Lucas Leão Nascimentoa,
Rafael A. C. Souzaab,
Paulo H. Horta Nunesa,
João F. C. S. Costad,
Ivo A. Ricardoaef,
Eduardo M. Oliveiraae,
Klaus Krambrock
d,
Chuanyi Wang
c and
Antonio Otavio T. Patrocinio
*ae
aLaboratory of Photochemistry and Materials Science, LAFOT-CM, LAFOT-CM Instituto de Química, Universidade Federal de Uberlândia, 38400-902, Uberlândia, MG, Brazil. E-mail: otaviopatrocinio@ufu.br
bFaculdade De Ciências Exatas e Tecnologia-FACET, Departamento de Ciências Exatas, Universidade Federal de Grande Dourados, 79084-970, Dourados, MS, Brazil
cSchool of Environmental Science and Engineering, Shaanxi University of Science and Technology, Xi'an, 710021, PR China
dDepartamento de Física, Universidade Federal de Minas Gerais, 31270-901, Belo Horizonte, MG, Brazil
eCentro de Excelência em Hidrogênio e Tecnologias Energéticas Sustentáveis – CEHTES, Parque Tecnológico Samambaia, 74690-631, Goiânia, GO, Brazil
fUniversidade Save, Faculdade de Ciências Naturais e Exactas, 0301-01 Chongoene, Gaza, Mozambique
First published on 18th August 2025
Solar-induced valorization of agroindustrial residues is a key issue for the so-called sustainable growth. In this study, a novel Nb-doped BiVO4 photoanode is described and applied for the photoreforming of crude glycerol and flegmass, real byproducts of biodiesel production and sugarcane ethanol refining, respectively. To the best of our knowledge, this is one of the first reports exploring flegmass as a substrate for photoreforming. Nb(V) doping significantly enhances the photoanode performance, with 5% Nb:
BiVO4 exhibiting 80% increase in photocurrent during long-term crude glycerol photoreforming, reaching 2.2 mA cm−2 at 1.0 V vs. RHE. When flegmass is employed, smaller photocurrents are observed in long-term experiments due to the lower organic content. Nevertheless, the 5% Nb
:
BiVO4 photoanode exhibits a 200% increase in photocurrent compared to the pristine oxide. Detailed characterization reveals that Nb5+ doping occurs dominantly at VO4 tetrahedral sites, also leading to oxygen vacancies as evidenced by EPR spectroscopy. Electrochemical impedance spectroscopy showed higher charge carrier densities and reduced charge transfer resistance in the Nb-modified photoanodes. Photoreforming of crude glycerol with 5% Nb
:
BiVO4 after 6 hours of simulated sunlight irradiation achieved 91.4% substrate conversion, producing green H2 at a rate of 4.2 μmol cm−2 h−1, much higher than 2.3 μmol cm−2 h−1 obtained by undoped BiVO4. EPR spin trapping experiments indicated that the main reaction mechanism involves direct hole transfer to adsorbed glycerol on the catalyst surface, forming organic radical species. In flegmass photoreforming, 5% Nb
:
BiVO4 achieved H2 evolution rates of 2.4 μmol H2 per cm2 per h, consistently outperforming the bare BiVO4 photoanode, highlighting its potential for valorizing low-cost, biomass-derived residues into valuable chemicals and renewable fuels.
Sustainability spotlightThe efficient and sustainable valorization of industrial waste streams is essential for promoting a circular economy and mitigating environmental impacts. This work reports a Nb-doped BiVO4 photoanode for the solar-driven reforming of real industrial waste streams, crude glycerol and flegmass, into value-added compounds and green hydrogen. Niobium doping enhances charge carrier dynamics and photocatalytic activity, enabling a scalable strategy for waste treatment coupled with clean fuel production. This work directly supports the UN Sustainable Development Goals, particularly SDG 7 (Affordable and Clean Energy), SDG 9 (Industry, Innovation and Infrastructure), and SDG 12 (Responsible Consumption and Production). |
Photo-assisted reactions have been extensively studied for waste treatment and valorization.1–7 In particular, the photoreforming process utilizes industrial residues as sacrificial agents for low-carbon H2 production, integrating principles of green chemistry with environmental remediation. Moreover, beyond H2 generation, the selective oxidation of organic substrates can yield high-value chemicals, enhancing the overall economic viability of the process.
