Open Access Article
Nguyen Thi Huyenabc,
Thi Viet Ha Luu
d,
Tran Le
ab and
Huu Phuc Dang
*e
aFaculty of Physics & Engineering Physics, VNUHCM-University of Science, Ho Chi Minh City, Vietnam
bVietnam National University Ho Chi Minh City, Ho Chi Minh City, Vietnam
cFaculty of Physics, Ho Chi Minh City University of Education, Ho Chi Minh City, Vietnam
dFaculty of Chemical Engineering, Industrial University of Ho Chi Minh City, Ho Chi Minh City, Vietnam
eFaculty of Fundamental Science, Industrial University of Ho Chi Minh City, Ho Chi Minh City, Vietnam. E-mail: danghuuphuc@iuh.edu.vn
First published on 15th September 2025
The development of efficient and stable photoanodes is critical for advancing photoelectrochemical (PEC) water splitting technologies. In this work, bismuth vanadate (BiVO4) photoanodes were fabricated using a two-step method combining the pulse electrodeposition of bismuth and spin-coating of a vanadium precursor [VO(acac)2], followed by thermal annealing. By systematically varying the pulse voltages and vanadium precursor volume, a series of samples were produced. The sample labeled Bi-576 (deposited at 1.5–1.7 V with 0.6 μL VO(acac)2) exhibited the highest PEC performance. This optimized sample achieved a photocurrent density of 1.33 mA cm−2 at 1.23 V vs. RHE, with an applied bias photon-to-current efficiency (ABPE) of 20% and a charge injection efficiency of 60.1% under AM 1.5G illumination. Structural analysis via X-ray diffraction revealed a preferential (121) crystal orientation and reduced crystallite size, promoting directional charge transport and suppressing recombination. Raman and X-ray photoelectron spectroscopy confirmed the presence of Bi3+, V5+, and strong V–O bonding, along with surface oxygen species that enhance charge separation and interfacial transfer. Field-emission scanning electron microscopy showed a porous, interconnected morphology that increased the electrochemical active surface area (ECSA). Electrochemical impedance spectroscopy and Mott–Schottky analysis revealed a high donor density of 8.65 × 1020 cm−3 and a long interfacial time constant (τint) of 31.46 ms, both contributing to efficient charge transport. Stability tests showed that Bi-576 retained over 82% of its photocurrent after 10 hours of continuous operation, indicating excellent long-term durability. These results demonstrate that tuning the pulse deposition conditions and precursor chemistry enables the rational design of BiVO4 photoanodes with optimized structural and electronic properties. This scalable approach offers a promising route for the development of high-performance photoanodes for solar-driven water splitting.
Electrodeposition has emerged as a cost-effective, scalable, and versatile technique for fabricating metal oxide photoanodes, including BiVO4. Compared to other solution-based or vacuum-based methods, electrodeposition enables direct growth on conductive substrates, good film adhesion, and tunable microstructure through control of deposition parameters.8,9 Importantly, the electrodeposition process can influence not only the thickness and porosity of the film but also the grain connectivity and defect density, which are closely linked to the charge separation and transport properties of BiVO4. Therefore, optimizing the electrodeposition conditions represents a promising strategy to tailor the structural and electronic properties of BiVO4 for enhanced PEC water splitting performance. Two common techniques employed in electrodeposition are direct current (DC) and pulsed current (PC) electrodepositions. Both methods have unique characteristics that influence the growth mechanism of the deposited material and the performance of the BiVO4 film in PEC applications. DC electrodeposition involves applying a constant current to an electrolyte solution, resulting in the continuous deposition of the desired material onto the substrate. This method is relatively simple and easy to control but can lead to the formation of films with non-uniform grain sizes, high internal stress, and poor adhesion, particularly when the deposition rate is too high.10 These issues may negatively impact the crystallinity, surface area, and charge-carrier mobility of the resulting BiVO4 film. In contrast, pulse-current (PC) electrodeposition involves applying current pulses at regular intervals, allowing for control of both the on-time and off-time of the current.11 This technique enables a more controlled nucleation and growth process, leading to films with finer grain sizes, reduced internal stress, and better mechanical properties than the DC-deposited films.12,13 PC electrodeposition also offers the advantage of better control over the composition and morphology of the film, which is crucial for optimizing PEC performance. Several studies have explored the influence of deposition parameters such as pH, temperature, and precursor concentration on the PEC performance of BiVO4 films.14 However, a direct comparison of the growth mechanisms and PEC performance of BiVO4 films deposited using DC and PC electrodeposition methods has not been thoroughly investigated. Such a comparative study would provide valuable insights into the advantages and limitations of each technique and could aid in optimizing the electrodeposition process for synthesizing high-performance BiVO4 photoanodes.
