Fitri Nur Indah
Sari
a,
Ping-Chang
Chuang
b,
Shih-Ching
Huang
a,
Chia-Yu
Lin
*ac and
Yi-Hsuan
Lai
*bc
aDepartment of Chemical Engineering, National Cheng Kung University, No. 1, University Road, Tainan City 701, Taiwan. E-mail: cyl44@mail.ncku.edu.tw
bDepartment of Materials Science and Engineering, National Cheng Kung University, No. 1, University Road, Tainan City 701, Taiwan. E-mail: yhlai@gs.ncku.edu.tw
cHierarchical Green-Energy Materials (Hi-GEM) Research Center, National Cheng Kung University, Tainan City 70101, Taiwan
First published on 21st July 2025
Rapid industrialisation has resulted in severe greenhouse gas emissions through extensive fossil fuel extraction and an increasing volume of solid waste disposal. This perspective review examines the photoelectrochemical (PEC) valorisation of organic waste as a promising solution for addressing the energy crisis and environmental pollution. The current stage of PEC organic valorisation has yet to meet industrial requirements, hindered by its relatively low efficiency, limited robustness, and poor scalability compared to conventional technologies. A better understanding of the working principle of the PEC reaction mechanism, the properties of state-of-the-art photoelectrodes, and the organic waste pre-treatment process is required to pave the path toward practical implementation. In this perspective review, we demonstrate that the strategies employed in the design of photoanodes, including doping, heterojunction formation, co-catalyst modification, nanostructuring, and crystal facet engineering, have different effects on activity, selectivity, and stability. In addition to the hydrogen evolution reaction, selected organic reduction reactions for the synthesis of value-added chemicals in a PEC cell are also discussed, followed by recent progress on integrated PEC cells and their practical assessment for solar fuel and value-added chemical production. Since stability and scalability are also essential parameters beyond efficiency for assessing the techno-economic feasibility of a PEC organic valorisation system, we additionally addressed stability, scalability and the compatibility of photoelectrodes in organic waste valorisation. Finally, conclusions and future perspectives on feasible strategies for PEC valorisation are discussed. We hope this review will serve as a helpful guide in designing effective, robust, and scalable PEC organic waste valorisation systems, making it a viable technology for real-world applications.
Photoelectrochemical (PEC) valorisation has been demonstrated to be one of the most promising technologies for addressing energy and environmental pollution issues. It supports carbon-neutral sustainability by converting organic waste into value-added chemicals using sustainable energy sources (Fig. 1).8–11 In particular, PEC reactions generally operate at room temperature under ambient conditions, thereby eliminating the substantial energy demands associated with the high temperatures and pressures typically required in conventional waste valorisation methods.12,13 In addition, PEC valorisation utilises solar energy, the most abundant renewable energy resource, to drive chemical reactions using organic waste as feedstock. Therefore, solar energy can be stored in the form of solar fuels in the co-production of value-added chemicals through PEC valorisation. Compared to water splitting, one of the most extensively studied reactions in PEC systems, organic waste valorisation offers an alternative pathway for generating more economically valuable products. For example, oxygen (O2) is the oxidation product in water splitting and has an economic value of only 0.04 USD per kg.8 In contrast, formic acid (FA), which can be generated from various organic oxidation processes, possesses an economic value approximately ten times higher than that of O2.14
![]() | ||
Fig. 1 Overview of PEC organic waste valorisation to drive value-added chemical and solar fuels production in a circular economy. |
Decreasing the energy demand in PEC reactions and enhancing the production rate of products are additional advantages of replacing water oxidation with organic oxidation. Water splitting consists of an oxygen evolution reaction (OER) and a hydrogen evolution reaction (HER):15,16
![]() | (1) |
![]() | (2) |
Overall reaction: 2H2O → O2 + 2H2 ΔG° = 237 kJ mol−1 | (3) |
On the other hand, replacing the OER with organic oxidation, i.e., coupling HER with organic oxidation, results in a thermodynamically favourable or marginally endothermic reaction (Table 1). For example, the HER coupled with glycerol valorisation to dihydroxyacetone (DHA) requires a marginal endothermic ΔG° of only 10.3 kJ mol−1. The potential required to oxidise glycerol to DHA is approximately at 0.05 V vs. the reversible hydrogen electrode (VRHE). Another example is ethylene glycol (EG), a monomeric unit of polyethylene terephthalate (PET), whose valorisation to FA coupled with the HER constitutes a thermodynamically favourable reaction. The potential required to oxidise EG to FA is located at a −0.25 VRHE (Table 1).
Overall reaction | Cathodic half-cell reaction | Anodic half-cell reaction |
---|---|---|
a Derived from ref. 6.
b Obtained by ![]() ![]() ![]() ![]() ![]() |
||
C6H6O3 (HMF) + 2H2O ↔ C6H4O5 (FDCA) + 3H2 | 6H+ + 6e− ↔ 3H2 | C6H6O3 (HMF) + 2H2O → C6H4O5 (FDCA) + 6H+ + 6e− |
aΔG° = 173.7 kJ mol−1, b![]() |
![]() |
a
![]() |
C2H6O2 (EG) + 2OH− ↔ 2 HCOO−(FA) + 3H2 | 6H2O + 6e− ↔ 3H2 + 6OH− | C2H6O2 (EG) + 8OH− → 2 HCOO−(FA) + 6H2O + 6e− |
cΔG° = −147.5 kJ mol−1, d![]() |
![]() |
b
![]() |
C2H6O2 (EG) + H2O ↔ C2H4O3 (GLA) + 2H2 | 4H2O + 4e− ↔ 2H2 + 4OH− | C2H6O2 (EG) + 4OH− → C2H4O3 (GLA) + 3H2O + 4e− |
cΔG° = 105.2 kJ mol−1, d![]() |
![]() |
b
![]() |
C3H8O3 (glycerol) ↔ C3H6O3 (DHA) + H2 | 2H+ + 2e− ↔ H2 | C3H8O3 (glycerol) → C3H6O3 (DHA) + 2H+ + 2e− |
cΔG° = 10.3 kJ mol−1, d![]() |
![]() |
b
![]() |
C3H8O3 (glycerol) ↔ C3H6O3 (GLAD) + H2 | 2H+ + 2e− ↔ H2 | C3H8O3 (glycerol) → C3H6O3 (GLAD) + 2H+ + 2e− |
cΔG° = 37.3 kJ mol−1, d![]() |
![]() |
b
![]() |
C3H8O3 (glycerol) + 3OH− ↔ 3 HCOO− (FA) + 4H2 | 8H2O + 8e− ↔ 4H2 + 8OH− | C3H8O3 (glycerol) + 11 OH− → 3 HCOO− (FA) + 8H2O + 8e− |
cΔG° = 65.6 kJ mol−1, d![]() |
![]() |
b
![]() |
C6H12O6 (glucose) + 6OH− ↔ 6 HCOO− (FA) + 6H2 | 12H2O + 12e− ↔ 6H2 + 12OH− | C6H12O6 (glucose) + 18OH− → 6 HCOO− (FA) + 12H2O + 12e− |
cΔG° = 47.9 kJ mol−1, d![]() |
![]() |
b
![]() |
C6H12O6 (glucose) + H2O ↔ C6H12O7 (GNA) + H2 | 2H+ + 2e− ↔ H2 | C6H12O6 (glucose) + H2O → C6H12O7 (GNA) + 2H+ + 2e− |
cΔG° = 70.9 kJ mol−1, d![]() |
![]() |
b
![]() |
C6H12O6 (glucose) + 2H2O ↔ C6H10O8 (GRA) + 3H2 | 6H+ + 6e− ↔ 3H2 | C6H12O6 (glucose) + 2H2O → C6H10O8 (GRA) + 6H+ + 6e− |
cΔG° = 179.1 kJ mol−1, d![]() |
![]() |
b
![]() |
Therefore, using a PEC system to drive a PEC organic valorisation reaction coupled with H2 evolution generally results in a higher product formation rate than using the same PEC system to drive water splitting. Fig. 2a and b show that the photogenerated holes can provide greater kinetic energy, i.e., a higher overpotential (ηOX), to drive organic oxidation, as organic oxidation requires less thermodynamic energy than the OER. Severe surface charge recombination caused by the sluggish kinetics of the OER would result in the onset potential for OER (Eon,OER) deviating significantly from the flat band potential of the photoanode (EF,n), as shown in Fig. 2c. In contrast, surface charge recombination can generally be mitigated by replacing the OER with a suitable organic oxidation reaction, in which the onset potential of organic valorisation (Eon,VR) can be negatively shifted toward the EF,n of the photoanode. Taking α-Fe2O3 as an example, the onset potential shifted from 1.1 VRHE to 0.85 VRHE by replacing the OER with the EG oxidation reaction,9 while it shifted from 1.0 VRHE to 0.7 VRHE by replacing the OER with the glucose oxidation reaction.18 The deviation range between the onset potential and the EF,n of the photoanode is crucial in determining the overall PEC cell performance. If an integrated PEC system comprising a photocathode and a photoanode is employed, the operating photocurrent (JOP) is determined by the intersection of the photocurrents of the two half-reactions (oxidation and reduction) (Fig. 2c). The difference between Eon,OER and Eon,VR for the same photoanode resulted in the JOP of PEC water splitting (JOP1) being lower than that of PEC organic valorisation coupled with the same reduction reaction (JOP2).