Photoelectroreforming is a promising approach that combines cost-effective and abundant photocatalysts with the enhanced kinetics offered by electrocatalysis. In this process, sunlight drives the desired reactions, while a small external bias improves charge separation and helps prevent undesired electron recombination.5,8,9 This approach employs semiconductor oxides, such as TiO2, WO3, Fe2O3, and BiVO4, immobilized on transparent conductive substrates to function as photoanodes in a photoelectrochemical cell. Among these, BiVO4 stands out due to its favorable light absorption properties, relatively high photocurrent, and near-optimal theoretical solar-to-hydrogen (STH) efficiency.10 However, BiVO4 is highly susceptible to photocorrosion and exhibits short-lived carriers, resulting in sluggish kinetics and a lack of long-term stability.11 Many approaches have been studied to address this phenomenon. Ke and co-workers reported that surface doping of BiVO4 photoanodes with Mo5+ or W5+ significantly increased photocurrent and stability for water oxidation.12 Liang and co-authors decorated the BiVO4 surface with MoOx nanoparticles, increasing the conversion of poly(ethylene-terephthalate) into formate, acetate, and H2, which the authors attributed to improved charge transfer within the materials.7 Kalanur and Seo reported Nb(V)-doped BiVO4 photoanodes with enhanced water splitting performance when compared to pristine BiVO4. Doping with Nb(V) increased the carrier density in the bulk of BiVO4, improved the charge separation, charge transfer, and diffusion, and decreased the charge transfer resistance at the photoanode/electrolyte interface.13 So far, such strategies have been poorly explored on photoelectroreforming, especially those employing real biomass-derived residues.
Biomass-derived industrial waste represents the largest renewable carbon source globally.14 Most of these are lignocellulosic solid residues, known for their recalcitrant nature, which poses a considerable challenge to employing them as sacrificial agents in light-driven reactions.1,15–17 However, several promising alternatives exist, particularly in the form of liquid waste streams. One notable example is glycerol, the primary byproduct of biodiesel production, currently available in large surplus. As an easily oxidizable triol in its refined form, glycerol has been widely recognized as an effective electron donor for photoreforming.4,5,18,19 Additionally, some oxidation products of glycerol, such as dihydroxyacetone (DHA), glyceraldehyde, and formic acid, are valuable chemicals with diverse applications. Glycerol global production reached approximately 3 to 4 million tons annually as of recent estimates.20,21 Costly refinement processes, however, are often required to convert crude glycerol for commercial uses, involving purification steps such as distillation and ion exchange. Moreover, photoreforming of crude glycerol is difficult to accomplish, as pointed out by Ribao and co-workers, due to the presence of impurities and the increased turbidity of crude glycerol suspensions that severely compromise conversion efficiency.22 Furthermore, the high viscosity of glycerol hinders its diffusion through the photocatalyst, requiring significant dilution to achieve practical conversion rates. Thus, there is a need for processes that enable the direct conversion of crude glycerol, eliminating costly additional steps.
Flegmass is another example of a widely available liquid waste stream. This aqueous distillation residue originates from the sugarcane processing industry (bio-ethanol production) and is composed of residual short-chain alcohols (e.g., methanol and ethanol), traces of organic acids, esters, aldehydes, and ketones, as well as isoamyl alcohol and its derivatives.23 Not to be mistaken for distillery stillage, flegmass is one of the most significant waste streams in the sugarcane processing plants, especially in Brazil, where its production reaches around 150 million cubic meters annually.24 The use of flegmass as a sacrificial agent for photoreforming is advantageous because this aqueous residue can be used directly, eliminating the need for prior processing and/or dilution. However, to our knowledge, there are no studies exploring the valorisation of flegmass.23,25
Herein, for the first time, a photoelectrochemical cell assembled with Nb(V)-doped BiVO4 (Nb:
BiVO4) was applied for the photoreforming of crude glycerol and flegmass residues, collected from a biodiesel plant and a sugarcane processing refinery, respectively. The morphological and electronic properties of the photoanode were evaluated as a function of dopant concentration, followed by photoelectrochemical assays employing real residues for hydrogen production and biomass valorization.