In this study, we present a comparative analysis of the growth and PEC performance of BiVO4 films synthesized using DC and PC electrodeposition. The primary objective of this study is to understand how different deposition techniques influence the structural, morphological, and electrochemical properties of BiVO4 films and how these properties affect the ability of the films to function as photoanodes in PEC water splitting. We hypothesized that PC electrodeposition, with its controlled nucleation and growth processes, would yield BiVO4 films with superior morphology and electrochemical properties compared to DC-deposited films, leading to enhanced PEC performance.
The applied bias photon-to-current efficiency (ABPE) was calculated using the following equation:
![]() | (1) |
![]() | ||
| Fig. 1 (a) X-ray diffraction (XRD) patterns of Bi5DC, Bi536, Bi554, Bi556, Bi574, and Bi576 photoanodes and SnO2. (b) XRD patterns of photoanodes in the 2θ range of 28°–30°. | ||
The average crystallite size of the samples was calculated using the Scherrer equation:18
![]() | (2) |
The full width at half maximum (FWHM) analysis using the Scherrer equation revealed that Bi-576 possessed smaller crystallite domains (∼54.3 Å) compared to Bi-552 (∼367.2 Å), suggesting a higher density of grain boundaries that can facilitate charge separation. Although the FESEM images revealed larger aggregated particles, the smaller XRD-derived crystallites indicated a polycrystalline character within individual particles, contributing to enhanced PEC activity. The smaller crystallite size of Bi576 is attributed to the optimized deposition conditions, which enhance the nucleation rate and result in finer-grained structures. These findings highlight the significant role of crystallinity and preferred orientation in determining the PEC performance of the BiVO4 photoanodes in general. The strong (121) peak intensity and reduced crystallite size of Bi576 contributed to its superior photocatalytic properties by enhancing charge separation and reducing recombination losses. The Bi5DC sample, which corresponds to five deposition layers and has a lower intensity at the diffraction peaks, shows the largest average crystallite size of approximately 367.2 Å. On the other hand, samples under different deposition conditions had smaller crystallite sizes, which were 367.2, 119.4, 85.3, 99.5, and 54.3 Å for Bi536, Bi554, Bi556, Bi574, and Bi576, respectively.
Raman spectroscopy was used to investigate the bonding structure and phase composition of the BiVO4 films. Fig. 2 shows the Raman spectra with characteristic peaks at 202, 322, 356, 709, and 817 cm−1, confirming the presence of the monoclinic scheelite phase.19,20 Among these, the most prominent peak at 817 cm−1 is attributed to the symmetric stretching vibrations of the V–O bond within the VO4 tetrahedra, a signature of the photoactive phase of BiVO4.19 The peak at 202 cm−1 corresponds to external lattice vibrations,21 while the peaks at 322 and 356 cm−1 are associated with asymmetric and symmetric bending modes of the VO4 unit.22 The peak at 709 cm−1 is attributed to the asymmetric V–O stretching. Nikam and Joshi21 reported that the monoclinic phase is a superior photocatalyst compared to the tetragonal phase because it can react thermodynamically and is more stable.