PEC organic waste valorisation may also provide an alternative approach to mitigate the instability of photoelectrodes in PEC water splitting. The thermodynamic stability of a photoelectrode depends on the relative thermodynamic potential of the semiconductor and the potential of the target redox reaction. Taking water oxidation as an example, if the thermodynamic oxidation potential (Eox) of a photoanode is more negative than the potential of the OER (1.23 VRHE), the photoanode would undergo photocorrosion by its photogenerated holes.19 Most non-oxide semiconductors, such as ZnS, CdS, ZnSe, etc., and some oxide semiconductors, such as ZnO, are susceptible to oxidation based on thermodynamic predictions.19 However, if these photoanodes are used for glycerol valorisation, for example, they are thermodynamically resistant to photocorrosion, as their Eox is more positive than the potential for glycerol valorisation to DHA (0.05 VRHE). Given these advantages, research on PEC organic waste valorisation has rapidly increased in recent years. This perspective review focuses on PEC organic waste valorisation. Organic waste substrates, including biomass, its derivatives, and plastics, are particularly highlighted in this review, since biomass derived from food and lignocellulosic sources accounts for the majority composition (∼60%) of solid waste. Although plastics represent a smaller fraction (12%) compared to plant-derived materials, their non-biodegradability poses a significantly detrimental effect on soil, water, marine life, and human health. Beyond the substrate used in PEC valorisation, variables such as solution pH, operation duration, and light intensity also influence PEC performance (Fig. 3). In particular, the properties of the photoelectrode represent the most crucial factor determining the upper limit of PEC valorisation system performance. Hence, in this perspective review, the working principle of a PEC reaction is introduced, followed by an overview of the materials and optical properties of the state-of-the-art photoelectrodes applied in PEC organic valorisation. Recent progress in designing efficient photoanode strategies for organic waste oxidation to generate value-added chemicals is subsequently discussed. Value-added chemical synthesis through reduction reactions using PEC cells is then covered, although the reported research in this area remains relatively limited compared to oxidation. Finally, recent developments in integrated PEC cells and their practical evaluation for solar fuel and value-added chemical production are reviewed, followed by a discussion of perspectives. Efficiency, stability, and scalability are the three key system requirements for the successful deployment of PEC systems.20 Long-term stability is an essential criterion for the practical implementation of PEC technologies. Without sustained performance over long-term operation, even highly efficient PEC devices cannot satisfy the economic, safety, and reliability standards required for commercialisation. Therefore, unlike most recent review papers, this work not only focuses on efficiencies but also addresses stability, scalability, and the compatibility of photoelectrodes and co-catalysts in integrated PEC cells. Finally, conclusions and future perspectives on feasible strategies for PEC valorisation are discussed. We hope this will guide readers in designing effective, robust, and scalable PEC systems for organic waste valorisation, particularly in the context of integrated PEC cells.
The performance of a photoelectrode in terms of photocurrent density (JPEC) can be described by eqn (4):
JPEC = Jmax × ηLHE × ηsep × ηint | (4) |
On the other hand, the EVB of the semiconductor should be more positive than the potential of an oxidation reaction, such as water oxidation or organic oxidation. This ensures that the holes created by light absorption possess sufficient oxidative power to oxidise water or organic waste.29–32 In PEC water splitting, the state-of-the-art photoanode includes TiO2,33–35 ZnO,35–37 BiVO4,38,39 WO3,38,40–43 α-Fe2O3.44–46 Those photoanodes have an EVB located at ≥2.5 VRHE (Fig. 5b), providing significant overpotential for water oxidation. As previously mentioned, the redox potential of most organic oxidations in organic valorisation research is less positive than that of the OER. Most of the photoanodes used in PEC water oxidation can be utilised and have been investigated as effective photoanodes in PEC organic oxidation. However, the photo-driven holes have a sufficient driving force for both the oxidation of water and organics. The selectivity of organic valorisation might remain an issue, as the parasitic reaction of the OER could occur.
The replacement of water oxidation with less energy-demanding organic oxidation, at least, partially alleviates the thermodynamic and kinetic constraints in selecting photoanodes. For example, nanosheet-structured WO3 lost 50% of its initial photoactivity during water splitting in a neutral solution within a few minutes.53 However, the nanosheet-structured WO3 does not lose 50% of its initial photoactivity until 12 h of operation in the glycerol valorisation reaction.54 The stability difference between water oxidation and glycerol valorisation on WO3 possibly resulted from difference in their reaction kinetics. Peroxo-species are readily formed during sluggish water oxidation, which leads to the corrosion of WO3. In contrast, peroxo-species are possibly eliminated during glycerol valorisation. Similar results have also been reported for BiVO4.55 However, the mechanisms by which photocorrosion is alleviated when replacing water oxidation with organic valorisation might vary among different photoanodes. Eliminating surface charge recombination may be another contributing factor to why a photoanode, such as BiVO4, typically exhibits better stability for organic (e.g., EG, glycerol) valorisation than for water oxidation. To date, PEC organic valorisation in the half-cell has achieved a maximum operation time of up to 150 h,56 and 80 h has been reported for an integrated cell.18 This highlights that the current progress in stability remains significantly below the criteria required for practical implementation and warrants further effort. The long-term stability of PEC valorisation based on recently reported studies is further discussed in a later section.