Raman spectra (Fig. 1b) show peaks characteristic of the monoclinic phase, particularly the signal at 830 cm−1 assigned to the asymmetric stretching of the tetrahedral V–O bonds. Bands around 320 and 367 cm−1 were attributed to asymmetric and symmetric deformation modes of the VO43− tetrahedron. Additionally, the BiVO4 sample also shows a less intense stretching peak at 248 cm−1 relative to the Bi–O symmetric vibration of the tetragonal phase. These bands are broadened as Nb content increases, with a pronounced effect on the bands associated with the deformations in the VO43− tetrahedra. This change indicates loss of the short-range arrangement due to Nb5+ doping at V5+ sites. Additionally, the slight shift of Raman features toward lower wavenumbers suggests changes in the short-range symmetry of the VO43− tetrahedra,26,31 further supporting the hypothesis that Nb5+ doping occurs at V5+ sites. Additional characterization studies were carried out for the 5% Nb:
BiVO4 sample and compared with pristine BiVO4.
X-ray photoelectron spectroscopy (XPS) was performed to analyze the surface composition and electronic properties of the different BiVO4 samples. Fig. 2a shows the XPS survey spectra of all synthesized materials, indicating no significant contamination. The high-resolution XPS spectra in Fig. 2b show two distinct oxygen peaks for both the pristine and the 5% Nb-doped BiVO4 samples. The peak at lower binding energy corresponds to lattice oxygen (Olig), while the less intense peak at higher binding energy is attributed to oxygen-adsorbed species and can be correlated with oxygen vacancies (Ovac) in the BiVO4 surface.32,33 The relative intensity of the oxygen vacancy peak is twice as high in the 5% Nb:
BiVO4 sample compared to BiVO4, suggesting an increased concentration of oxygen vacancies due to niobium doping. In the spectral region of the Nb 3d peak (Fig. 2c) the expected doublet for Nb5+ species is observed, further confirming the presence of Nb(V) at the surface of 5% Nb
:
BiVO4.13,34,35 Elementary quantification of the surface reveals a niobium content of 5.1 ± 0.1% wt% for the 5% Nb
:
BiVO4 sample, while the vanadium content decreased by 4.9 ± 0.1% relative to the pristine sample (Table S1) corroborating the hypothesis that Nb5+ replaces V5+ in the lattice.
EPR spectra measured at room temperature with a microwave frequency of about 9.43 GHz for the different solid-state powders are shown in Fig. 2d. The spectra show a broad paramagnetic resonance line (line with about 75(5) mT) with a g factor of about 2.05(5), typical of paramagnetic salts of transition metal ions (Nb4+ and V4+) or paramagnetic niobium–oxygen vacancy complexes. The highest EPR signal intensity is observed for the BiVO4 sample doped with 5% of niobium, followed by the sample with 1% of Nb doping. The pristine sample does not show this broad paramagnetic signal. The results suggest that part of the Nb doped into the BiVO4 samples is incorporated in the samples as Nb4+, which is paramagnetic with spin S = ½ compared with Nb5+ that is diamagnetic. At first glance, this result is contradictory to the XPS analysis, in which only Nb5+ signals were clearly observed, but this is actually explained by the sensitivity of each technique. By comparing the EPR signal intensity with that of an anhydrous CuSO4 standard powder sample, the concentration of the broad EPR signal in the 5% Nb:
BiVO4 sample is estimated to be approximately 1015 cm− 3. Assuming that Nb4+ is principally incorporated at the surface, possibly complexed with oxygen vacancies, the concentration of these defects in the surface region would be about 3 orders of magnitude higher. This concentration is still below the detection limit for XPS spectroscopy, explaining the missing Nb4+ transitions in the XPS spectra. One can conclude, then, that the increased oxygen vacancy concentration on the Nb
:
BiVO4 samples can be associated with the formation of Nb4+ species. High-Resolution Transmission Electron Microscopy (HRTEM) images (Fig. 3a–d) show both BiVO4 and 5% Nb
:
BiVO4 samples with interplanar spaces of 0.31 nm attributed to the (112) plane of monoclinic BiVO4, in good agreement with the XRD spectra in which high relative intensities for the (112) lattice plane are observed for all samples.36,37 Moreover, EDS elemental mapping of 5% Nb
:
BiVO4 powder (Fig. S1(a and f)) shows the presence of niobium at the surface of the material.