The equation expresses the relationship between the metal and oxygen bond lengths, and the Raman vibrational frequency is as follows:23
| ν (cm−1) = 21349 × 10−1,9176R(Å) | (3) |
Among all the samples, Bi576 exhibited the highest intensity and a noticeable blue-shift of the 817 cm−1 peak, indicating a shorter V–O bond length and improved phase purity. This shift implies a stronger orbital overlap between the V and O atoms, which can enhance the electronic conductivity. The increased Raman intensity further suggests enhanced crystallinity and fewer structural defects, leading to reduced trap states and improved carrier lifetimes in the films. These vibrational enhancements are attributed to the optimized pulse electrodeposition parameters used for Bi576, such as the precise pulse duration and current modulation. These conditions likely facilitated uniform nucleation and preferential formation of the monoclinic scheelite phase. The improved bonding structure directly supports a higher charge carrier mobility and reduces the electron–hole recombination, as reflected in the superior PEC performance of the Bi576 sample. Overall, Raman analysis confirmed that Bi-576 possesses a highly ordered and electronically favorable structure, reinforcing its role as the most efficient photoanode among the samples studied.
Diffuse reflectance spectroscopy (DRS) was employed to investigate the optical absorption characteristics and estimate the bandgap energy (Eg) of the BiVO4 films (Fig. 3). This technique measures the intensity of light reflected from a sample irradiated with a broad-spectrum light source and provides insights into the electronic excitation processes in semiconductors. The optical bandgap (Eg) was calculated using the Tauc plot method, which is based on the relationship between the absorption coefficient (α) and the photon energy (hν), as described by the Wood and Tauc model:24
![]() | (4) |
![]() | ||
| Fig. 3 (a) Absorption spectra and (b) bandgap width of Bi-5DC, Bi-536, Bi-554, Bi-556, Bi-574, and Bi-576 photoanodes. | ||
The Eg values determined for Bi554, Bi574, Bi552, Bi576, and Bi556 were 2.30, 2.39, 2.40, 2.44, and 2.41 eV, respectively. These results indicate that all samples exhibit electronic transitions in the visible region, which is consistent with the requirements for efficient solar light absorption in PEC applications. Notably, Bi576 exhibited the widest bandgap (2.44 eV), which correlates with its enhanced crystallinity and improved structural order, as confirmed by XRD and Raman analyses. These bandgap values are in good agreement with previously reported results for BiVO4 photoanodes synthesized via electrochemical and hydrothermal techniques.26,27 The slight bandgap variations across the samples reflect subtle differences in the microstructure and stoichiometry influenced by the pulse electrodeposition parameters, which ultimately impact the charge carrier generation and separation efficiencies.
Field-emission scanning electron microscopy (FESEM) images (Fig. 4) were used to evaluate the morphology of the BiVO4 samples synthesized via electrochemical deposition. The surface morphology of the bismuth metal (Bi55) film (Fig. 4a) exhibits a complex wavy network structure with loosely packed layers and uneven thickness, indicating uncontrolled nucleation owing to suboptimal pulse parameters. In contrast, the bismuth metal (Bi57) film (Fig. 4b) shows a denser fibrous network, where small clusters are more uniformly distributed and have well-defined edges. This improved morphology reflects the enhanced control over nucleation and grain growth achieved by tuning the pulse deposition conditions. These results highlight the effectiveness of pulse electrodeposition in modulating the structural evolution of the film. Distinct morphological changes were observed across the samples upon conversion to BiVO4 via spin-coating VO(acac)2 and thermal annealing. At lower vanadium precursor volumes (e.g., Bi554 and Bi556), the surfaces appeared smoother and more branched, suggesting partial reactions and moderate grain interconnections. Increasing the VO(acac)2 volume, as in Bi574 and Bi576, led to the formation of well-defined spherical clusters with interconnected chains, particularly for Bi576. This sample exhibited a highly porous and complex structure that increased the active surface area and facilitated electrolyte diffusion, which are favorable features for PEC applications. Overall, both the pulse electrodeposition technique and the VO(acac)2 volume played decisive roles in determining the final BiVO4 morphology. Pulse deposition enabled finer control of nucleation and grain connectivity, as demonstrated by the superior microstructure of Bi57 and the derived BiVO4 film. Meanwhile, a sufficient amount of VO(acac)2 ensured complete reaction and crystal growth, leading to dense and porous architectures. The synergistic effect of the optimized deposition conditions and precursor content resulted in a highly porous and interconnected Bi576 photoanode, which correlated well with its enhanced photoelectrochemical performance.