TiO2 has been widely investigated in PEC water splitting; however, the application of simple TiO2 in PEC organic valorisation is relatively limited. TiO2 generally exhibits low selectivity toward selected value-added chemicals in PEC organic valorisation (Table 2). Surface modification with a secondary material is generally required to achieve high selectivity. Inspiring examples include Bi2O3 heterojunction with TiO2 for glycerol valorisation to DHA,56 ultrathin BiOx-covered TiO2 converting succinic acid into C2H4,63 and Bi2O3-modified TiO2 converting cellulose into FA.64
Photoelectrode | Modification | Substrate | Spectrum and intensity of light | Supporting electrolyte | FE [%] of main product@E | J [mA cm−2]/φPEa@E | t 0.6 b [h]@E | Ref. |
---|---|---|---|---|---|---|---|---|
a φ PE is defined by the ratio of JPEC/Jmax, where j is the measured photocurrent and Jmax is the calculated theoretical photocurrent under 1 sun and AM 1.5 G, revealed in Fig. 5. b Defined as the time needed for the photocurrent to reach 60% of its respective initial photocurrent. c Operating at an external bias in a two-electrode system. d Operating at bias-free in a two-electrode system. | ||||||||
TiO 2 -based photoelectrode | ||||||||
TiO2 | N/A | Glycerol (0.1 M) | 385 nm, 10 mW cm−2 | Na2SO4 (0.5 M, pH 7) | FA: ≈33@1 VRHE | ≈2/N.A.@1.3 VRHE | N/A | 102 |
TiO2 | N/A | Cellulose (0.15 M) | λ < 500 nm | NaOH (2 M) | Malonic acid: ≈45 at 1 VRHE | ≈5/N.A.@0.5 VRHE | >3@0 VRHE | 103 |
TiO2 | Nanostructure | Glycerol (0.1 M) | AM 1.5 G, 100 mW cm−2 | Na2SO4 (1 M, pH 2) | GLAD: 32@1.23 VRHE | ≈0.9/0.84@0.5 VRHE | >10@1.23 VRHE | 104 |
TiO2/Bi2O3 | Heterojunction | Cellulose | AM 1.5 G, 100 mW cm−2 | NaOH (0.1 M) | FA: 83.9 ± 4.9 at 0.5 VRHE | ≈0.36/0.34@0.5 VRHE | >6@0.5 VRHE | 64 |
Bi2O3/TiO2 | Heterojunction/Co-catalyst | Glycerol (0.1 M) | AM 1.5 G, 100 mW cm−2 | Na2SO4 (0.5 M, pH 2) | DHA: 62.2@1.0 VRHE | ≈2.4/2.2@1.2 VRHE | N/A | 56 |
Ag@LDH@TiO2 | Co-catalyst | Glycerol (0.1 M) | AM 1.5 G, 100 mW cm−2 | Na2SO4 (0.5 M, pH 7) | DHA: 55@1.2 VRHE | ≈2.3/2.1@1.2 VRHE | >4@1.2 VRHE | 105 |
TiO2/nanoNiP | Co-catalyst | PET lysate | AM 1.5 G, 100 mW cm−2 | KOH (1 M, pH 14) | FA: 57.1 ± 1.7@c0.5 V | ≈0.6/0.57@1.2 VRHE | >2@0.5 VRHE | 106 |
![]() |
||||||||
WO 3 -based photoelectrode | ||||||||
WO3 | N/A | HMF (5 mM) | 100 mW cm−2 | NaPi (0.1 M, pH 4) | N/A | ≈1.5/N.A@1.2 VRHE | N/A | 67 |
WO3 | N/A | Cyclohexane (18 mL) | 100 mW cm−2 | tBuOH (12 mL) and HNO3 (2 mL) | KA oil: 76@c0.5 V | ≈2/N.A@1.2 VRHE | N/A | 65 |
WO3 | N/A | Glycerol (0.1 M) | 350 mW cm−2 | Na2SO4 (0.5 M, pH 5.5) | GLAD: ≈30, DHA: ≈20@1.2 VRHE | ≈0.35/N.A@1.2 VRHE | >5@1.2 VRHE | 66 |
WO3 | Nanostructure | Glucose (0.1 M) | AM 1.5 G, 100 mW cm−2 | NaCl (0.5 M, pH 4) | Gluconic acid: 20@1.23 VRHE | 6/1.5@1.2 VRHE | >20@1.23 VRHE | 107 |
m-WO3 | Crystal phase | HMF (5 mM) | AM 1.5 G, 100 mW cm−2 | NaBi (0.1 M, pH 4) | N/A | ≈0.9/0.23@1.2 VRHE | ≈1@1.1 VRHE | 68 |
WO3 (202) | Crystal facet | Glycerol (2 M) | AM 1.5 G, 100 mW cm−2 | Na2SO4 (0.1 M, pH 2) | GLAD: ≈75, DHA: ≈20@1.2 VRHE | ≈3.4/0.85@1.2 VRHE | >12@1.2 VRHE | 69 |
WO3 | Crystal facet | Methane | 100 mW cm−2 | Na2SO4 (0.1 M, pH 2) | EG: 23.9@1.3 VRHE | ≈3/N.A@1.2 VRHE | >12@0.9 VRHE | 108 |
WO3/TiO2 | Heterojunction | Glycerol (0.1 M) | AM 1.5 G, 100 mW cm−2 | Na2SO4 (0.5 M, pH 6) | GLAD&DHA: 70@1.2 VRHE | 2.89/0.72@1.23 VRHE | >1@1.23 VRHE | 109 |
N-WO3 | Dopant | Glycerol (0.1 M) | 365 nm | Na2SO4 (0.5 M, pH 1) | CO: 40@0.6 VRHE | 1.5/N.A.@0.7 VRHE | c>30 V | 110 |
![]() |
||||||||
BiVO 4 -based photoelectrode | ||||||||
BiVO4 | N/A | Glycerol (2 M) | AM 1.5 G, 100 mW cm−2 | Na2B4O7 (0.1 M, pH 9.4) | N/A | 1.3/0.17@1.2 VRHE | ≈7@0.7 VRHE | 76 |
BiVO4 | N/A | Lignin (5 mg mL−1) | λ > 400 nm,100 mW cm−2 | KHCO3 buffer (10 mM, pH 8.2) | N/A | ≈0.53/N/A@0.6 VAg/AgCl | N/A | 71 |
BiVO4 | N/A | HMF (5 mM) | AM 1.5 G, 100 mW cm−2 | Borate buffer (0.5 M, pH 9.2) containing 7.5 mM TEMPO | FDCA: 93–94@1.54 VRHE | N/A | N/A | 70 |
Mo:BiVO4/NiCo-LDH | Dopant/Co-catalyst | EG (1 M) | 100 mW cm−2 | KOH (0.1 M) | FA: ≈100@1.2 VRHE | ≈2.3/N.A.@1.2 VRHE | >3@1.0 VRHE | 111 |
Ta:BiVO4 | Dopant | Glycerol (0.1 M) | λ > 422 nm, AM 1.5 G, 100 mW cm−2 | H2SO4 (25 mM) | DHA: 61@1.0 VRHE | 3.07/0.4@1.23 VRHE | >2@1.0 VRHE | 112 |
NiOx (OH)y/W:BiVO4 | Co-catalyst | Glycerol (0.1 M) | AM 1.5 G, 100 mW cm−2 | KBi (0.5 M, pH 9.3) | FA: ≈30, DHA: ≈19@1.2 VRHE | 3.5/0.49@1.2 VRHE | >24@0.8 VRHE | 55 |
BiVO4 | Structural | Glycerol (0.1 M) | AM 1.5 G, 100 mW cm−2 | Na2SO4 (0.5 M, pH 2) | DHA: 29.1@1.23 VRHE | 6/0.8@1.23 VRHE | >4@1.23 VRHE | 74 |
BiVO4 | Nanostructure | Glycerol (0.1 M) | AM 1.5 G, 100 mW cm−2 | Na2SO4 (0.5 M, pH 2) | DHA: ≈30@0.6 VRHE | 3.7/0.49@1.2 VRHE | >5@1.0 VRHE | 75 |
Bi-rich BiVO4−x | Surface atom engineered | Glycerol (0.1 M) | AM 1.5 G, 100 mW cm−2 | Na2SO4 (0.5 M, pH 2) | DHA: ≈45@1.23 VRHE | 4.26/0.57@1.2 VRHE | >5@1.23 VRHE | 113 |
BiVO4(010) | Crystal facet | Glycerol (0.1 M) | AM 1.5 G, 100 mW cm−2 | Na2B4O7 (0.1 M, pH 2) | N/A | ≈1.2/0.16@1.2 VRHE | >2@1.1 VRHE | 114 |
Mo:BiVO4(001) | Crystal facet/dopant/nanostructure | Glycerol (0.5 M) | AM 1.5 G, 100 mW cm−2 | Na2SO4 (0.1 M, pH 2) | DHA: 23@1.23 VRHE | 7.45/0.99@1.23 VRHE | >100@0.8 VRHE | 115 |
![]() |
||||||||
Fe 2 O 3 -based photoelectrode | ||||||||
Fe2O3 | N/A | Cyclohexene (10 mM) | AM 1.5 G, 100 mW cm−2 | CH3CN (20% H2O) containing 0.1 M tetrabutylammonium tetrafluoroborate | Cyclohexene oxide: 35.2 ± 1.6@0.8 VAg/AgCl | ≈0.20/0.017@1.2 VAg/AgCl | >2@0.8 VAg/AgCl | 116 |
Fe2O3 | N/A | Glycerol (2 M) | AM 1.5 G, 100 mW cm−2 | NaOH (1 M) | FA: ≈ 46@1.1 VRHE | 2.01/0.17@1.23 VRHE | >36@1.23 VRHE | 117 |
Ni-Pi/α-Fe2O3 | Co-catalyst | PET hydrolysate | AM 1.5 G, 100 mW cm−2 | NaOH (1 M, pH 13.6) | FA: ≈ 87@1.1 VRHE | ≈2/0.17@1.2 VRHE | >12@1.1 VRHE | 9 |
NiOOH/α-Fe2O3 | Co-catalyst | Sawdust-derived sugar | AM 1.5 G, 100 mW cm−2 | KOH (1 M) | FA: ≈90@1 VRHE | ≈2/0.17@1.1 VRHE | >100@1.0 VRHE | 18 |
Fe2O3/Ni(OH)x | Co-catalyst | PET hydrolysate (50 mM) | N/A | KOH (1 M) | FA: ≈100@1.2 VRHE | ≈4/0.34@1.2 VRHE | N/A | 118 |
Fe2O3|m-CuO | Co-catalyst/heterojunction | Glucose (20 mM) | AM 1.5 G, 100 mW cm−2 | NaOH (0.1 M) | FA: 97.3 ± 2.8@1.0 VRHE | ≈0.8/0.07@1.2 VRHE | >2@1.0 VRHE | 119 |
Fe2O3 | N/A | Methanol (95% vol in water) | 365 nm, 100 mW cm−2 | NaOH (0.1 M) | Formaldehyde: ≈100%@1.1 VRHE | ≈0.7/N.A.@0.55 VAg/AgCl | N/A | 99 |
Ti-doped Fe2O3 | Dopant | Polyimide (≈60 mg mL−1) | AM 1.5 G, 100 mW cm−2 | KOH (1 M) | N/A | ≈1.5/0.13@1.2 VRHE | >4@1.2 VRHE | 120 |
