; Yu, K.; Li, H.; Cai, Y.; Zhang, Z.; Zhu, H.; Liu, Y.; Cao, H.; Zhang, H.; et al. Band structure and charge transfer of AgI/Ag/Ag2O/BiVO4 Z-scheme heterojunction photoanodes for efficient photoelectrochemical water splitting. New J. Chem. 2022, 46, 10800–10807. [Google Scholar]
  122. Wang, L.; Zhang, T.; Su, J.; Guo, L. Room-Temperature Photodeposition of Conformal Transition Metal Based Cocatalysts on BiVO4 for Enhanced Photoelectrochemical Water Splitting. Nano Res. 2020, 13, 231–237. [Google Scholar] [CrossRef]
  123. LLi, D.; Chen, R.; Wang, P.; Li, Z.; Zhu, J.; Fan, F.; Shi, J.; Li, C. Effect of Facet-Selective Assembly of Cocatalyst on BiVO4 Photoanode for Solar Water Oxidation. ChemCatChem 2019, 11, 3763–3769. [Google Scholar] [CrossRef]
  124. Guo, A.; Wu, X.; Ali, S.H.; Shen, H.; Chen, L.; Li, Y.; Wang, B. Modified Photoanode by in Situ Growth of Covalent Organic Frameworks on BiVO4 for Oxygen Evolution Reaction. RSC Adv. 2024, 14, 9609–9618. [Google Scholar] [CrossRef] [PubMed]
  125. Feng, F.; Mitoraj, D.; Gong, R.; Gao, D.; Elnagar, M.M.; Liu, R.; Beranek, R.; Streb, C. High-Performance BiVO4 Photoanodes: Elucidating the Combined Effects of Mo-Doping and Modification with Cobalt Polyoxometalate. Mater. Adv. 2024, 5, 4932–4944. [Google Scholar] [CrossRef]
  126. Liang, Y.; Tsubota, T.; Mooij, L.P.A.; Van De Krol, R. Highly Improved Quantum Efficiencies for Thin Film BiVO4 Photoanodes. J. Phys. Chem. Commun. 2011, 115, 17594–17598. [Google Scholar] [CrossRef]
  127. Abdi, F.F.; Han, L.; Smets, A.H.M.; Zeman, M.; Dam, B.; Van De Krol, R. Efficient Solar Water Splitting by Enhanced Charge Separation in a Bismuth Vanadate-Silicon Tandem Photoelectrode. Nat. Commun. 2013, 4, 2195. [Google Scholar] [CrossRef]
  128. Su, J.; Guo, L.; Bao, N.; Grimes, C.A. Nanostructured WO3/BiVO4 Heterojunction Films for Efficient Photoelectrochemical Water Splitting. Nano Lett. 2011, 11, 1928–1933. [Google Scholar] [CrossRef]
  129. Zhao, S.; Jia, C.; Shen, X.; Li, R.; Oldham, L.; Moss, B.; Tam, B.; Pike, S.; Harrison, N.; Ahmad, E.; et al. The Aerosol-Assisted Chemical Vapour Deposition of Mo-Doped BiVO4 Photoanodes for Solar Water Splitting: An Experimental and Computational Study. J. Mater. Chem. A 2024, 12, 26645–26666. [Google Scholar] [CrossRef]
  130. Cheng, S.; Cheng, Y.; Zhou, T.; Li, S.; Xie, D.; Li, X. Enhanced Photoelectrochemical Performance of BiVO4 Photoanodes Co-Modified with Borate and NiFeOx. Micromachines 2025, 16, 866. [Google Scholar] [CrossRef]
  131. Fu, C.; Xu, B.; Dong, L.; Zhai, J.; Wang, X.; Wang, D.-Y. Highly Efficient BiVO4 Single-Crystal Nanosheets with Dual Modification: Phosphorus Doping and Selective Ag Modification. Nanotechnology 2021, 32, 325701. [Google Scholar] [CrossRef]
  132. Gil-Rostra, J.; Castillo-Seoane, J.; Guo, Q.; Sobrido, A.J.; González-Elipe, A.R.; Borrás, A. Photoelectrochemical Water Splitting with ITO/WO3/BiVO4/CoPi Multishell Nanotubes Enabled by a Vacuum and Plasma Soft-Template Synthesis. ACS Appl. Mater. Interfaces 2023, 15, 9250–9262. [Google Scholar] [CrossRef]
  133. Samuel, E.; Joshi, B.; Kim, M.-W.; Swihart, M.T.; Yoon, S.S. Morphology Engineering of Photoelectrodes for Efficient Photoelectrochemical Water Splitting. Nano Energy 2020, 72, 104648. [Google Scholar] [CrossRef]
  134. Wang, T.; Yu, Q.; Zhang, S.; Kou, X.; Sun, P.; Lu, G. Rational Design of 3D Inverse Opal Heterogeneous Composite Microspheres as Excellent Visible-Light-Induced NO2 Sensors at Room Temperature. Nanoscale 2018, 10, 4841–4851. [Google Scholar] [CrossRef] [PubMed]
  135. Zhang, H.; Cheng, C. Three-Dimensional FTO/TiO2/BiVO4 Composite Inverse Opals Photoanode with Excellent Photoelectrochemical Performance. ACS Energy Lett. 2017, 2, 813–821. [Google Scholar] [CrossRef]
  136. Marinković, D.; Righini, G.C.; Ferrari, M. Synthesis, Optical, and Photocatalytic Properties of the BiVO4 Semiconductor Nanoparticles with Tetragonal Zircon-Type Structure. Photonics 2025, 12, 438. [Google Scholar] [CrossRef]
  137. Zhao, H.; Wei, X.; Pei, Y.; Han, W. Enhancing Photoelectrocatalytic Efficiency of BiVO4 Photoanodes by Crystal Orientation Control. Nanomaterials 2024, 14, 1870. [Google Scholar] [CrossRef]
  138. Hu, Z.; Tong, G.; Lin, D.; Nian, Q.; Shao, J.; Hu, Y.; Saeib, M.; Jin, S.; Cheng, G.J. Laser Sintered Graphene Nickel Nanocomposites. J. Mater. Process. Technol. 2016, 231, 143–150. [Google Scholar] [CrossRef]