The TEM and HRTEM images of the Bi576 photoanode (Fig. 4k and l) further validate the crystallographic features suggested by XRD. Clear lattice fringes with an interplanar spacing of 0.309 nm can be observed, which are indexed to the (121) plane of monoclinic BiVO4 (JCPDS No. 14-0688). This direct lattice-resolved evidence is consistent with the enhanced (121) diffraction peak intensity detected in the XRD pattern, confirming the preferential orientation along the (121) plane. Such orientation is known to facilitate efficient charge separation and migration by reducing carrier recombination pathways. The combination of structural ordering, porous morphology, and orientation control accounts for the superior ηinj and photocurrent density achieved by the Bi576 electrode. Therefore, the HRTEM results provide strong microstructural support for the correlation between the crystal orientation and the remarkable PEC activity observed.
Linear sweep voltammetry (LSV) measurements (Fig. 5a) were conducted to assess the photoelectrochemical (PEC) water-splitting performance of the BiVO4 photoanodes in 0.5 M Na2SO4 solution (pH ≈ 6.5) under AM 1.5G illumination (100 mW cm−2). The photocurrent density at 1.23 V vs. RHE followed the increasing order: Bi5DC < Bi536 < Bi574 < Bi554 < Bi556 < Bi576, corresponding to 0.21, 0.50, 0.74, 0.78, 1.04, and 1.33 mA cm−2, respectively. This trend clearly demonstrates that the Bi576 sample exhibited the highest PEC activity. The superior performance of Bi576 can be attributed to a combination of favorable structural and morphological features. XRD analysis revealed a dominant (121) crystallographic orientation, which facilitated anisotropic charge transport and reduced bulk recombination. Raman spectroscopy confirmed enhanced V–O bonding with a shorter bond length, indicative of improved orbital overlap and electronic conductivity. Additionally, the FESEM images show a highly porous, interconnected morphology that increases the electrochemical active surface area and promotes efficient charge transfer at the electrode–electrolyte interface. Importantly, the improved PEC performance of Bi576 also correlated with the precise control of the pulse electrodeposition parameters and optimal amount of the VO(acac)2 precursor. These synergistic factors enhance the structural integrity and surface complexity of the film, leading to superior light absorption, charge separation, and photocatalytic activity. This confirms the critical role of the deposition strategy and precursor design in optimizing BiVO4-based photoanodes for solar water-splitting applications.
Using sulfite oxidation as a kinetic benchmark, the Bi576 photoanode exhibits a surface charge transfer efficiency ηinj = ηtransfer = 60.1% at 1.23 VRHE, outperforming Bi574 (43.3%) and Bi556 (33.2%), as shown in Table 1. The bulk charge separation efficiency reaches ηsep = 29.5% at the same bias, yielding an overall ηoverall = ηsep × ηinj = 17.7%. These values, obtained under AM 1.5G illumination in neutral 0.5 M Na2SO4 and computed following recent protocols,28–30 confirm that the superior PEC activity of Bi576 arises from both efficient bulk charge separation and markedly enhanced interfacial transfer.