![]() |
||||||||
Other photoelectrodes | ||||||||
Bi2WO6 | N/A | Glycerol (10% v/v) | λ > 350 nm, 120 mW cm−2 | K2SO4 (0.1 M, pH 6) | N/A | 0.69/0.22@1.3 VRHE | N/A | 121 |
CuWO4 | Nanostructure | Glycerol (1 M) | AM 1.5 G, 100 mW cm−2 | Pi (0.1 M, pH 7) | GLAD: 60 ± 6, DHA: 30 ± 1@1.23 VRHE | ≈0.6/0.07@1.23 VRHE | >38 h@1.23 VRHE | 54 |
CoNiFe-LDH/Ta3N5 | Nanostructure/Co-catalyst | Glycerol | AM 1.5 G, 100 mW cm−2 | NaOH (1.0 M) | FA: 60@d0 V | 3.59/0.29@1.23 VRHE | >0.5 h@d0 V | 122 |
Owing to its widespread application in the OER, the WO3 photoanode has also been intensively studied in PEC organic valorisation in recent years (Table 2). Examples include the conversion of cyclohexane to KA oil (a mixture of cyclohexanol and cyclohexanone),65 glycerol into value-added C3 products,66 and HMF to furandicarboxaldehyde (DFF) and FDCA.67 The crystal structure and predominantly exposed facets of WO3 also significantly affect the selectivity of PEC organic valorisation,68,69 which are also discussed in a later section.
Like TiO2 and WO3, BiVO4 has been intensively investigated for PEC organic valorisation. Pioneering work on BiVO4 for PEC biomass valorisation was reported by K.-S. Choi's group.70 In the presence of 2,2,6,6-tetramethylpiperidine 1-oxyl (TEMPO), a faradaic efficiency (FE) of over 90% toward FDCA from HMF is achievable using BiVO4 as the photoanode.70 Beyond the valorisation of HMF, BiVO4 has also been investigated for the PEC valorisation of lignin71,72 and methanol,73 and has been particularly well explored for glycerol valorisation (Table 2).55,74–76
α-Fe2O3 has also shown great potential and has attracted considerable attention in PEC organic valorisation. α-Fe2O3 has been investigated in glycerol, glucose, cyclohexene, methanol, and EG valorisation (Table 2). In addition, α-Fe2O3 is particularly suitable for polymer-waste valorisation. Because depolymerisation of plastics by alkaline hydrolysis is generally required prior to PEC valorisation, the high stability of α-Fe2O3 in strongly alkaline solutions makes it the most frequently applied photoanode for plastic valorisation.
As mentioned in the previous section, most work on PEC organic valorisation has focused on earth-abundant photoanodes such as TiO2, WO3, and α-Fe2O3. Given its extensive investigation in PEC reactions, BiVO4 is also reviewed here, though Bi and V are not highly earth-abundant elements. Other earth-abundant photoelectrodes beyond those mentioned above, including CuWO4,77–79 Ta3N5,80,81n-Si,82,83 ZnO,84 Cu2O,49,84 and CuBi2O4,85–87 have also been intensively studied for PEC water splitting; nevertheless, they have received significantly less, or no, attention in PEC valorisation (Table 2). For instance, CuWO4, an n-type semiconductor with a bandgap of 2.2 eV, is capable of utilising visible light up to approximately 550 nm to drive PEC oxidation reactions and exhibits chemical stability from acidic to slightly alkaline conditions.88,89 Only recently has CuWO4 been demonstrated as a versatile photoelectrode for converting glucose, fructose, and glycerol into several value-added chemicals, including FA, glycolate (the conjugate base of GLA), DHA, and GLAD.54
![]() | (5) |
However, in PEC organic waste valorisation, ηSTF values are generally low, regardless of Rproduct, due to the typically small ΔG. In particular, some organic waste reactions have a negative ΔG, which limits the applicability of ηSTF. Given the substantial variation in ΔG across different organic waste valorisation reactions, relying on ηSTF would result in inequitable comparisons among solar-driven valorisation systems. As an alternative, areal activity (Rareal) can serve as another indicator, providing a unified quantification of product generation per unit of illuminated area and reaction time.1
![]() | (6) |
To recognize the economic value of each type of organic waste valorisation process, the solar-to-value (STV) creation rate (rSTV) proposed by E. Reisner's group has been introduced:1
![]() | (7) |
However, the cost of PEC cells should also be considered and is not negligible, particularly when noble metals are used in the composition of the photoelectrodes or when the synthesis of photoelectrodes involves arduous procedures. Therefore, rSTV is further modified as follows:
![]() | (8) |
Although the ηSTF, Rareal, and rSTV parameters can be used to estimate the overall performance of an integrated PEC cell, additional indicators are necessary to evaluate half-cell performance. Since half-cell performance is predominantly governed by the intrinsic properties of the photoelectrodes, the inherent performance of a photoelectrode (φPE) using the following equation is suggested for the assessment of the half-cell performance of a PEC cell:
![]() | (9) |
The φPE allows for the assessment of the intrinsic performance of a photoelectrode by neglecting reaction substrates in a PEC system. It could be a promising indicator for comparing the performance of a PEC cell using the same photoelectrode for different organic valorisation reactions. The ηint generally deviates from unity during water oxidation due to its sluggish kinetics. However, ηint can be assumed to be unity, as most photoelectrodes provide significantly greater driving force than that required for water oxidation. Table 2 summarises the key parameters of reported PEC organic valorisation systems and categories of photoelectrodes based on their composition and modification methods. The φPE in Table 2 is calculated using eqn (9) and assuming ηint is unity.
The FE and power-saved metrics are advantageous for comparing the performance of different photoelectrodes under the same PEC organic valorising reaction. FE, defined as the percentage of passed charge used in the desired reaction, is an essential metric for evaluating selectivity in both EC and PEC systems. FE can be described by the following reaction:
![]() | (10) |
Some literature presents selectivity using the ratio of the amount of a desired product to the total amount of product or the ratio of the amount of a desired product to the amount of substrate. If there is an unknown product, it will lead to a deviation using the former method, while the selectivity can be enhanced within a specific reaction time by lowering the initial amount of substrate using the latter method. Therefore, these two types of selective presentations may not be sufficiently objective, and FE will be used primarily in the discussion about selectivity.