The powder samples were then deposited onto clean FTO substrates using the screen-printing method. Scanning Electron Microscopy (SEM) images of the screen-printed photoanodes (Fig. S2a and f) show that the BiVO4-based films are formed by the agglomeration of spherical particles formed during the sintering step. Cross-section SEM images (Fig. S3) were taken to measure the thickness of the films. The measured average thickness was 1.9 ± 0.2 μm. EDS mapping of the film surface and the cross-section evidenced the homogeneous distribution of Nb, Bi and V on the film, as expected. Diffuse reflectance measurements (Fig. 4a) indicate that Nb5+ doping did not induce any drastic changes in the optical properties of BiVO4. The strong absorption band up to 500 nm accounts for the intense yellow color of the photoanodes, Fig. 4a (inset). All synthesized samples displayed a band gap energy (Eg) of 2.45 ± 0.05 eV (Fig. 4b) consistent with other BiVO4 samples obtained through hydrothermal synthesis.27,28,38
Electrochemical Impedance Spectroscopy (EIS) was employed to gain insights into the charge transport process in the obtained photoanodes. The EIS data (Fig. S4) show that both pristine BiVO4 and Nb-doped samples exhibit similar behavior under dark conditions, as expected, since the crystalline properties were not significantly affected by Nb5+ doping. However, under illuminated conditions (Fig. 5a) the 5% Nb:
BiVO4 sample demonstrated significantly lower total impedance, when compared to the 1% Nb
:
BiVO4 and undoped samples, with both displaying a more capacitive response, characterized by higher phase angles when compared to the 5% Nb
:
BiVO4 sample (Fig. 5b).
The EIS data were fitted using a simplified Randle's circuit as a model (Fig. 5a, inset), in which Rs is the resistance of the electrolyte, Cdl is the capacitance of the double-layer, and Rct is the charge transfer resistance. The fitting data parameters (Table S2) indicate no relevant changes in the electrochemical properties between BiVO4 and 1% Nb:
BiVO4. However, the 5% Nb
:
BiVO4 films showed a much lower charge transfer resistance of 634 Ω under simulated sunlight, when compared to 1290 Ω obtained by the pristine BiVO4 film under the same conditions. This behavior was further investigated using Mott–Schottky plots (Fig. 5c and d) which show that the 5% Nb
:
BiVO4 films exhibit a charge carrier density of −8.15 × 1019 cm−3, twofold higher than that compared to the pristine sample of −4.40 × 1019 cm−3. In contrast, the flat-band potential (Vfb) was not significantly affected by Nb5+ doping, with both samples displaying similar Vfb values ranging from −0.012 (BiVO4) to 0.069 V vs. RHE (5% Nb
:
BiVO4), consistent with the literature.13,39,40 Thus, Nb5+ doping greatly improved charge transport, resulting in an increased charge carrier density across the photoanode surface, but without significantly shifting either the conduction band or the bandgap energy of BiVO4.
EPR spin trapping experiments with DMPO (0.1 M) in distilled water suspensions (pH ∼ 6), containing 5% Nb:
BiVO4 powder under magnetic stirring, show the formation of radical adducts under white light illumination. Fig. 6a and b display the EPR spectra of DMPO spin adducts as a function of time in pure water and mixed water/glycerol (10%) suspensions. The radical adducts were identified by spectral simulation of the spectra using the Easyspin® routine in Matlab (Fig. 6c). The two DMPO adducts were identified as DMPO–˙OH and DMPO–˙R. The EPR spectrum of the DMPO–˙OH adduct is characterized by a hyperfine interaction between the electron spin S = ½ and the nuclear spin I = 1 (from 14N) and the nuclear spin I = ½ (from 1H) leading to aN = aHβ = 1.49 mT, resulting in the well-known four-line spectrum with intensity ratios of 1
:
2
:
2
:
1.41 The EPR spectrum of the DMPO–˙R adduct is described by a six-line spectrum with equal intensities and hyperfine parameters of aN = 1.57 mT and aHβ = 2.29 mT. This latter DMPO adduct is less specific, but it is probably of organic nature, attributed to a carbon-centered radical adduct of DMPO (aHβ > aN)42 and references therein, different from the DMPO–˙OH adduct, which is oxygen-centered (aN > aHβ). In a mixed water/glycerol (10%) suspension, DMPO formed two different spin adducts, while only one spin adduct was observed in pure water. These results indicate that the organic radical is produced from glycerol, while hydroxyl radicals are produced from water. The concentration of the organic radicals depends on the ratio of the water/glycerol mixture. A higher glycerol ratio increased the related EPR DMPO–˙R adduct concentration (not shown).