  139. Feng, W.; Wang, Y.; Huang, X.; Wang, K.; Gao, F.; Zhao, Y.; Wang, B.; Zhang, L.; Liu, P. One-Pot Construction of 1D/2D Zn1-xCdx S/D-ZnS(en)0.5 Composites with Perfect Heterojunctions and Their Superior Visible-Light-Driven Photocatalytic H2 Evolution. Appl. Catal. B Environ. 2018, 220, 324–336. [Google Scholar] [CrossRef]
  140. Prasad, U.; Young, J.L.; Johnson, J.C.; McGott, D.L.; Gu, H.; Garfunkel, E.; Kannan, A.M. Enhancing Interfacial Charge Transfer in a WO3/BiVO4 Photoanode Heterojunction through Gallium and Tungsten Co-Doping and a Sulfur Modified Bi2O3 Interfacial Layer. J. Mater. Chem. A 2021, 9, 16137–16149. [Google Scholar] [CrossRef]
  141. Nong, S.; Wang, M.; Wang, X.; Li, Y.; Yu, S.; Tang, C.; Li, G.; Xu, L. A Multifunctional Guanosine-Based Carbon Dots for Dead Microbial Imaging and Synergistic Broad-Spectrum Antimicrobial Therapy. Chem. Eng. J. 2024, 485, 150123. [Google Scholar] [CrossRef]
  142. Wang, A.; Chen, Y.; Liu, X.; Li, R.; Zhang, Z.; Zhang, F.; Cao, D.; Gao, Z.; Mi, B. Plasma treatment induced oxygen vacancies and N doping in the BiVO4 photoanode toward improved charge transfer and facilitated water oxidation. ACS Appl. Mater. Interfaces 2025, 17, 12064–12073. [Google Scholar] [CrossRef]
  143. Qian, F.; Tian, J.; Bai, Z.; Liu, L.; Cao, M.; Li, J.; Chen, S. CoNi2S4 Cocatalyst Modified TiO2 Nanorods/Nanoflower Homojunction for Efficient Photoelectrochemical Water Splitting. J. Alloys Compd. 2024, 991, 174567. [Google Scholar] [CrossRef]
  144. Chen, M.; Chang, X.; Li, C.; Wang, H.; Jia, L. Ni-Doped BiVO4 Photoanode for Efficient Photoelectrochemical Water Splitting. J. Colloid Interface Sci. 2023, 640, 162–169. [Google Scholar] [CrossRef] [PubMed]
  145. Yang, J.W.; Park, I.J.; Lee, S.A.; Lee, M.G.; Lee, T.H.; Park, H.; Kim, C.; Park, J.; Moon, J.; Kim, J.Y.; et al. Near-Complete Charge Separation in Tailored BiVO4-Based Heterostructure Photoanodes toward Artificial Leaf. Appl. Catal. B Environ. 2021, 293, 120217. [Google Scholar] [CrossRef]
  146. Liu, B.; Wang, X.; Zhang, Y.; Xu, L.; Wang, T.; Xiao, X.; Wang, S.; Wang, L.; Huang, W. A BiVO4 photoanode with a VOx layer bearing oxygen vacancies offers improved charge transfer and oxygen evolution kinetics in photoelectrochemical water splitting. Angew. Chem. Int. Ed. 2023, 62, e202217346. [Google Scholar] [CrossRef] [PubMed]
  147. Bhat, S.S.M.; Lee, S.A.; Suh, J.M.; Hong, S.-P.; Jang, H.W. Triple Planar Heterojunction of SnO2/WO3/BiVO4 with Enhanced Photoelectrochemical Performance under Front Illumination. Appl. Sci. 2018, 8, 1765. [Google Scholar] [CrossRef]
  148. Fan, X.; Chen, Q.; Zhu, F.; Wang, T.; Gao, B.; Song, L.; He, J. Preparation of Surface Dispersed WO3/BiVO4 Heterojunction Arrays and Their Photoelectrochemical Performance for Water Splitting. Molecules 2024, 29, 372. [Google Scholar] [CrossRef]
  149. Cao, Y.; Xing, Z.; Wang, B.; Tang, W.; Wu, R.; Li, J.; Ma, M. Surface Engineering of WO3/BiVO4 to Boost Solar Water-Splitting. Catalysts 2020, 10, 556. [Google Scholar] [CrossRef]
  150. Ziwritsch, M.; Müller, S.; Hempel, H.; Unold, T.; Abdi, F.F.; Van De Krol, R.; Friedrich, D.; Eichberger, R. Direct Time-Resolved Observation of Carrier Trapping and Polaron Conductivity in BiVO4. ACS Energy Lett. 2016, 1, 888–894. [Google Scholar] [CrossRef]
  151. Huang, Y.; Guo, Z.; Liu, H.; Zhang, S.; Wang, P.; Lu, J.; Tong, Y. Heterojunction Architecture of N-Doped WO3 Nanobundles with Ce2S3 Nanodots Hybridized on a Carbon Textile Enables a Highly Efficient Flexible Photocatalyst. Adv Funct Mater. 2019, 29, 1903490. [Google Scholar] [CrossRef]
  152. Nair, V.; Perkins, C.L.; Lin, Q.; Law, M. Textured Nanoporous Mo:BiVO4 Photoanodes with High Charge Transport and Charge Transfer Quantum Efficiencies for Oxygen Evolution. Energy Environ. Sci. 2016, 9, 1412–1429. [Google Scholar] [CrossRef]
  153. Rettie, A.J.E.; Lee, H.C.; Marshall, L.G.; Lin, J.-F.; Capan, C.; Lindemuth, J.; McCloy, J.S.; Zhou, J.; Bard, A.J.; Mullins, C.B. Combined Charge Carrier Transport and Photoelectrochemical Characterization of BiVO4 Single Crystals: Intrinsic Behavior of a Complex Metal Oxide. J. Am. Chem. Soc. 2013, 135, 11389–11396. [Google Scholar] [CrossRef] [PubMed]
  154. Wu, X.; Oropeza, F.E.; Qi, Z.; Einert, M.; Tian, C.; Maheu, C.; Lv, K.; Hofmann, J.P. Influence of Mo Doping on Interfacial Charge Carrier Dynamics in Photoelectrochemical Water Oxidation on BiVO4. Sustain. Energy Fuels 2023, 7, 2923–2933. [Google Scholar] [CrossRef]