| Samples | J ∼ 1.23 V (mA) | Onset potential (VRHE) | ABPE (%) | Cdl (μFcm−2) | ECSA | Rct | t (ms) | N (cm−3) | ηinj (%) | ηsep (%) |
|---|---|---|---|---|---|---|---|---|---|---|
| Bi5DC | 0.21 | 0.55 | 0.64 | 38.00 | 512 | 256.60 | 10.04 | 2.66 × 1020 | 27.20 | 10.39 |
| Bi536 | 0.50 | 0.64 | 1.24 | 43.00 | 538 | 124.00 | 12.04 | 3.58 × 1020 | 31.20 | 21.4 |
| Bi554 | 0.68 | 0.55 | 16.80 | 48.00 | 600 | 256.60 | 15.66 | 39.11 | 22.14 | |
| Bi556 | 1.04 | 0.82 | 14.80 | 48.20 | 598 | 142.00 | 13.42 | 5.77 × 1020 | 33.18 | 41.83 |
| Bi574 | 0.74 | 0.27 | 16.30 | 63.19 | 790 | 92.15 | 29.91 | 43.33 | 22.73 | |
| Bi576 | 1.33 | 0.60 | 20.00 | 75.74 | 947 | 82.79 | 31.46 | 8.65 × 1020 | 60.10 | 29.46 |
The applied bias photon-to-current efficiency (ABPE), calculated from the LSV data (Fig. 5b), further confirmed the superior photoelectrochemical performance of the Bi576 photoanode in this study. Among all the samples, Bi576 exhibited the highest ABPE value of approximately 20% at a low bias voltage (below 0.8 V vs. RHE), underscoring its efficient solar-to-hydrogen conversion capability. The remarkable ABPE performance of Bi576 is attributed to the synergistic effect of improved crystallographic orientation, enhanced V–O bond strength, and a highly porous morphology. The higher vanadium precursor volume enabled a more complete reaction between the Bi and V species during annealing, contributing to improved charge separation and reduced surface recombination. These attributes allowed Bi576 to utilize photogenerated charge carriers more efficiently, translating into a greater applied bias efficiency. This result highlights how tuning the deposition parameters and precursor chemistry can effectively modulate the internal electric field and surface reactivity of BiVO4 films, both of which are essential for maximizing the PEC efficiency under solar irradiation.
Electrochemical impedance spectroscopy (EIS) was performed to investigate the interfacial charge transfer dynamics at the electrode–electrolyte interface under illumination (Fig. 5c). The Nyquist plots revealed semicircles of varying diameters, from which the charge-transfer resistance (Rct) was extracted. The Bi576 sample exhibited the smallest semicircle, indicating the lowest Rct among all the photoanodes, whereas Bi5DC and Bi536 showed much larger semicircles, corresponding to higher interfacial resistance. The lower Rct value of Bi576 is a direct consequence of its interconnected porous morphology, which provides abundant reaction sites and facilitates faster ion diffusion and charge transport kinetics. Additionally, the improved structural ordering and V–O bond strength enhance the electronic conductivity, reducing the resistance to charge injection at the semiconductor/electrolyte interface. These findings are consistent with the LSV and ABPE results and reinforce the conclusion that optimized pulse electrodeposition, coupled with sufficient vanadium precursor loading, is essential for producing BiVO4 photoanodes with excellent charge-transfer characteristics and PEC performance.
Bode phase plots (Fig. 5d) were analyzed to estimate the interfacial time constant (τint) on the BiVO4 surface, which reflects the kinetics of the interfacial charge transfer. The (τint) increased in the following order: Bi5DC < Bi536 < Bi574 < Bi554 < Bi556 < Bi576, with Bi576 achieving the highest lifetime of 31.46 ms. A longer interfacial time constant implies slower recombination kinetics and prolonged charge carrier availability for participation in the oxygen evolution reaction (OER). This significant improvement in Bi576 originates from its well-organized crystal structure and highly porous interconnected network, which collectively reduce trap-assisted recombination. Moreover, enhanced phase purity and reduced defect density, as shown by Raman spectroscopy, further support efficient hole transport to the electrolyte interface. The prolonged carrier lifetime is directly correlated with a higher photocurrent density and ABPE, reinforcing the critical role of both crystallographic and morphological optimizations in PEC performance.