Acidic pH also favors the formation of these organic radicals. The reaction kinetics are strongly influenced by the pH of the solution. When the same spin trapping experiments were performed at pH ∼2, the same DMPO spin adducts were observed as at pH ∼6; however, with increased concentrations (Fig. 6d), a similar result was reported elsewhere.13 In contrast, in mixed water/glycerol (10%) suspensions at pH ∼2, we also noticed some formation of spin adducts in the dark (not shown).
When the spin trapping experiments were carried out with the 5% Nb:
BiVO4 photoanode coupled with a Pt cathode under a bias of about ∼1.0 V vs. RHE in the presence of DMPO (0.2 M) in mixed water/glycerol (10%) solution and K2SO4 electrolyte (0.1 M) at pH ∼6, the EPR results (Fig. 6e) reveal the formation of the same spin adducts observed before, but with higher intensities, indicating an accelerated kinetics of the reactions and higher yield for the formation of ˙OH radicals. Comparing different BiVO4 photoanodes (Fig. 6f) the higher Nb5+ doped sample (5%) showed higher efficiency (∼35%) in the formation of DMPO–˙OH adducts under white light illumination than the undoped BiVO4 or 1% Nb
:
BiVO4 ones. These results indicate a significantly enhanced ability to generate organic radicals and hydroxyl radicals, which aligns with the EIS results, as the 5% Nb
:
BiVO4 photoanode also exhibited a higher charge carrier density compared to pristine BiVO4. Moreover, radical formation is largely dependent on the applied bias and illumination, correlating with the much lower charge transfer resistance due to surface defects (oxygen vacancies and/or Nb4+ ions) as observed for 5% Nb
:
BiVO4 films under illumination.
Our findings with DMPO EPR spin trapping align well with the reaction mechanism proposed before (ref. 13). Dong Liu et al. utilized a nanoporous BiVO4 photoanode for the photoelectrochemical oxidation of glycerol and discovered that glycerol strongly adsorbs onto the BiVO4 surface. In the initial step of the reaction, the adsorbed glycerol molecules are converted into glycerol radicals through direct hole oxidation. These radicals then react with water from the electrolyte, followed by a dehydration step, leading to different subproducts of glycerol. In our experiments using DMPO EPR spin trapping, we also observed the generation of organic radicals, although their exact nature remains unknown. Additionally, we detected hydroxyl radicals (see Fig. 6a and b), which may play an important role in guiding the complex photoreforming reaction mechanisms of glycerol towards different subproducts.
Light-chopped linear sweep voltammograms (LSVs) (Fig. 7) show all tested photoanodes with a similar photoresponse profile, consisting of a sharp increase in photocurrent upon light exposure and a prompt decrease when the light is switched off. The LSVs with glycerin (Fig. 7a) show that the 5% Nb:
BiVO4 photoanode achieved the highest photocurrent, approximately 70% higher than that of the pristine BiVO4 sample, which in turn produced similar photocurrents to the 1% Nb
:
BiVO4 photoanodes. When crude glycerol was used (Fig. 7b), the same trend was observed, although the overall photocurrent was 20% lower. Finally, with flegmass (Fig. 7c), the 5% Nb
:
BiVO4 photoanodes yielded the same level of photocurrent as with pure glycerin, showing nearly 100% increase compared to the other photoanodes. Under all tested conditions, the 5% Nb
:
BiVO4 films consistently outperformed the others, while the 1% Nb
:
BiVO4 and pristine BiVO4 films produced similar results, suggesting that lower Nb5+ doping levels did not significantly impact BiVO4 performance.