  155. Chaudhari, S.A.; Patil, V.V.; Jadhav, V.A.; Thorat, P.; Sutar, S.S.; Dongale, T.D.; Parale, V.; Patil, V.; Mhamane, D.S.; Mali, M.G.; et al. Conductivity Boosted BiVO4 for Enhanced OER and Supercapacitive Performance: Stability Insights with Modeling, Predictions, and Forecasting Using Machine Learning Technique. Energy Mater. 2025, 5, 500082. [Google Scholar] [CrossRef]
  156. Wang, Y.; Zhang, J.; Balogun, M.-S.; Tong, Y.; Huang, Y. Oxygen Vacancy–Based Metal Oxides Photoanodes in Photoelectrochemical Water Splitting. Mater. Today Sustain. 2022, 18, 100118. [Google Scholar] [CrossRef]
  157. Li, R.; Zhan, F.; Wen, G.; Wang, B.; Qi, J.; Liu, Y.; Feng, C.; La, P. Facile Synthesis of a Micro–Nano-Structured FeOOH/BiVO4/WO3 Photoanode with Enhanced Photoelectrochemical Performance. Catalysts 2024, 14, 828. [Google Scholar] [CrossRef]
  158. Creasey, G.H.; McCallum, T.W.; Ai, G.; Tam, B.; Rodriguez Acosta, J.W.; Mohammad Yousuf, A.; Fearn, S.; Eisner, F.; Kafizas, A.; Hankin, A. Mechanically and Photoelectrochemically Stable WO3|BiVO4|NiFeOOH Photoanodes Synthesised by a Scalable Chemical Vapour Deposition Method. J. Mater. Chem. A 2025, 13, 11585–11604. [Google Scholar] [CrossRef]
  159. Li, Z.; Xie, Z.; Li, W.; Aziz, H.S.; Abbas, M.; Zheng, Z.; Su, Z.; Fan, P.; Chen, S.; Liang, G. Charge Transport Enhancement in BiVO4 Photoanode for Efficient Solar Water Oxidation. Materials 2023, 16, 3414. [Google Scholar] [CrossRef]
  160. Zhang, S.; Ahmet, I.; Kim, S.-H.; Kasian, O.; Mingers, A.M.; Schnell, P.; Kolbach, M.; Lim, J.; Fischer, A.; Mayrhofer, K.J.J.; et al. Different Photostability of BiVO4 in Near-pH-Neutral Electrolytes. ACS Appl. Energy Mater. 2020, 3, 9523–9527. [Google Scholar]
  161. Wang, L.; Zhang, Y.; Li, W.; Wang, L. Recent Advances in Elaborate Interface Regulation of BiVO4 Photoanode for Photoelectrochemical Water Splitting. Mater. Rep. Energy 2023, 3, 100232. [Google Scholar] [CrossRef]
  162. Lei, R.; Tang, Y.; Qiu, W.; Yan, S.; Tian, X.; Wang, Q.; Chen, Q.; Wang, Z.; Qian, W.; Xu, Q.; et al. Prompt Hole Extraction Suppresses V5+ Dissolution and Sustains Large-Area BiVO4 Photoanodes for Over 2100 h Water Oxidation. Nano Lett. 2023, 23, 11785–11792. [Google Scholar] [CrossRef]
  163. Lee, K.W.; Kim, D.S.; Jung, S.Y.; Kang, J.; Cho, H.K. Oxygen Network Formation for Efficient Charge Transport and Durable BiVO4 Photoanodes with Ultra-Thin TiO2 Layer in Solar Water Splitting. Appl. Surf. Sci. Adv. 2025, 27, 100737. [Google Scholar] [CrossRef]
  164. Zhou, G.; Aletsee, C.C.; Lemperle, A.; Rieth, T.; Mengel, L.; Gao, J.; Tschurl, M.; Heiz, U.; Sharp, I.D. Water Oxidation and Degradation Mechanisms of BiVO4 Photoanodes in Bicarbonate Electrolytes. ACS Catal. 2025, 15, 13048–13058. [Google Scholar] [CrossRef] [PubMed]
  165. Cao, X.; Yu, X.; Chen, X.; Ye, R. Boosting the Photoelectrochemical Performance of BiVO4 by Borate Buffer Activation: The Role of Trace Iron Impurities. Chem. Commun. 2024, 60, 10330–10333. [Google Scholar] [CrossRef] [PubMed]
  166. Gong, H.; An, S.; Qin, W.; Kuang, Y.; Liu, D. Stabilizing BiVO4 Photoanode in Bicarbonate Electrolyte for Efficient Photoelectrocatalytic Alcohol Oxidation. Molecules 2024, 29, 1554. [Google Scholar] [CrossRef]
  167. Khan, H.; Kim, M.-J.; Balasubramanian, P.; Jung, M.-J.; Kwon, S.-H. Improving the Photoelectrochemical Stability of a Bismuth Vanadate Photoanode for Solar Water Oxidation. Cell Rep. Phys. Sci. 2023, 4, 101652. [Google Scholar] [CrossRef]
  168. Jung, G.; Moon, C.; Martinho, F.; Jung, Y.; Chu, J.; Park, H.; Hajijafarassar, A.; Nielsen, R.; Schou, J.; Park, J.; et al. Monolithically Integrated BiVO4/Si Tandem Devices Enabling Unbiased Photoelectrochemical Water Splitting. Adv. Energy Mater. 2023, 13, 2301235. [Google Scholar] [CrossRef]
  169. Sitaaraman, S.R.; Grace, A.N.; Zhu, J.; Sellappan, R. Photoelectrochemical Performance of a Nanostructured BiVO4/NiOOH/FeOOH–Cu2O/CuO/TiO2 Tandem Cell for Unassisted Solar Water Splitting. Nanoscale Adv. 2024, 6, 2407–2418. [Google Scholar] [CrossRef]
  170. Liu, B.; Wang, X.; Zhang, Y.; Zhu, M.; Zhang, C.; Li, S.; Ma, Y.; Huang, W.; Wang, S. A Standalone Bismuth Vanadate-Silicon Artificial Leaf Achieving 8.4% Efficiency for Hydrogen Production. Nat. Commun. 2025, 16, 2792. [Google Scholar] [CrossRef]
  171. Shu, X.; Wang, R.; Yang, L.; Liu, S.; Yin, Y.; Liang, X.; Hu, K.; Zhang, M. Bifunctional BiVO4/Cu2O Photoelectrode for Bias-Free Solar Water Splitting Tandem Cells. Sol. Energy Mater. Sol. Cells 2025, 282, 113386. [Google Scholar] [CrossRef]