Power-saved metrics enable the evaluation of the effects of input solar illumination on a half-cell.90 In a three-electrode system, the power stored at a specific current density I (Psave(I), W m−2) is derived by the product of the current density I and the potential difference of driving a half-reaction at current density I between a selected working electrode in the dark and a photoelectrode in the light:
Psaved(I) = I(A m−2) × (Edark(I) − Elight(I)(V)) | (11) |
The Edark(I) and Elight(I) are the potential of driving a half-reaction at current density I of the dark electrode and photoelectrode, respectively.
![]() | ||
Fig. 6 (a) Linear sweep voltammetry (LSV) curves for EC oxidation using CoNi0.25P/NF and other electrocatalysts. Reproduced with permission under the terms of a CC BY 4.0 license.91 Copyright 2021, The Author(s), published by Springer Nature. (b) Photocurrent (solid lines) of BiVO4/NiCo-LDH before and after i–t measurements in an EG containing KOH solution. Reproduced with permission.92 Copyright 2024, John Wiley and Sons. |
The advantages of using a PEC system over an EC system can be highlighted by the work on glycerol valorisation in a membrane-separated continuous-flow PEC cell reported in 2024.95 In this work, direct PEC oxidation of glycerol paired with the dark H2 evolution or CO2 reduction was achieved using a Si photoanode coupled with a silver nanoparticle-coated carbon cloth cathode. On the other hand, in the EC system, a Ni sheet was used as the anode. Intense sunlight of up to 10 suns was applied to the PEC cell, and a photocurrent exceeding 110 mA cm−2 was achieved for glycerol oxidation (Fig. 7a). Compared to the EC cell, the PEC cell required less energy input from the external voltage source to attain the same current density (Fig. 7b). Approximately 1 V less cell voltage was needed for the PEC cell compared to the EC cell to generate a current density of 100 mA cm−2 (Fig. 7b). Notably, this study also demonstrated that the product distribution of glycerol oxidation in the PEC cell differs from that of the EC cell at the same current density (Fig. 7c). The FE for value-added products was higher in the PEC cell than in the EC cell (≈82% vs. ≈63%). Oxygen quantification experiments confirmed that the PEC system can effectively suppress the parasitic OER. This difference may arise from the inherent discrepancy in potential dependence between PEC and EC systems. The oxidative force in the EC system depends on the applied potential, whereas in the PEC system, it is determined and fixed by the EVB position and is independent of the applied potential.
![]() | ||
Fig. 7 (a) LSV of a Ni/Si photoanode under chopped light illumination and a Ni sheet-based electrode in the dark for glycerol oxidation. (b) Cell voltage difference between the Ni/Si photoanode and the Ni sheet-based anode at a current density of 100 mA cm−2 for glycerol oxidation. (c) FE of glycerol oxidation products after 30 min at a current density of 100 mA cm−2. Reproduced with permission under the terms of a CC BY 4.0 license.95 Copyright 2024, The Author(s), published by Springer Nature. |
Nonetheless, the PEC system undergoes complex charge transfer processes, beginning with the extraction of photogenerated charges, followed by hole transfer to the semiconductor surface, and subsequently to the organic reactant. Therefore, an effective photoelectrode design that facilitates rapid charge transfer is necessary for PEC organic valorisation.
The rate law analysis on α-Fe2O3 for PEC water oxidation was first demonstrated by Le Formal et al.96 Photoinduced absorption (PIA) spectroscopy combined with transient photocurrent (TPC) measurements was employed to elucidate the quantitative relationship between the reaction rate and the accumulation of photogenerated holes on the α-Fe2O3 surface.96 PIA enables monitoring of long-lived photogenerated charge carriers, i.e., surface hole density for the photoanode, while TPC monitors the net flux of holes transferred to the electrolyte. The reaction kinetics under steady-state PEC oxidation conditions was then estimated by rate law analysis (eqn (12)):96–98
J = kr[hole]β | (12) |
It was found that a transition occurs from a first-order reaction at low densities of accumulated surface holes (slow reaction) to a third-order reaction at high densities of accumulated surface holes (fast reaction), indicating that the accumulation of multiple holes is required to drive water oxidation efficiently over α-Fe2O3.96
In addition to PIA, the rate law analysis can also be conducted using EC impedance spectroscopy (EIS) measurement. Zhang et al. determined the surface hole density on the α-Fe2O3 photoanode using this EIS measurement and subsequently correlated it with the photocurrent density to calculate the reaction order of water oxidation.97 They found that the reaction order was approximately 1 and 2 for water oxidation in near-neutral and alkaline conditions, respectively, highlighting the dependency of the water oxidation reaction order on the solution pH.
Shifting the focus toward PEC organic valorisation, PIA coupled with TPC measurements has also been applied to derive the kinetic parameters and elucidate the mechanism of PEC alcohol oxidation.99,100 Mesa et al. found that methanol, when PEC oxidised to formaldehyde on α-Fe2O3 photoanodes, followed a second-order reaction with rate orders of 1.88 and 1.89 at applied potentials of 0 V vs. Ag/AgCl (VAg/AgCl) and 0.55 VAg/AgCl, respectively (Fig. 8a).99 They confirmed that the rate order on α-Fe2O3 is determined solely by the density of holes accumulated at the surface of the photoelectrode and is independent of the applied potential. This observation corresponds to the earlier discussion comparing EC and PEC systems. In contrast to EC systems (Section 2.3), in which the oxidative force varies with the applied potential, the energy, i.e., the oxidative force of the holes, in a PEC system is determined by the EVB of photoanodes.
![]() | ||
Fig. 8 (a) Rate law analyses showing the reaction order of PEC methanol oxidation on α-Fe2O3 at 0.55 VAg/AgCl (dark red) and 0.00 VAg/AgCl (light red), and on TiO2 at −0.80 VAg/AgCl (blue). Reproduced with permission under the terms of a CC BY 4.0 license.99 Copyright 2017, American Chemical Society. (b) Rate law analyses showing the reaction order of PEC water oxidation at 1.3 VRHE (blue), ethanol oxidation at 1.2 VRHE (black), and acetaldehyde oxidation at 1.1 VRHE (red) on α-Fe2O3. Reproduced with permission.100 Copyright 2022, John Wiley and Sons. |
It has been found that α-Fe2O3 exhibits highly selective aldehyde production through PEC alcohol oxidation.99,100 The underlying reason was also elucidated by operando spectroelectrochemical PIA coupled with TPC, in parallel with the Arrhenius analyses.100 Rate law analyses suggest that α-Fe2O3 exhibits a second-order reaction for ethanol oxidation, while the reaction order of PEC acetaldehyde oxidation on α-Fe2O3 is only 0.5 (Fig. 8b). Temperature-dependent rate law analyses further demonstrated that activation energy for ethanol oxidation by photogenerated holes in α-Fe2O3 is much lower than that for acetaldehyde oxidation (195 vs. 398 meV). These results indicate that the high selectivity for aldehyde production on α-Fe2O3 results from its much more facile kinetics toward ethanol oxidation compared to acetaldehyde oxidation, which suppresses the overoxidation of acetaldehyde. Compared to α-Fe2O3, TiO2 exhibits only a modest selectivity toward acetaldehyde production, and overoxidation of acetaldehyde was observed. The disparate selectivity toward acetaldehyde production from ethanol over different metal oxide photoelectrodes is correlated with their varying EVB.101 The measured acetaldehyde oxidation activation energies are 45 and 427 meV, respectively, for photogenerated holes in TiO2 and α-Fe2O3. The study highlights the significance of band potential in determining the selectivity of PEC organic valorisation.
Understanding reaction kinetics in PEC systems can guide the design of photoanodes for PEC organic valorisation, emphasising the importance of achieving a high density of accumulated holes and ensuring the suitability of the EVB for a photoanode. This can be achieved by designing photoelectrodes with a suitable electronic band structure, increasing the intensity of illumination using concentrated solar light, tuning hole densities through doping, modifying co-catalysts, nanostructuring, and creating heterojunctions. Strategies for enhancing the performance of photoanodes for organic valorisation is further discussed in Section 4.