Long-term photoreforming of crude glycerol was conducted using pristine BiVO4 and 5% Nb:
BiVO4 films, as the latter exhibits the best-performance among the Nb-doped photoanodes. The 5% Nb
:
BiVO4 sample yielded a higher oxidative photocurrent (Fig. 8a) converting 91.4% of glycerol, compared to 73.3% for pristine BiVO4, after 6 hours of irradiation (A.M. 1.5 100 mW cm−2). The H2 evolution rate was also significantly higher for 5% Nb
:
BiVO4, reaching 4.2 μmol H2 per cm2 per h versus 2.3 μmol H2 per cm2 per h with BiVO4 (Fig. 8b). The faradaic efficiency for H2 evolution was 96% for 5% Nb
:
BiVO4 and 98% for BiVO4, indicating that crude glycerol photoreforming is viable for sustained green H2 production.
Regarding oxidation products, formic acid was the main product for both pure glycerin and raw glycerol, with a C-based selectivity of 89% for 5% Nb:
BiVO4 and 71% for BiVO4. Other minor products, including oxalic acid (2%), acetic acid (5%), and dihydroxyacetone (<1%), were also detected, as shown in NMR and HPLC data in the supplementary information (Fig. S6 and S7). Moreover, CO2 evolution was also qualitatively detected during the photoelectroreforming assays, suggesting that the remaining amount of the consumed glycerol was eventually mineralized. XPS analysis of the films subjected to 8-hour-long crude glycerol photoelectroreforming assays was conducted to evaluate the long-term stability of the photoanode. The results were compared to fresh prepared photoanodes (Fig. S8). The films exposed to the glycerol solution exhibited a higher relative carbon content as part of the organic species remained adsorbed on the surface. It was possible, however, to clearly identify the presence of Nb5+, Bi3+ and V5+ signals, with similar chemical shifts as those observed in the fresh photoanode. The determined Nb
:
Bi ratio in the surface of the photoanode after the photoelectrochemical assay was 0.15, while for the fresh sample, it was 0.17. SEM images and EDS analysis of the used photoanode (Fig. S2d–f) showed that the film morphology and composition remained very similar to those observed for fresh prepared samples. Moreover, ICP-OES analysis of the anolyte has only detected traces (<0.5 ppm) of Bi(III) and Nb(V), evidencing the stability of the 5% Nb
:
BiVO4 in the K2SO4 electrolyte containing 10% v/v crude glycerol.
When flegmass was used as a sacrificial agent in long-term photoreforming experiments, the current density was consistently higher for 5% Nb:
BiVO4 (Fig. 9a), which also resulted in a higher H2 evolution rate of 2.4 μmol H2 per cm2 per h, compared to 0.42 μmol H2 per cm2 per h achieved with the pristine BiVO4 photoanode (Fig. 9b). As flegmass is a more diluted and complex matrix than glycerol, the quantification and identification of liquid oxidation products proved challenging. HPLC analysis of the crude residue revealed the presence of ethanol and isoamyl alcohol as the main components (Fig. S9). Alternatively, the consumption of organic species during photoelectroreforming was evaluated based on Chemical Oxygen Demand (COD) removal. 5% Nb
:
BiVO4 exhibited superior COD removal of up to 70%, compared to 38% for BiVO4 after 24 h irradiation (Fig. 10a). HPLC analysis in different time intervals (Fig. 10b) reveals the build-up of acetic acid as the primary oxidation product, with some formic acid appearing at longer irradiation times. These results are promising, especially given that flegmass was used as received, without dilution or pre-treatment, an important consideration for enabling practical applications. The superior performance observed for the 5% Nb
:
BiVO4 photoanodes can be explained by the increased charge carrier density, decreased charge transfer resistance at the BiVO4/electrolyte interface, and the higher presence of reactive surface states due to Nb5+ doping.
Other researchers have explored similar strategies to enhance the biomass photoelectroreforming capabilities of BiVO4. Some representative results are shown in Table 1. Bora and co-workers reported Mo-doped BiVO4 photoanodes with improved current density in the presence of glycerol.18 This enhancement was attributed to the formation of a capacitance layer and a reduction in charge transfer resistance on the photoelectrode, described as a photocharging phenomenon. However, no direct comparison was made with pristine BiVO4 photoanodes. Similarly, Tateno et al. investigated the effect of Ta-doping on BiVO4 photoanodes for the photoelectrochemical oxidation of glycerol.43 They observed a significant increase in photocurrent of over 60% compared to undoped BiVO4, along with enhanced photoanode stability in acidic electrolytes. Notably, Ta doping did not significantly alter the band positions or structural properties of BiVO4, similarly to what we have observed for Nb-doped samples. However, Nb doping resulted in a superior increase of 80% in photocurrent, evidencing its effectiveness in improving BiVO4-based photoanodes.