  172. Ahmet, I.Y.; Ma, Y.; Jang, J.-W.; Henschel, T.; Stannowski, B.; Lopes, T.; Vilanova, A.; Mendes, A.; Abdi, F.F.; Van De Krol, R. Demonstration of a 50 Cm2 BiVO4 Tandem Photoelectrochemical-Photovoltaic Water Splitting Device. Sustain. Energy Fuels 2019, 3, 2366–2379. [Google Scholar] [CrossRef]
  173. Wan, S.; Jin, J.; Dong, C.; Lu, Y.; Zhong, Q.; Zhang, K.; Park, J.H. Promoting Water Oxidative Hydrogen Peroxide Production by Dynamic Anion Exchange on BiVO4 Photoanodes. ACS Catal. 2023, 13, 14845–14852. [Google Scholar] [CrossRef]
  174. Shi, S.; Song, Y.; Jiao, Y.; Jin, D.; Li, Z.; Xie, H.; Gao, L.; Sun, L.; Hou, J. BiVO4-Based Heterojunction Photocathode for High-Performance Photoelectrochemical Hydrogen Peroxide Production. Nano Lett. 2024, 24, 6051–6060. [Google Scholar] [CrossRef]
Figure 8. (a) UV–Vis absorption spectra of Mo-BiVO4/CoPOM, Mo-BiVO4 and pristine BiVO4 photoanodes with (b) Tauc plots for bandgap determination. Reprinted with permission from Ref. [125], Copyright © 2024, Royal Society of Chemistry. (c) Projected density of states (PDOS) comparison for undoped t-BiVO4 (left) and m-BiVO4 (right) host structures. Reprinted with permission from Ref. [129], Copyright © 2024, Royal Society of Chemistry. (d) UV–Vis absorption spectra of BiVO4, B/BiVO4, and B/BiVO4/NiFeOx photoanodes and (e) corresponding Tauc plots. Reprinted with permission from Ref. [130], Copyright © 2025, MDPI.
Figure 8. (a) UV–Vis absorption spectra of Mo-BiVO4/CoPOM, Mo-BiVO4 and pristine BiVO4 photoanodes with (b) Tauc plots for bandgap determination. Reprinted with permission from Ref. [125], Copyright © 2024, Royal Society of Chemistry. (c) Projected density of states (PDOS) comparison for undoped t-BiVO4 (left) and m-BiVO4 (right) host structures. Reprinted with permission from Ref. [129], Copyright © 2024, Royal Society of Chemistry. (d) UV–Vis absorption spectra of BiVO4, B/BiVO4, and B/BiVO4/NiFeOx photoanodes and (e) corresponding Tauc plots. Reprinted with permission from Ref. [130], Copyright © 2025, MDPI.
Nanomaterials 15 01494 g008
Figure 9. (a) JV curves and (b) chronoamperometric curves (at 1.23 VRHE) of bare BiVO4, BiVO4/PbS(n) QDs, and BiVO4/PbS(5) QDs/ZnS photoanodes (n: the number of PbS SILAR cycles). Reprinted with permission from Ref. [56], Copyright © 2023, MDPI. (c) Scheme of synthetic route of dually modified BiVO4 photocatalyst. Reproduced from Ref. [131], under Creative Commons CC BY license. (d) LSV diagrams under illumination with blue light for NT_#1/CoPi, NT_#2/CoPi and NT_#3/CoPi electrodes. (e) Band diagram describing the fate of photoelectron and photohole upon light absorption by the BiVO4 semiconductor. Reproduced from Ref. [132], under Creative Commons CC BY license.
Figure 9. (a) JV curves and (b) chronoamperometric curves (at 1.23 VRHE) of bare BiVO4, BiVO4/PbS(n) QDs, and BiVO4/PbS(5) QDs/ZnS photoanodes (n: the number of PbS SILAR cycles). Reprinted with permission from Ref. [56], Copyright © 2023, MDPI. (c) Scheme of synthetic route of dually modified BiVO4 photocatalyst. Reproduced from Ref. [131], under Creative Commons CC BY license. (d) LSV diagrams under illumination with blue light for NT_#1/CoPi, NT_#2/CoPi and NT_#3/CoPi electrodes. (e) Band diagram describing the fate of photoelectron and photohole upon light absorption by the BiVO4 semiconductor. Reproduced from Ref. [132], under Creative Commons CC BY license.
Nanomaterials 15 01494 g009
Figure 10. (a) Photoluminescent emission spectra of tz-BiVO4 nanoparticles in ethylene glycol, (b) photodegradation curves of MO solution (5 mg/L) by different tz-BiVO4 samples and Degussa P25 (1 g/L) under UV-Vis lighting of MO in the presence of the tz-BiVO4 photocatalyst. Reprinted with permission from Ref. [136], Copyright © 2025, MDPI. (c) Schematic illustration of the synthesis and characterizations of BiVO4 catalytic films. Reprinted with permission from Ref. [137], Copyright © 2024, MDPI.
Figure 10. (a) Photoluminescent emission spectra of tz-BiVO4 nanoparticles in ethylene glycol, (b) photodegradation curves of MO solution (5 mg/L) by different tz-BiVO4 samples and Degussa P25 (1 g/L) under UV-Vis lighting of MO in the presence of the tz-BiVO4 photocatalyst. Reprinted with permission from Ref. [136], Copyright © 2025, MDPI. (c) Schematic illustration of the synthesis and characterizations of BiVO4 catalytic films. Reprinted with permission from Ref. [137], Copyright © 2024, MDPI.