(a) The first group, PE, PS, PP, and PVC, is composed of C–C bonds in the main chain and is synthesised via addition polymerisation. The high stability of these hydrocarbon chains limits their applicability in reforming reactions.
(b) The second group, comprising PET, PLA, and PU, features C–O or C–N bonds in the backbone chain and is synthesised via condensation polymerisation. Oxygenated plastics account for approximately 15% of global primary plastic production.131 Owing to their polarity and ester bonds, they are typically the target substrates in solar-driven plastic reforming.7,132
The suppression of charge recombination by enhancing the internal electric field has been recently demonstrated to boost the performance of a photoelectrode in organic valorisation. Kong et al. introduced nitrogen into WO3 (N-WO3) to enhance the PEC valorisation of polyol biomass into CO.110 Density functional theory (DFT) calculation revealed that nitrogen doping regulates the electronic structure polarity by creating an asymmetric distribution of charge (Fig. 9a). The asymmetric distribution of charge enhances the internal electric field, effectively preventing the recombination of charge carriers. As a result, N-WO3 exhibited a higher photocurrent density in PEC glycerol valorisation than pristine WO3 (Fig. 9b). N-WO3 presented a higher CO evolution rate than WO3 at a similar CO selectivity throughout the voltage range of 0.4–0.9 VRHE (Fig. 9c and d). N-WO3 also exhibited promising long-term stability in an acidic solution (0.5 M Na2SO4, pH 1), with around 80% selectivity towards CO, 40% CO FE, and 10% carbon yield after 50 h (Fig. 9e). The excellent stability was attributed to the absence of any phase or structural changes during the PEC process. Similarly, Kim et al. demonstrated an enhancement in PET microplastic valorisation over α-Fe2O3 by Zr doping (Zr:α-Fe2O3).146 PET microplastic was obtained by grinding commercial PET into microplastic (≤5 mm), followed by immersion in 5 M NaOH for 3 days at 80 °C to obtain EG. Zr doping served as an electron donor, accelerating electron transport kinetics and promoting more upward band bending. Zr:α-Fe2O3 exhibited a faster production rate of FA and acetate compared to α-Fe2O3. Moreover, Zr:α-Fe2O3 demonstrated excellent long-term stability over seven cycles, each lasting 12 h. Notably, there was no change in phase or surface oxidation state after the cycling test, and no metal leaching was observed, indicating excellent structural and chemical stability.
![]() | ||
Fig. 9 (a) Calculated isosurface of electron density and electrostatic potential of WO3 and N-WO3. (b) LSV curves of N-WO3 in the absence and presence of glycerol under dark or light irradiation. Evolution rate and CO selectivity of (c) N-WO3 and (d) WO3. (e) Long-term stability of PEC glycerol oxidation toward CO generation at 0.6 VRHE. Reproduced with permission.110 Copyright 2022, John Wiley and Sons. (f) Schematic of PEC glycerol oxidation over BiVO4 and Ta:BiVO4 in acidic medium. (g) Long-term stability of PEC glycerol oxidation over BiVO4, Ta:BiVO4, and TaOx/BiVO4 at 1.0 VRHE. Reproduced with permission.112 Copyright 2022, American Chemical Society. |
Beyond regulating the photocurrent of the pristine photoelectrode, increasing the selectivity for the target product and enhancing stability can also be realised by incorporating a suitable dopant. The enhancement of photocurrent, selectivity of DHA in PEC glycerol oxidation, and immunity to the dissolution in an acidic medium were achieved through Ta doping of BiVO4 (Ta:BiVO4), as shown in Fig. 9f and g.112 Doping with Ta in BiVO4 was found to effectively enhance the photocurrent density and FE of DHA from 1.70 to 3.07 mA cm−2 at 1.23 VRHE and from 41% to 61%, respectively. In terms of stability, BiVO4 retained only approximately 70% of the photocurrent in the acidic solution after 2 h of operation during glycerol oxidation. In contrast, both Ta:BiVO4 and TaOx-coated BiVO4 (TaOx/BiVO4) remained stable in photocurrent during 2 h PEC glycerol valorisation. These results confirm that the presence of Ta or TaOx on the surface of BiVO4 protects BiVO4 from dissolution in an acidic medium, thus providing high stability.
PEC organic waste valorisation is typically performed in water, and the OER inevitably occurs as a side reaction if the photo-generated holes possess sufficient oxidative force toward OER. As previously mentioned, organic oxidation in common valorisation studies generally requires less oxidative force than OER. Selecting a guest semiconductor with insufficient EVB to oxidise water can eliminate the OER side reaction during PEC organic compound valorisation. Chuang et al. constructed a type-II p–n heterojunction of nanoFe2O3 and CuO as a nanoFe2O3|CuO photoanode for PEC glucose oxidation into FA.119 With regard to the EVB position of nanoFe2O3, there is competition between OER and glucose oxidation. The incorporation of CuO effectively inhibits the OER side reaction because of the insufficient driving force of holes on its EVB. Additionally, CuO is an efficient catalyst for glucose valorisation to FA and also acts as a co-catalyst in the nanoFe2O3|CuO heterojunction photoelectrode, promoting FA formation (Fig. 10a). The nanoFe2O3|CuO outperformed nanoFe2O3 by exhibiting a stable photocurrent density, while nanoFe2O3 exhibited a 33% decay photocurrent density after 2 h of operation (Fig. 10b). Furthermore, nanoFe2O3|CuO also exhibited nearly 100% of FE for FA production at 1.0 VRHE (Fig. 10c). Another analogous case was presented by Luo et al., showing that the addition of a p-type semiconductor of Bi2O3 to n-type TiO2 (Bi2O3/TiO2) significantly enhanced charge separation, transfer, and light absorption, thereby promoting larger photocurrent density in the PEC glycerol oxidation.56 The Bi2O3/TiO2 exhibited a higher DHA selectivity than TiO2 alone (65.7% vs. 8.2%). In particular, DFT calculation also revealed that Bi2O3 tends to adsorb the middle hydroxyl of glycerol, selectively generating DHA rather than formic acid (Fig. 10d). Meanwhile, the Bi2O3/TiO2 maintained a high conversion rate (>200 mmol m−2 h−1) and FE of DHA for 150 h (Fig. 10e) due to its excellent structural and chemical stability.
![]() | ||
Fig. 10 (a) Band diagram of the nanoFe2O3|m-CuO. (b) CPE measurements of nanoFe2O3 and nanoFe2O3|m-CuO, and (c) their corresponding FE and production rate of FA in 0.1 M NaOH with 20 mM glucose under solar light irradiation (AM 1.5 G, 100 mW cm−2). Reproduced with permission.119 Copyright 2022, Royal Society of Chemistry. (d) Schematic of PEC glycerol oxidation over TiO2 and Bi2O3/TiO2 heterojunction and (e) long-term stability test of Bi2O3/TiO2 photoanode for PEC glycerol oxidation. Reproduced with permission.56 Copyright 2022, American Chemical Society. |
n–n heterojunctions have also been employed to promote the reaction rate for PEC organic valorisation. The formation of a type-II heterojunction between defective WO3 (m-H-WO3) and TiO2, coupled with interfacial and defect engineering, effectively accelerates charge separation and mass transfer.109 A photocurrent of 2.89 mA cm−2 was achieved at 1.23 VRHE for the heterojunction photoelectrode, which is more than 1.5 times higher than that of WO3. Likewise, the improved photocurrent stability of m-H-WO3/TiO2, 3 times higher than that of m-WO3, was demonstrated over 60 min. m-WO3 exhibited poor photocurrent stability, with a 60% decay occurring within only 35 min. The superior photocurrent stability in m-H-WO3/TiO2 was attributed to its massive oxygen vacancies, which acted as electron mediators. The production rate of GLYAD and DHA reached 353 mmol m−2 h−1 at 1.2 VRHE for the heterojunction photoelectrode, whereas WO3 achieved only 133 mmol m−2 h−1.