Photoanode | Undoped BiVO4 photocurrent densitya | Metal-doped BiVO4 photocurrent densitya | Experimental remarksb | Ref. |
---|---|---|---|---|
a Estimated based on reported LSV curves (±0.1 mA cm−2).b The performance values were taken directly from the respective references and were obtained under varying experimental conditions (e.g., electrolyte composition, sacrificial agent, light source, and measurement protocols). Direct quantitative comparison between different studies should therefore be made with caution. | ||||
W![]() ![]() |
NA | 3.1 mA cm−2 | 0.5 M potassium borate buffer with 0.1 M glycerol | 46 |
Mo![]() ![]() |
NA | 4.6 mA cm−2 | Glycerol 20% v/v with 0.1 M phosphate buffer. After photocharging effect | 18 |
Ta![]() ![]() |
1.6 mA cm−2 | 2.6 mA cm−2 | Glycerol 1.0 M with H2SO4 25 mM | 43 |
Nb![]() ![]() |
1.2 mA cm−2 | 2.2 mA cm−2 | Glycerol 10% v/v with 0.1 M K2SO4 | This work |
Nb![]() ![]() |
0.8 mA cm−2 | 2.4 mA cm−2 | Flegmass with added 0.1 M K2SO4 | This work |
In summary, Nb doping at 1% and below did not significantly affect the properties of BiVO4. However, when the Nb content was increased to 5 wt%, charge transfer across the photoanode bulk and surface was enhanced, as evidenced by the higher charge carrier density and significantly lower charge transfer resistance of the 5% Nb:
BiVO4 photoanode compared to pristine BiVO4, which was attributed to the introduction of Nb5+ in the lattice and surface defects (oxygen vacancies and/or Nb4+ species). This improvement was reflected in the photoreforming results, where 5% Nb
:
BiVO4 demonstrated more than twice the photocurrent and hydrogen evolution.
The improved photoelectrochemical performance 5% Nb:
BiVO4 photoanodes compared to pristine BiVO4 was also observed when real waste streams (crude glycerol and flegmass) were employed in the photoelectrochemical cells. The use of real biomass residues in photoelectroreforming is an unexplored field in the literature. A few examples can be found using photocatalytic systems based on suspensions of semiconductor nanoparticles. Skaf et al. showed that for photocatalytic reforming, activity losses of up to 3.5 times occur when crude glycerol is used instead of purified glycerol.44 Villachica-Llamosas et al. evaluated the hydrogen evolution from pure glycerol solutions prepared with natural waters, revealing decreased efficiencies when compared to high-purity water.45 Our system stands out by demonstrating high selectivity and performance with non-treated crude glycerol, with no additional conditioning, indicating that Nb-doped BiVO4 tolerates real waste conditions. The excellent performance is accompanied by improved stability under irradiation, which serves as proof of concept for the possible large-scale application of photoelectroreforming.
The BiVO4-based photoanodes were deposited on clean FTO substrates using the screen printing method.47 The paste for screen printing was prepared following the methodology previously described elsewhere.48 Briefly, 3 g of photocatalyst powder were thoroughly ground in a mortar containing 15 mL of ethanol, 2.5 mL of water and 0.5 mL of acetic acid, and then transferred to a round flask with the further addition of 50 mL of ethanol, 10 g of terpineol and 15 g of a 10 wt% ethyl cellulose solution in ethanol. The mixture was sonicated and stirred until the material was completely dispersed. Finally, the solvent was removed with a rotary evaporator. The films were manually deposited using a 200 mesh per inch polyester screen, followed by annealing at 500 °C for 30 min to remove organic components and increase their mechanical stability.