Nanomaterials 15 01494 g010
Figure 11. (a) LSV of SnO2/WO3 photoanode measured using a three-electrode configuration set up in aqueous phosphate buffer (pH 7.0) with 0.5 M Na2SO3, (b) EIS of SnO2/WO3/BiVO4. Reprinted with permission from Ref. [147], Copyright © 2028, MDPI. (c) Schematic illustration of the formation and the PEC water oxidation mechanism of the WO3/BiVO4 arrays. The red flakes represent WO3 and the green particles represent BiVO4 (d) LSV curves of the photoelectrode. Reprinted with permission from Ref. [148], Copyright © 2024, MDPI. (e) IV curves for FTO, WB-7 cycle, WB-30 s, WB-5 min, and WB-10 min samples. Reprinted with permission from Ref. [149], Copyright © 2020, MDPI.
Figure 11. (a) LSV of SnO2/WO3 photoanode measured using a three-electrode configuration set up in aqueous phosphate buffer (pH 7.0) with 0.5 M Na2SO3, (b) EIS of SnO2/WO3/BiVO4. Reprinted with permission from Ref. [147], Copyright © 2028, MDPI. (c) Schematic illustration of the formation and the PEC water oxidation mechanism of the WO3/BiVO4 arrays. The red flakes represent WO3 and the green particles represent BiVO4 (d) LSV curves of the photoelectrode. Reprinted with permission from Ref. [148], Copyright © 2024, MDPI. (e) IV curves for FTO, WB-7 cycle, WB-30 s, WB-5 min, and WB-10 min samples. Reprinted with permission from Ref. [149], Copyright © 2020, MDPI.
Nanomaterials 15 01494 g011
Figure 12. (a) LSV curves of as-prepared BiVO4-based samples measured both in the dark and under illumination (λexc = 435 nm, 100 mW/cm2) in 0.1 M phosphate buffer and 0.5 M Na2SO3 solutions, showing enhanced photocurrent under illumination. (b,c) Mott–Schottky plots of pristine BVO and Mo-doped BVO, respectively, highlighting changes in donor density and flat-band potential upon Mo incorporation. Reprinted with permission from Ref. [154], Copyright © 2023, Royal Society of Chemistry. (d) LSV curves comparing BiVO4 and BiVO4/carbon composite electrodes under PEC conditions, (e) LSV curves for OER measured before and after stability testing at a scan rate of 1 mV/s, demonstrating the effect of stability treatment, and (f) EIS Nyquist plots of the electrodes before and after CA stability testing, revealing changes in charge transfer resistance. Reprinted with permission from Ref. [155], Copyright © 2025, OAE Publishing Inc.
Figure 12. (a) LSV curves of as-prepared BiVO4-based samples measured both in the dark and under illumination (λexc = 435 nm, 100 mW/cm2) in 0.1 M phosphate buffer and 0.5 M Na2SO3 solutions, showing enhanced photocurrent under illumination. (b,c) Mott–Schottky plots of pristine BVO and Mo-doped BVO, respectively, highlighting changes in donor density and flat-band potential upon Mo incorporation. Reprinted with permission from Ref. [154], Copyright © 2023, Royal Society of Chemistry. (d) LSV curves comparing BiVO4 and BiVO4/carbon composite electrodes under PEC conditions, (e) LSV curves for OER measured before and after stability testing at a scan rate of 1 mV/s, demonstrating the effect of stability treatment, and (f) EIS Nyquist plots of the electrodes before and after CA stability testing, revealing changes in charge transfer resistance. Reprinted with permission from Ref. [155], Copyright © 2025, OAE Publishing Inc.
Nanomaterials 15 01494 g012
Figure 13. (a) SEM surface morphology images of BiVO4/WO3 photoanodes, (b) surface photovoltage efficiency (ηsurface) curves, and (c) dark current curves of WO3, BiVO4/WO3, and FeOOH/BiVO4/WO3 electrodes. Reprinted with permission from Ref. [157], Copyright © 2024, MDPI. (d) Chronoamperometric stability tests of BiVO4 and WO3|BiVO4 electrodes at 1.23 VRHE in 0.1 M KPi (pH 7) and photograph showing visible degradation on the irradiated area of WO3|BiVO4 after 24 h testing, (e) schematic optical pathway of the PEC cell used for preliminary measurements of WO3|BiVO4 electrodes. Reprinted with permission from Ref. [158], Copyright © 2025, Royal Society of Chemistry. (f,g) JV curves of BiVO4 and BiVO4/CoPi photoanodes under dark conditions and simulated sunlight illumination, respectively, and (h) corresponding ABPE values. Reprinted with permission from Ref. [159], Copyright © 2023, MDPI.
Figure 13. (a) SEM surface morphology images of BiVO4/WO3 photoanodes, (b) surface photovoltage efficiency (ηsurface) curves, and (c) dark current curves of WO3, BiVO4/WO3, and FeOOH/BiVO4/WO3 electrodes. Reprinted with permission from Ref. [157], Copyright © 2024, MDPI. (d) Chronoamperometric stability tests of BiVO4 and WO3|BiVO4 electrodes at 1.23 VRHE in 0.1 M KPi (pH 7) and photograph showing visible degradation on the irradiated area of WO3|BiVO4 after 24 h testing, (e) schematic optical pathway of the PEC cell used for preliminary measurements of WO3|BiVO4 electrodes. Reprinted with permission from Ref. [158], Copyright © 2025, Royal Society of Chemistry. (f,g) JV curves of BiVO4 and BiVO4/CoPi photoanodes under dark conditions and simulated sunlight illumination, respectively, and (h) corresponding ABPE values. Reprinted with permission from Ref. [159], Copyright © 2023, MDPI.
Nanomaterials 15 01494 g013
Figure 14. (a) Time profiles of photocurrent density i (black), film thickness h (red) the mark * indicates the bubble removal points, (b) dissolution rates of BiVO4 held at 1.6 VRHE in borate, (c,e) top view of scanning electron microscopy (SEM), (d,f) cross-section view of scanning transmission electron microscopy (STEM). Reproduced from Ref. [160], under Creative Commons CC BY license. (g) Schematic diagram of the water oxidation of NiFe-MOFs/BiVO4. (h) Stability of Co(OH)x/p-Cu2S/n-BiVO4 series photoanode in 0.5 M KPi. (i) Schematic illustration of the restraint of photocorrosion in KBi–V. Reprinted with permission from Ref. [161], Copyright © 2023, Wiley-VCH GmbH.