Transient photocurrent measurements (Fig. 5e) were performed under chopped AM 1.5G illumination to evaluate the photostability and responsiveness of the BiVO4 photoanodes. Among all the samples, Bi576 exhibited the most stable and repeatable photocurrent response with minimal current decay during successive light on/off cycles. This behavior is indicative of the efficient separation and transport of photogenerated charge carriers, as well as low recombination rates at the surface. The robust transient response of Bi576 can be attributed to its optimized morphology and superior crystallinity. The interconnected porous structure enhances charge carrier collection, while the reduced structural defects, confirmed by Raman and XRD analyses, limit recombination at the grain boundaries. Moreover, the stronger V–O orbital interactions, as suggested by the blue-shifted Raman peak, contributed to better electronic conductivity and sustained photocurrent output under illumination. This photostability highlights the reliability of Bi576 for long-term PEC operation and confirms that rationally tuned electrodeposition conditions not only improve the performance but also enhance the durability of the device.
The electrochemical active surface area (ECSA) of the BiVO4 photoanodes was estimated from cyclic voltammetry (CV) curves (Fig. S1) using the double-layer capacitance (Cdl) method (Fig. 5f). Bi576 exhibited the highest Cdl value (75.74 μF cm−2), corresponding to the largest ECSA among all the samples. This indicates a greater density of electrochemically accessible sites, which is essential for facilitating surface redox reactions, such as water oxidation. The enhanced ECSA of Bi-576 stems from its hierarchical surface morphology, which is composed of porous, interconnected nanostructures. These features increase the contact area with the electrolyte and promote efficient charge transfer, consistent with the low Rct from EIS and high photocurrent from LSV. Collectively, the increased ECSA, prolonged interfacial time constant, and strong transient photocurrent stability confirm that Bi576 benefits from a synergistic enhancement of both bulk and interfacial charge dynamics, directly contributing to its superior PEC water-splitting performance.
The Mott–Schottky (MS) plots of the BiVO4 photoanodes (Fig. 5g) were measured at a frequency of 1 kHz in 1 M KOH electrolyte under dark conditions. All the samples exhibited positive slopes in the linear region, confirming their n-type semiconductor characteristics. The flat band potentials (Vfb) were estimated by extrapolating the linear portions of the plots to the x-axis, yielding approximate values of −0.08 V for Bi-576 and –0.12 V for Bi-556, respectively. These relatively negative (Vfb) values are favorable for photoelectrochemical water oxidation because they enhance band bending and charge separation at the semiconductor–electrolyte interface.
In addition to Vfb, the donor density Nd—an important parameter reflecting the majority carrier concentration—was calculated from the Mott–Schottky equation:
From the slopes of the linear region in the MS plot, Nd can be determined using the rearranged form as follows:
Bi-576 exhibited the smallest slope, indicating the highest donor density among the tested samples. This higher Nd contributes to improved electrical conductivity and enhanced charge transport within the photoanode. Conversely, the steeper slopes observed for Bi-536 and Bi-5DC imply lower carrier concentrations, which can limit the PEC performance owing to poorer charge transport properties. The combination of a more favorable flat band potential and higher donor density in Bi-576 underscores its superior electronic structure, which further explains its enhanced photocurrent and ABPE values. These Mott–Schottky results are consistent with the EIS, transient photocurrent, and LSV analyses, confirming the critical role of tailored electrodeposition in tuning the semiconductor properties of BiVO4 photoanodes.