![]() | ||
Fig. 11 (a) High-resolution transition electron microscopy (HRTEM) image shows the surface modification of the BiVO4 photoanode with nanoFe:Ni-Bi. (b) LSV curves at a scan rate of 10 mV s−1 of (i and i′) the pristine BiVO4, (ii and ii′) BiVO4|nanoNi-Bi, and (iii and iii′) BiVO4|nanoFe:Ni-Bi photoanodes in near-neutral pH containing 0.1 M methanol in the dark (i′–iii′) and under light illumination (i–iii). (c) Production rate of formate (Rformate) and FEformate from PEC methanol oxidation using the pristine BiVO4 and BiVO4|nanoFe:Ni-Bi photoanodes. Reproduced with permission.154 Copyright 2020, Elsevier. (d) HRTEM image illustrating NiOx(OH)y co-catalyst incorporation on the surface of the W:BiVO4 electrode. (e) Production rate and (f) FE with total charge passed (Qinput) obtained in 0.5 M KBi. Reproduced with permission under the terms of a CC BY-NC 3.0 license.55 Copyright 2021, Royal Society of Chemistry. |
Co-catalyst modification has also been widely adopted for α-Fe2O3 in the valorisation of organic waste. For instance, CuO acts both as a heterojunction and co-catalyst in nanoFe2O3|CuO for PEC glucose valorisation, as discussed in the heterojunction section (Fig. 10a–c). Taking advantage of the corrosion-resistant properties of α-Fe2O3 in alkaline solutions, this material is of particular interest in polymer waste valorisation. However, the FE of some target products, such as FA, is often limited. Li et al. demonstrated that surface modification of Fe2O3 photoanode with Ni(OH)x co-catalyst significantly enhanced the FE of FA from PET valorisation, increasing it from 10% to nearly 100% (Fig. 12a–c).118 In this study, Ni(OH)x co-catalyst was loaded on the surface of α-Fe2O3 using a hydrothermal method under a mild temperature. Ni2+ in Ni(OH)x is proposed to be first oxidized to the higher valence of Ni3+ by the photo-generated holes from Fe2O3 and then reduced back to Ni2+ by oxidising EG, a monomer of PET, to FA. A similar result was observed with a Ni phosphate (Ni-Pi) co-catalyst modified α-Fe2O3 for EG and PET valorisation.9 Monomer EG was also obtained by grinding waste PET bottles into micro-sized powder, followed by alkaline hydrolysis in 1 M NaOH at 90 °C for 24 h. Ni phosphate (Ni-Pi) co-catalyst was deposited on α-Fe2O3 using a simple photo-deposition method (Fig. 12d–f), and the Ni-Pi-modified α-Fe2O3 (Ni-Pi/α-Fe2O3) photoanode demonstrated an enhancement of FE (87%) toward FA production from EG valorisation compared to unmodified α-Fe2O3 (40%). The experimental results confirmed that the enhanced PEC performance accounted for the enhancement in charge separation and charge transfer kinetics by Ni-Pi. This study also compared the performance of Ni-Pi-modified α-Fe2O3 with α-Fe2O3 modified by other co-catalysts. Ni-Pi demonstrated the highest activity and FE on α-Fe2O3 compared to CoOOH, NiOOH, and Co-Pi. DFT also revealed that Ni-Pi exhibits a stronger absorption ability toward the EG and promotes efficient desorption of the product of FA compared to NiOOH. Ni-Pi/α-Fe2O3 photoanode also successfully oxidised waste PET plastic bottle, achieving a high FA FE (82%) and functioned stably for 12 h (6 cycles) due to its structural and chemical stability.
![]() | ||
Fig. 12 (a) Schematic of the surface modification of the Fe2O3 photoanode with Ni(OH)x for selective oxidation of EG into FA. Enhancement of (b) photocurrent and (c) the yield and FE of FA with the charge passed using Fe2O3/Ni(OH)x photoanodes modified with different concentrations of Ni(OH)x. Reproduced with permission.118 Copyright 2022, American Chemical Society. (d) TEM image of the Ni-Pi/α-Fe2O3 photoanode. (e) LSV curves and (f) FE for α-Fe2O3 with different co-catalysts. FE was derived at 1.1 V for 2 h. Reproduced with permission.9 Copyright 2023, Elsevier. |
V2O5 has been demonstrated as an example of the effect of nanostructuring on the performance of methanol oxidation over V2O5 (Fig. 13a–d).10 Exfoliation of V2O5 into nanosheet bundles can be readily achieved by direct top-down exfoliation of sub-micron V2O5 plates using formamide as the exfoliation agent. The photocurrent for methanol oxidation doubled at 1.3 VRHE if the V2O5 microplate was exfoliated. An impedance study confirms that exfoliating V2O5 reduces charge transport resistance, owing to a shorter travel distance for photo-generated holes. In addition, the oxygen vacancy (VO) increases upon exfoliation, thereby enhancing the conductivity and active sites of V2O5.
![]() | ||
Fig. 13 SEM images of (a) microV2O5 and (b) exfoliated V2O5. (c) TEM image of the exfoliated V2O5. (d) Photocurrent density recorded at 1.3 VRHE of (i) microV2O5 and (ii) exfoliated V2O5 in 0.1 M Na2SO4 electrolyte containing 0.1 M methanol. Reproduced with permission under the terms of a CC BY 4.0 license.10 Copyright 2022, The Author(s), published by Elsevier. (e) SEM image of nanosheet structured CuWO4 (f) FE derived from PEC valorisation of glucose (0.8 M), fructose (0.8 M) and glycerol (1.0 M) at 1.23 VRHE at the variation of pH. Reproduced with permission.54 Copyright 2024, John Wiley and Sons. (g) Schematic of the effect of microscale fluid in the PEC glycerol valorisation. (h) LSV curves of the different pore size of TiO2 nanotube in Na2SO4 (0.5 M, pH = 2) with and without the addition of glycerol. (i) FE of the main products formed from glycerol oxidation using TiO2 nanotube photoanodes with different pore sizes. Reproduced with permission.104 Copyright 2024, American Chemical Society. |
As mentioned in the previous section, the majority of work in PEC organic valorisation focuses on state-of-the-art photoanodes such as TiO2, WO3, Fe2O3, and BiVO4. However, except for TiO2, the application of these photoelectrodes is limited to specific reactions due to the narrow operational range in which they can function stably. Most recently, nanosheet-structured CuWO4 (nanoCuWO4) has been investigated across a wide range of PEC valorisation systems, including glucose, fructose, and glycerol, owing to its broad operational stability (Fig. 13e and f).54 FE of 76% ± 5% for FA and 61% ± 8% for GLA were achieved at pH 10.2 over CuWO4 from PEC glucose and fructose valorisation, respectively. Notably, the FEs of the primary products, GLAD and DHA, were over 85% at neutral and slightly acidic pH using CuWO4 for PEC glycerol valorisation. NanoCuWO4 exhibited a significantly higher photocurrent at 1.23 VRHE than a planar CuWO4 for PEC water oxidation and glycerol valorisation, demonstrating that nanoCuWO4 provides more active sites in PEC reactions. Notably, the nanoCuWO4 maintained a high total FE of value-added chemicals (100%) over 72 h of photoelectrolysis and retained 60% of its initial photocurrent density after prolonged operation of more than 38 h. This excellent stability was attributed to the preservation of phase properties following long-term operation.
Tuning product selectivity has also been reported to be achieved through nanoconfinement. Lu et al. introduced a nanoconfined environment into the glycerol valorisation process using one-dimensional TiO2 nanotubes (Fig. 13g–i).104 Nanotubes with three pore sizes, 50 nm (TNT-50), 75 nm (TNT-75) and 100 nm (TNT-100), were investigated, among which TNT-50 exhibited the highest photocurrent for PEC glycerol oxidation. The nanoconfined environment accelerated the mass transfer process by the microscale fluid effects. The diffusion rate of glycerol on TiO2 nanotubes increased almost fivefold when the tube diameter was reduced from 100 to 50 nm. The faster diffusion rate inhibited the overoxidation of glycerol and enhanced the selectivity of C3 products, such as GLAD. The TiO2 nanotube also demonstrated promising stability, with no significant photocurrent degradation observed during 10 h of operation.