X-ray diffraction analysis (XRD) was performed with an XRD600 powder diffractometer (Shimadzu) operating at 40 kV and 30 mA employing Cu Kα radiation. Raman spectroscopy was conducted using a LabRAM HR Evolution spectrometer (Horiba). X-ray photoelectron (XPS) spectroscopy measurements were carried out using a Thermo Scientific model K-Alpha spectrometer, equipped with an Al Kα X-ray source. All spectra were corrected using the C 1s peak with a binding energy of 284.6 eV. Transmission Electron Microscopy (TEM) was conducted on a JEOL JEM 2100 High Resolution Transmission Electron Microscope operated at 200 kV and equipped with an Oxford Instruments X-Max 80T Energy Dispersive Spectrometer (EDS) system. Carbon support film-coated copper TEM grids (200 mesh) were used to support the samples. HRTEM images were taken using a Gatan Model 794 Slow Scan CCD Camera.
Film morphologies were evaluated by Scanning Electron Microscopy (SEM) using a TESCAN Vega3 microscope. Diffuse Reflectance Spectra (DRS) were obtained using a UV 2600 spectrophotometer (Shimadzu) equipped with an integrating sphere. The electrochemical characterization was carried out on an Autolab PGSTAT204 (Metrohm) potentiostat/galvanostat equipped with an Electrochemical Impedance Spectroscopy (EIS) FRA32M module. The irradiation source was a 300 W Xe arc lamp (Oriel) equipped with an AM 1.5 G filter; the light intensity was measured using a power meter 1916-R (Newport) connected to an optical sensor 818-UV/DB.
A PTFE customized H-shaped photoelectrochemical cell was used for all (photo)electrochemical experiments. The catholyte chamber was filled with 0.1 M K2SO4 electrolyte and the anolyte chamber with glycerol 10% v/v in 0.1 M K2SO4 or bare flegmass with added 0.1 M K2SO4. Platinum was used as the counter electrode and Ag/AgCl as the reference electrode. A Nafion® proton selective membrane was placed between the anolyte and catholyte chambers. Backside illumination was provided through a 3.8 cm2 quartz circular window placed at the front of the anolyte chamber.
The final composition of the anolyte was characterized by Chemical Oxygen Demand (COD), and the oxidation products were identified and quantified by High-Performance Liquid Chromatography (HPLC) and H1 Nuclear Magnetic Resonance (NMR) analysis. HPLC analysis was performed on a Waters e2695 system equipped with a DAD-RID detector and using an Aminex HPX-87H column (300 × 7.8 mm, 5 μm) under isocratic conditions. The mobile phase was a mixture of MeCN:
H2O 35
:
65 (30 min) and H2SO4 (0.005 mol L−1). A flow rate of 0.6 mL min−1 was employed, and the signals were detected at 254 nm. Quantification was made by calibration curves, obtained through the analysis of mixtures of chemical standards at known concentrations. H1 NMR analyses were carried out on a Bruker spectrometer model Ascend 400 MHz, and dimethyl sulfoxide was used as an internal standard as described elsewhere.5 All samples were filtered using PES 0.22 μm syringe filters. The gas products were identified and quantified using a Clarus 580 (PerkinElmer) GC-TCD equipped with a molecular sieve and a Porapak N packed column. Calibration curves, obtained by injecting known quantities of analytical grade H2, were used to quantify the green H2 produced during the photoreforming assays.
Electron paramagnetic resonance (EPR) spectroscopy was conducted on solid photocatalyst powders to examine residual paramagnetic defects. EPR, combined with the spin trapping method, was utilized to explore the reaction mechanisms involved in the photocatalytic tests. For this purpose, the spin traps 5,5-dimethyl-1-pyrroline N-oxide (DMPO) and N-tert-butyl-α-phenylnitrone (PBN) were dissolved in distilled water and in a water/methanol mixture, respectively, with each solution containing 10% analytical glycerol. Different photocatalysts were then added to these solutions under magnetic stirring, and the samples were illuminated with a white LED lamp (16 mW cm−2). At different time intervals, aliquots were taken, and EPR spectra were recorded. A similar procedure was performed using solutions without glycerol for comparison. Additionally, spin-trapping experiments were conducted using different BiVO4 photoanodes paired with a Pt cathode under an applied bias and white light illumination. All EPR spectra were acquired using a Magnettech MiniScope MS 400 (Germany) at room temperature.
Supplementary information includes additional structural, morphological, and electrochemical characterizations (XPS, SEM/TEM-EDS, EIS), as well as product analyses (1H NMR, HPLC) of crude glycerol and flegmass, supporting the findings of this study. See DOI: https://doi.org/10.1039/d5su00376h.
This journal is © The Royal Society of Chemistry 2025 |