Table 2 summarizes the PEC performance of single-layer BiVO4 electrodes prepared by mainstream methods under comparable AM 1.5G testing conditions. SILAR-prepared films can deliver relatively high photocurrents (e.g., ∼1.9 mA cm−2 at 1.23 VRHE in ref. 31), but such approaches often require multiple deposition–annealing cycles and face scalability challenges. Spin-coated/sol–gel films (ref. 1) and spray-pyrolyzed films (ref. 35) generally report lower photocurrents of ∼0.6–1.0 mA cm−2, limited by poor charge separation and transport. Several electrodeposited BiVO4 photoanodes achieve modest baseline values, such as 0.32 mA cm−2 (ref. 32), 0.61 mA cm−2 (ref. 35), and ∼1.1 mA cm−2 (ref. 33), with ABPE values typically below 0.3% and ηinj <30%. Reactive magnetron sputtered BiVO4 exhibits higher performance (≈2.1 mA cm−2 in 1 M K-borate, ref. 34), though still constrained by limited injection efficiency (∼44%) and reliance on alkaline electrolytes. Atomic layer deposition (ALD, ref. 35) yields ∼1.17 mA cm−2 in the phosphate + sulfite electrolyte, but the use of a sacrificial hole scavenger prevents direct benchmarking against water oxidation. Hydrothermally synthesized BiVO4 (ref. 36) reports stable operation but with modest photocurrents (<2.0 mA cm−2). In contrast, our optimized Bi576 electrode achieves a balanced and competitive performance in neutral electrolyte (0.5 M Na2SO4, pH ≈ 6.8): 1.30 mA cm−2 at 1.23 VRHE, an ABPE of 20%, and a significantly higher ηinj of 60.1%. Compared with earlier electrodeposited films (ηinj typically 20–30%), this result highlights much more efficient interfacial charge transfer and reduced recombination. The performance advantage originates from our intensity-tuned pulse electrodeposition coupled with precursor-volume control, which produces (i) a porous interconnected morphology facilitating electrolyte infiltration, (ii) preferred (121) crystal orientation favoring charge transport, and (iii) an elevated donor density (Nd ≈ 8.65 × 1020 cm−3) that strengthens space-charge separation. Importantly, the work also verifies oxygen and hydrogen gas evolution and demonstrates >82% retention after 10 h of operation, which many literature reports lack. These structural and electronic optimizations enable more efficient utilization of photogenerated charges and position our approach among the most efficient and well-balanced single-layer BiVO4 photoanodes synthesized by electrochemical methods to date, offering a practical and scalable pathway beyond ALD or sputtering while surpassing the efficiency and stability of conventional electrodeposited counterparts.
| BiVO4 preparation method | Photocurrent density @1.23 VRHE (mA cm−2) | ABPE (%) | Stability | Electrolyte | Ref |
|---|---|---|---|---|---|
| SILAR (thin films, optimized cycles) | 1.95 | 0.5–0.8 | — | 0.5 M Na2SO4 (pH 7) | 31 |
| SILAR (nanostructured BiVO4) | 1.2 | ∼0.8 | >1 h, >90% retained | 0.5 M Na2SO4 (pH 7) | 37 |
| Electrodeposition (BiOI precursor → BiVO4) | 0.32 | — | — | 0.1 M KPi buffer (pH 7) | 32 |
| Electrodeposition (Bi metal → BiVO4) | ∼1.1 | 0.15 | >2 h, stable | 0.1 M KPi buffer (pH 7) | 33 |
| Electrodeposition of BiVO4 | ∼1.0 | — | Stable | 0.5 M Na2SO4 (pH 7) | 38 |
| Reactive magnetron sputtering | 2.1 | 0.4 | — | 0.5 M Na2SO4 (pH 7) | 34 |
| Spin coating | ∼0.8 | 0.06 | — | 0.5 M Na2SO4 (pH 7) | 1 |
| Atomic layer deposition | 1.17 | — | 20% drop after 2 h | 0.5 M Na2SO4 | 35 |
| Spray pyrolysis | 1.66 | — | ∼5 min | 0.5 M borate buffer (pH 9.5) | 39 |
| 0.58 | — | — | 0.5 M Na2SO4 (pH 7.2) | 40 | |
| Hydrothermal synthesis | 0.5 | 0.5 M Na2SO4 (pH 6.8) | 36 | ||
| 0.075 | 1 M KOH (pH 13.3) | 41 | |||
| Electrodeposition of BiVO4 | 1.30 | 0.2 | 18% drop after 10 h | 0.5 M Na2SO4 (pH 6.8) | This work |