Figure 14. (a) Time profiles of photocurrent density i (black), film thickness h (red) the mark * indicates the bubble removal points, (b) dissolution rates of BiVO4 held at 1.6 VRHE in borate, (c,e) top view of scanning electron microscopy (SEM), (d,f) cross-section view of scanning transmission electron microscopy (STEM). Reproduced from Ref. [160], under Creative Commons CC BY license. (g) Schematic diagram of the water oxidation of NiFe-MOFs/BiVO4. (h) Stability of Co(OH)x/p-Cu2S/n-BiVO4 series photoanode in 0.5 M KPi. (i) Schematic illustration of the restraint of photocorrosion in KBi–V. Reprinted with permission from Ref. [161], Copyright © 2023, Wiley-VCH GmbH.
Nanomaterials 15 01494 g014
Figure 15. (a) Jt curve of the NiOx/BiVO4 at 1.23 VRHE. The electrolyte was replaced at 108 and 208 h, respectively. (b) Schematic illustration of the proposed anticorrosion mechanism of NiOx/BiVO4. Reprinted with permission from Ref. [162], Copyright © 2023, American Chemical Society. (c) PEC performances of P-BiVO4 (black line) and P-BiVO4/TiO2 (sky-blue line) photoanodes in general potassium borate electrolyte (dashed lines) and in PB electrolyte with added scavenger (solid lines), under AM 1.5G illumination; (d) PEC performances of P-BiVO4 (black line) and P-BiVO4/TiO2 (sky-blue line) photoanodes in general potassium borate (PB) electrolyte (dashed lines) and in PB electrolyte with added scavenger (solid lines). (e) Schematic for charge kinetics at the bulk, surface, and interface of the BiVO4 with optimized overlayer. Reprinted with permission from Ref. [163], Copyright © 2025, Wiley-VCH GmbH.
Figure 15. (a) Jt curve of the NiOx/BiVO4 at 1.23 VRHE. The electrolyte was replaced at 108 and 208 h, respectively. (b) Schematic illustration of the proposed anticorrosion mechanism of NiOx/BiVO4. Reprinted with permission from Ref. [162], Copyright © 2023, American Chemical Society. (c) PEC performances of P-BiVO4 (black line) and P-BiVO4/TiO2 (sky-blue line) photoanodes in general potassium borate electrolyte (dashed lines) and in PB electrolyte with added scavenger (solid lines), under AM 1.5G illumination; (d) PEC performances of P-BiVO4 (black line) and P-BiVO4/TiO2 (sky-blue line) photoanodes in general potassium borate (PB) electrolyte (dashed lines) and in PB electrolyte with added scavenger (solid lines). (e) Schematic for charge kinetics at the bulk, surface, and interface of the BiVO4 with optimized overlayer. Reprinted with permission from Ref. [163], Copyright © 2025, Wiley-VCH GmbH.
Nanomaterials 15 01494 g015
Figure 16. (a) Schematic illustration of water splitting cell configuration in KHCO3 electrolyte. Reprinted with permission from Ref. [164], Copyright © 2025, American Chemical Society. (b) Transient photocurrent response of photoanodes in KPi buffer, unbuffered KHCO3, and CO2-saturated KHCO3 electrolyte under chopped illumination. Reprinted with permission from Ref. [164], Copyright © 2025, American Chemical Society. (c) Jt curves of BiVO4 photoanodes at 0.6 VRHE before and after Fe impurity removal, with photocurrent change upon Fe2+ addition. Reprinted with permission from Ref. [165], Copyright © 2024, Wiley-VCH GmbH. (d) Stability of BVO photoanodes in aqueous and MeCN buffers containing 0.5 M NaHCO3 (pH 9) over 6 h. Reprinted with permission from Ref. [166], Copyright © 2024, MDPI.
Figure 16. (a) Schematic illustration of water splitting cell configuration in KHCO3 electrolyte. Reprinted with permission from Ref. [164], Copyright © 2025, American Chemical Society. (b) Transient photocurrent response of photoanodes in KPi buffer, unbuffered KHCO3, and CO2-saturated KHCO3 electrolyte under chopped illumination. Reprinted with permission from Ref. [164], Copyright © 2025, American Chemical Society. (c) Jt curves of BiVO4 photoanodes at 0.6 VRHE before and after Fe impurity removal, with photocurrent change upon Fe2+ addition. Reprinted with permission from Ref. [165], Copyright © 2024, Wiley-VCH GmbH. (d) Stability of BVO photoanodes in aqueous and MeCN buffers containing 0.5 M NaHCO3 (pH 9) over 6 h. Reprinted with permission from Ref. [166], Copyright © 2024, MDPI.
Nanomaterials 15 01494 g016
Figure 17. (a) Schematic illustration of the optical absorption mechanism and electron transport of nanoporous BiVO4 on the flat substrate and the conductive nanocone substrate. (b) J–V curves of the Mo:BiVO4 on the FTO-coated glass and the nanocone substrate tested in a 0.5 M KH2PO4 buffer solution (pH 7). Reprinted with permission from Ref. [33], Copyright © 2016, American Association for the Advancement of Science. (c) 2-electrode LSV response of BiVO4-Cu2O, Mo-BiVO4/FeOOH-Cu2O/MoS2 and Mo-BiVO4/TiO2/FeOOH-Cu2O/TiO2/MoS2 tandem cells and (d) unassisted stability test (j–t); (e) energy band diagram of BiVO4–Cu2O tandem PEC cell with respect to RHE potential. Reprinted with permission from Ref. [53], Copyright © 2022, Royal Society of Chemistry.
Figure 17. (a) Schematic illustration of the optical absorption mechanism and electron transport of nanoporous BiVO4 on the flat substrate and the conductive nanocone substrate. (b) J–V curves of the Mo:BiVO4 on the FTO-coated glass and the nanocone substrate tested in a 0.5 M KH2PO4 buffer solution (pH 7). Reprinted with permission from Ref. [33], Copyright © 2016, American Association for the Advancement of Science. (c) 2-electrode LSV response of BiVO4-Cu2O, Mo-BiVO4/FeOOH-Cu2O/MoS2 and Mo-BiVO4/TiO2/FeOOH-Cu2O/TiO2/MoS2 tandem cells and (d) unassisted stability test (j–t); (e) energy band diagram of BiVO4–Cu2O tandem PEC cell with respect to RHE potential. Reprinted with permission from Ref. [53], Copyright © 2022, Royal Society of Chemistry.
Nanomaterials 15 01494 g017
Figure 18. (a) LSV curves of BVO and PBVO-2 photoanodes in 1 M NaHCO3 electrolyte under AM 1.5G illumination. (b) Plots of the theoretical charge number obtained from the Jt curves and the actual quantities of H2 and H2O2 Reprinted with permission from Ref. [173], Copyright © 2023, American Chemical Society. (c) Chopped JV curves, (d) It curves under air and oxygen conditions of photoanodes. (e) Comparison of Faradaic efficiency of H2O2 generation by different photocathodes. Reprinted with permission from Ref. [174], Copyright © 2024, American Chemical Society.
Figure 18. (a) LSV curves of BVO and PBVO-2 photoanodes in 1 M NaHCO3 electrolyte under AM 1.5G illumination. (b) Plots of the theoretical charge number obtained from the Jt curves and the actual quantities of H2 and H2O2 Reprinted with permission from Ref. [173], Copyright © 2023, American Chemical Society. (c) Chopped JV curves, (d) It curves under air and oxygen conditions of photoanodes. (e) Comparison of Faradaic efficiency of H2O2 generation by different photocathodes. Reprinted with permission from Ref. [174], Copyright © 2024, American Chemical Society.
Nanomaterials 15 01494 g018
Table 1. Summary of materials, fabrication method, electrolyte condition, and PEC performance of BiVO4-based photoanodes.
Table 1. Summary of materials, fabrication method, electrolyte condition, and PEC performance of BiVO4-based photoanodes.
MaterialsFabrication MethodAnnealing Condition
(Temp., Time,
Atmosphere)
Electrolyte
Condition
Photocurrent Density @ 1.23 VRHE (mA/cm2)
Bare BiVO4 [47]Electrodeposition BiOI + VO(acac)2 drop450 °C, 2 h, air0.1 M Na2SO4
(pH 6)
0.65
NiOOH/FeOOH/BiVO4 [48]Electrodeposition BiOI + VO(acac)2 drop + cocatalysts deposition450 °C, 2 h, air0.1 M Na2SO4~1.3
CdS/NiFe-LDH/BiVO4 [49]BiOI + VO(acac)2 + CdS hydrothermal + NiFe-LDH deposition450 °C, 2 h, air0.5 M Na2SO43.10
BiVO4/NiO composite [50]BiOI + VO(acac)2 conversion + NiO HT450 °C, 2 h, air0.5 M Na2SO41.20
CoV-LDH/Ag/BiVO4 [51]BiOI + VO(acac)2→CoV-LDH/Ag coating450 °C, 2 h, air0.5 M Na2SO4 + 0.1 M glycerol7.15
Bare BiVO4 [52]BiOI electrodeposition + V2O5 electrodeposition + calcination475 °C, 2 h, air0.5 M potassium borate (pH 9.5)2.2
Mo-BiVO4/TiO2/FeOOH [53]BiOI + VO/(Mo dopant) + TiO2 + FeOOH450 °C, 2 h, air0.1 M Na2SO40.81
BiVO4/PbS QDs/ZnS [54]BiOI + VO(acac)2→BiVO4 + PbS/ZnS (SILAR)450 °C, 2 h, air0.5 M KH2PO4 + 1.0 M Na2SO35.19
Table 2. Summary of modification strategies for enhancing the PEC performance of BiVO4-based photoanodes.
Table 2. Summary of modification strategies for enhancing the PEC performance of BiVO4-based photoanodes.
TechniquesMaterials/Structures/MethodsFunctionality
1. Synthesis and Film Morphology Engineering
  • Electrodeposition [84]
  • Hydrothermal synthesis [85]
  • Hollow nanospheres [86]
  • Metal–organic decomposition [86]
  • RF sputtering [87]
(1) Controls morphology and film thickness.
(2) Increases surface area.
(3) Improves light absorption and bulk charge transport (ηbulk).
(4) Precisely controls composition and film density.
(5) Enables conformal coating on complex nanostructures.
2. Doping
  • Mo, W [88].
  • Mo and W [89].
  • Ti, Zr [90]
(1) Increases bulk conductivity.
(2) Tunes the bandgap.
(3) Reduces bulk recombination.
3. Surface and Cocatalyst Modification
  • FeOOH, NiOOH [20]
  • NiOOH thin layer [91]
  • RuO2, IrO2 [92]
  • CoPi [93]
(1) Lowers OER overpotential.
(2) Improves surface reaction kinetics (ηsurf).
(3) Increases chemical stability.
(4) Accelerates oxygen evolution reaction.
(5) Enhances charge transfer efficiency at the surface.
4. Heterojunction Engineering
  • WO3 [94]
  • Cu2O [95]
  • BiVO4/MoS2 [96]
  • BiVO4/CdS/TiO2 [97]
  • BiVO4/Ag NPs [98]
  • PbS Quantum Dots [54]
  • NiOx [99]
(1) Increases electron–hole separation efficiency.
(2) Improves quantum efficiency.
(3) Enhances light absorption.
(4) Boosts performance via surface resonance effect.
(5) Provides a direct pathway for hole extraction.
(6) Expands absorption into the near-infrared and enhances energy conversion efficiency.
5. Post-treatment and Passivation
  • Plasma (Ar, O2) [100]
  • Controlled annealing [101]
  • Strain engineering [102]
  • Small organic compounds [101,102]
  • ALD Al2O3/TiO2 [103]
(1) Passivates surface defects.
(2) Improves crystallinity.
(3) Tunes electronic band structure.
(4) Increases stability and reduces recombination.
(5) Creates a protective layer against photocorrosion.
6. Other Treatment Methods
  • Antireflection coatings [98]
  • Controlled defect creation [104]
  • Laser-sintering [105]
(1) Enhances photon absorption.
(2) Improves electrical conductivity.
(3) Promotes desired defect states for enhanced activity.