Fig. 6a shows the time-dependent gas evolution profiles of H2 and O2 over the Bi-576 photoanode under AM 1.5G illumination at 1.23 V vs. the RHE. The hydrogen generation rate steadily increased and reached approximately 115 μmol h−1 after 5 h, whereas the oxygen evolution rate followed a similar linear trend, reaching ∼58 μmol h−1. The observed H2
:
O2 molar ratio was close to the theoretical stoichiometric ratio of 2
:
1, confirming the high faradaic efficiency of the PEC process. These results validate that Bi576 can facilitate overall water splitting with excellent charge separation and utilization under continuous illumination. Fig. 6b illustrates the long-term photostability of Bi576 evaluated by chronoamperometry (I–t) at 1.23 V vs. RHE over 600 min of continuous illumination. The photocurrent density decreased slightly from ∼1.45 to ∼1.20 mA cm−2, maintaining more than 82% of its initial value after 10 h. This slow degradation suggests strong structural stability and resistance to photocorrosion, which is likely attributed to the optimized crystal orientation, enhanced carrier concentration, and porous morphology that allow efficient charge transport and reduce surface degradation.
![]() | ||
| Fig. 6 (a) Hydrogen evolution and oxygen evolution–time curves of Bi-576 per area (1 cm2) under AM 1.5G illumination. (b) Photostability tests (I–t curves at 1.23 VRHE) over the Bi-576 photoanode. | ||
The surface chemical composition and oxidation states of the Bi-576 photoanode were further investigated using X-ray photoelectron spectroscopy (XPS), as shown in Fig. 6. The high-resolution Bi 4f spectrum (Fig. 7a) exhibits two characteristic peaks at 158.63 eV and 163.97 eV, corresponding to Bi 4f7/2 and Bi 4f5/2, respectively.8,42 The ∼5.34 eV spin-orbit splitting is consistent with the +3-oxidation state of bismuth (Bi3+), confirming the correct incorporation of Bi into the BiVO4 lattice. In the V 2p region (Fig. 7c), two distinct peaks located at 515.70 eV (V 2p3/2) and 523.21 eV (V 2p1/2) are observed, which correspond to the V5+ oxidation state.43 The sharp and symmetric nature of these peaks indicates a well-crystallized and stoichiometric incorporation of vanadium, corroborating the Raman results that show strong V–O bonding and phase purity. The O 1s spectrum (Fig. 7b) is deconvoluted into two main components at 529.18 eV and 531.94 eV. The peak at 529.18 eV corresponds to lattice oxygen (O2−) bound within the BiVO4 structure, while the higher binding energy component at 531.94 eV is attributed to adsorbed hydroxyl groups (–OH) or surface oxygen species.44 The presence of these surface oxygen species suggests potential surface defect states or functional groups that could act as active sites for water oxidation reactions. Such surface features can facilitate interfacial charge transfer and enhance photocatalytic performance. Taken together, the XPS results confirm the successful formation of BiVO4 with the correct oxidation states of Bi3+ and V5+, as well as the presence of beneficial surface oxygen species. These findings align well with the structural and electrochemical analyses, reinforcing the conclusion that the optimized pulse electrodeposition method and precursor volume not only improved the crystallinity and morphology but also fine-tuned the electronic structure of the photoanode. The strong V–O orbital interaction and appropriate surface chemistry collectively contribute to the excellent PEC performance and photostability of the Bi576 photoanode.
![]() | ||
| Fig. 7 High-resolution (a) Bi 4f, (b) O 1s, and (c) V 2p peaks of Bi-576 photoanode before and after PEC stability testing for 60 min (0.5 M Na2SO4). | ||
Supplementary information is available. See DOI: https://doi.org/10.1039/d5na00667h.
| This journal is © The Royal Society of Chemistry 2025 |