Table 3. Summary of representative cocatalyst and surface overlayer strategies for enhancing the PEC performance of BiVO4 photoanodes.
Table 3. Summary of representative cocatalyst and surface overlayer strategies for enhancing the PEC performance of BiVO4 photoanodes.
Structure/StrategyCocatalyst or
Overlayer
Performance HighlightKey EffectRef.
One-step PEC depositionFeOOH (inner) + Co–Sil (outer)6.10 mA/cm2 @1.23 VDual-layer cocatalyst boosts charge separation and reduces recombination[115]
Fluoride-assisted in situ passivationF ionsLong-term stability >100 h @0.6 VSurface passivation and cocatalyst reactivation[116]
Porphyrin-based surface ligandCo–He (Co–O–V linkage)5.3 mA/cm2 @1.23 V,
Von = 0.07 V
Low overpotential and efficient hole transfer[117]
Organic ligand modificationCo2+ + BTC ligand4.82 mA/cm2, onset 0.22 VSurface passivation + cocatalyst anchoring[118]
Magnetic overlayerCo-doped Fe3O41.9× higher OER activity, Faradaic efficiency >85%Improves surface kinetics and protects BiVO4[119]
Dual cocatalyst immersionFeOOH + Co(OH)22.56 mA/cm2, 71.6% retention (10 h)Synergistic catalytic enhancement + stability[120]
Bilayer MOF cocatalystFe-MOF/Ni-MOF1.80 mA/cm2, V_on dropped from 0.9 V to 0.69 VFacilitated interfacial charge transfer[121]
Room-temp photodepositionCo-Pi, Ni-Bi, Mn-Pi/BiVO4Hole transfer efficiency up to 94.5 % @1.23 VRHEConformal, uniform cocatalyst deposition[122]
Facet-selective cocatalyst loadingSelective facet-modified MnOx0.74 mA/cm2 @1.23 VRHEMaximize catalytic activity by crystal facet control[123]
In situ solvothermal growthCOF–Azo1.38 mA/cm2 @1.23 VRHEImproves carrier separation, lowers impedance, and accelerates OER[124]
Bulk Mo doping + surface molecular catalyst depositionCoPOM4.32 mA/cm2 @1.23 VRHEConductivity enhancement + catalytic activation[125]
Table 4. Summary of ηsep and PEC activity in modified BiVO4 photoanodes. ηsep and ηinj extracted under AM 1.5G; all values measured at 1.23 VRHE unless noted.
Table 4. Summary of ηsep and PEC activity in modified BiVO4 photoanodes. ηsep and ηinj extracted under AM 1.5G; all values measured at 1.23 VRHE unless noted.
PhotoanodesJPEC @1.23 VRHE
(mA/cm2)
ηsep (%)ηinj (%)Modification MethodRef.
WO3/S:Bi2O3/(Ga,W):BiVO4/Co-Pi5.10N/RN/RInterface design[140]
Co3O4/BiVO4~2.3N/RN/RCocatalyst interface[141]
Plasma-treated N-doped BiVO41.39~4.6× higher vs. pristineN/RN doping
+ oxygen vacancies
[142]
Co:BiVO4/Mo:BiVO42.0977.886.5Homojunction (doped layers)[139]
Zn:BiVO4/Mo:BiVO42.7065.089.0Homojunction[143]
Ni-BiVO4/FeOOH3.02N/R73.3Homojunction
(+OER overlayer)
[144]
BiVO4/SnO2 (heterostructure)5.6197N/RHeterojunction
+ cocatalyst
[145]
Ov-BiVO4 (VOx engineered)6.299496VOx (oxygen vacancy-engineered)[146]
Table 5. Comparison of BiVO4 photoanodes for overall water splitting with complete reporting.
Table 5. Comparison of BiVO4 photoanodes for overall water splitting with complete reporting.
PhotoanodesBiasElectrolyteJPEC @1.23 VRHE
(mA/cm2)
STH (%)Ref
NiOOH/FeOOH/BiVO4/SnO2//TTO//TOPCon-SiUnbiased1.0 M potassium borate, pH 91.401.72[168]
Nanocone/Mo:BiVO4/Fe(Ni)OOHUnassistedPhosphate buffer, pH 75.82 ± 0.366.2[33]
BiVO4/NiOOH/FeOOH (top)//Cu2O/CuO/TiO2 (bottom)Unassisted0.1 M Na2SO4, pH 62.050.27[169]
BiVO4/FeOOH (oxygen vacancy gradient; FeOOH OEC)Unassisted1 M borate buffer (pH ≈ 9)7.08.4[170]
BiVO4/Cu2O/NiFe-LDHUnassisted0.1 M Na2SO4, pH 65.011.18[171]
Mo:BiVO4 + polycarbazole HTL (CPF-TCB) + NiFeCoOx OECUnassistedK–borate buffer (pH ≈ 9–10)≈6.6~9[110]
CoPi/W:BiVO4/NiUnassistedK-phosphate buffer (pH 7)1.52.1–6.3[172]
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

Nguyen, B.D.; Choi, I.-H.; Kim, J.-Y. Strategies for Enhancing BiVO4 Photoanodes for PEC Water Splitting: A State-of-the-Art Review. Nanomaterials 2025, 15, 1494. https://doi.org/10.3390/nano15191494

AMA Style

Nguyen BD, Choi I-H, Kim J-Y. Strategies for Enhancing BiVO4 Photoanodes for PEC Water Splitting: A State-of-the-Art Review. Nanomaterials. 2025; 15(19):1494. https://doi.org/10.3390/nano15191494

Chicago/Turabian Style

Nguyen, Binh Duc, In-Hee Choi, and Jae-Yup Kim. 2025. "Strategies for Enhancing BiVO4 Photoanodes for PEC Water Splitting: A State-of-the-Art Review" Nanomaterials 15, no. 19: 1494. https://doi.org/10.3390/nano15191494

APA Style

Nguyen, B. D., Choi, I.-H., & Kim, J.-Y. (2025). Strategies for Enhancing BiVO4 Photoanodes for PEC Water Splitting: A State-of-the-Art Review. Nanomaterials, 15(19), 1494. https://doi.org/10.3390/nano15191494

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