![]() | ||
Fig. 14 (a) Representation of the atomic structures of the lowest-energy ˙OH adsorption on twinning W atoms of different facets, in which, from left to right, the sequence is facets of (010), (100), and (001). The corresponding adsorption energy is presented at the bottom. (b) EPR analyses of ˙OH in the presence of WO3 NB, WO3 NP or WO3 NF after light illumination (c) PEC CH4-to-EG conversion selectivity at various potential using the photoanode of WO3 NB, WO3 NP or WO3 NF. Reproduced with permission.108 Copyright 2021, John Wiley and Sons. (d) Energy profiles derived from DFT calculation of glycerol oxidation on WO3 at the facets of (200) and (202). (e) The comparison of WO3 in OCP change between facets of (200) and (202) in a supporting electrolyte with subsequent injection of 0.1 M glycerol. (f) FT-IR absorbance spectra of the C–O bond vibration peaks of WO3 photoanodes with predominant facets of (200) and (202) in a Na2SO4 solution with or without glycerol. Reproduced with permission.69 Copyright 2022, American Chemical Society. (g) Schematic representation of the process of synthesising textured and untextured MBVO. (h) The ηsep and ηint (ηtrans) of textured and untextured MBVO. (i) The PEC stability of textured MBVO under glycerol oxidation and OER (inset) conditions. Reproduced with permission.115 Copyright 2024, John Wiley and Sons. |
Engineering the WO3 facet for optimising the generation of GLAD has also been reported (Fig. 14d–f).69 The production of GLAD originates from the two-electron oxidation of the primary hydroxyl group of glycerol. A WO3 photoanode with predominant (202) facets, denoted as WO3 (202), exhibited a higher production rate and selectivity for GLAD than WO3 with predominated (200) facets (WO3 (200)). This was attributed to the superior light absorption properties and higher carrier concentration of WO3 with predominated (202) facet. In addition, the DFT calculations revealed that the crystal facet of (202) showed stronger glycerol adsorption and facilitated its activation more effectively than the (200) facet. The (202) facet also exhibited easier desorption of GLAD. The stronger glycerol absorption tendency on WO3 (202) was evidenced by a more pronounced cathodic shift in open-circuit potential (OCP) after the addition of glycerol. Furthermore, Fourier-transform infrared (FT-IR) spectra indicated that, compared to WO3 (200), the C–O bond vibration peaks of the primary and secondary hydroxyl groups shifted to higher and lower wavenumbers, respectively, on WO3 (202). Both OCP and FT-IR studies experimentally demonstrated that WO3 (202) had a higher tendency for PEC GLAD production from glycerol. Notably, WO3 (202) also exhibited excellent photocurrent stability without any decay for 12 h, while the photocurrent density of WO3 (200) remained at approximately 70% of its initial photocurrent density at 1.2 VRHE.
In addition to the crystal facet effect in WO3, Wu et al. investigated the effect of the crystal phase of WO3, including monoclinic (m-WO3) and hexagonal (h-WO3), on HMF oxidation to DFF.68 m-WO3 outperformed h-WO3, attributed to enhanced light absorption, improved charge separation, optimised reactant adsorption and a higher oxidation capacity of the photo-generated holes. Consequently, m-WO3 exhibited a lower onset potential at 0.6 VRHE and a significant increase in photocurrent density, reaching 1.1 mA cm−2 at 1.3 VRHE (a sixfold enhancement over h-WO3). Under constant irradiation at 1.1 VRHE, m-WO3 retained 50% of its initial photocurrent.
Benefiting from the enhancement of reaction rate or selectivity by tuning the preferential crystal orientation has also been demonstrated in the case of BiVO4. Wu et al. synthesized textured Mo-doped BiVO4 (MBVO) with a (001) crystal orientation via a rapid-ramping annealing method and applied it for glycerol oxidation reaction (GOR) (Fig. 14g–i).115 The (001) crystal orientation in MBVO-textured provided a substantially enhanced ηsep of 90% at 1.23 VRHE, exceeding that of randomly oriented untextured MBVO (74%). Accordingly, the textured MBVO exhibited a photocurrent density of 7.45 ± 0.19 mA cm−2 at 1.23 VRHE, corresponding to 99% of the Jmax of BiVO4 (Table 2). Notably, the textured MBVO also showed promising stability of photocurrent density up to 100 h at 0.8 VRHE in GOR, originating from its high ηsep from (001) crystal orientation coupled with optimisation of photo-generated hole utilisation in glycerol oxidation. In another study, the difference in activity and selectivity toward DHA production from PEC glycerol valorisation between facets of (010) and (121) has been explored.114 Their FT-IR results indicated that (010) had a much stronger absorption signal for the C–O bond vibration peak of the secondary hydroxyl group than that of the facet (121). Although the DFT calculation aligned with the FT-IR result, there was only a subtle difference in the adsorption energy between facets of (010) and (121). This might result from the DFT used being the simplest one, without considering the solvent effect.
The above discussion on strategies related to the effect of achieving different levels of improvement on photoelectrode activity, target product selectivity, and PEC stability are summarised in Table 3.
Activity | Selectivity | Stability | |
---|---|---|---|
Doping | ●●○ | ●○○ | ●○○ |
Heterojunction | ●●○ | ●●○ | ●○○ |
Co-catalyst modification | ●●○ | ●●● | ●○○ |
Nanostructuring | ●●● | ●○○ | ○○○ |
Crystal facet and phase engineering | ●●○ | ●●● | ●○○ |
![]() | ||
Fig. 15 Schematic of (a) the PEC cell consisting of a BiVO4 photoanode and Ag cathode for HMF reduction into BHMF. Reproduced with permission.28 Copyright 2016, American Chemical Society. (b) Photo-electro-biochemical reactor with a TiO2 photoanode harvesting sunlight and a Co–N/CNT cathode to generate H2O2, which is provided to the biocatalyst for lignin valorisation. Reproduced with permission under the terms of a CC BY 4.0 license.155 Copyright 2019, The Author(s), published by Springer Nature. (c) Inorganic-biological hybrid PEC system with a Si/C photocathode for light harvesting and a LPMO catalysing α-chitin valorisation. (d) Long-term stability of the Si/C photocathode. Reproduced with permission.156 Copyright 2020, Elsevier. |
A PEC cell integrated with an enzyme, a biocatalyst, has been recently investigated for the valorisation of lignin and α-chitin (Fig. 15b). Myohwa Ko et al. demonstrated a PEC system consisting of a TiO2 photoanode for photovoltage generation and a Co–N/carbon nanotube (Co–N/CNT) cathode for oxygen reduction reaction.155 The cathodic product H2O2 acted as an electron acceptor for lignin peroxidase isozyme H8, which has high activity toward selective β-O-4 bonds in lignin cleavage. To prevent damage to the enzyme from high concentrations of H2O2, an additional cellulose membrane was used to separate the cathode and enzyme. The enzyme-assisted PEC catalyzed the lignin dimer depolymerisation with 93.7% conversion and 98.7% selectivity for 3,4-dimethoxybenzaldehyde, a vanillin derivative.
Generation of chitin oligosaccharides from α-chitin over a Si/C integrated with lytic polysaccharide monooxygenases enzyme mediator (LPMO) in the presence of a redox mediator is a rare example of using a photocathode for PEC reduction valorisation (Fig. 15c and d).156 The mediator, such as 2,6-dimethyl-1,4-benzoquinone (DMBQ), was first photoelectrochemically reduced by the Si/C photocathode. The reduced DMBQ subsequently provided the reductive power to LPMO, assisted by O2 activation, to convert α-chitin to oligosaccharide. The PEC system can be realised under benign environmental conditions for the sustainable valorisation of α-chitin. However, this system exhibited a photocurrent density decay of more than 60% after 12 h. Improving the stability of this system remains a subject for further investigation.
![]() | ||
Fig. 16 (a) Schematic of an integrated PEC cell comprising a nanoTiO2/nanoNiP photoanode paired with a CNT/nanoNiP cathode, and the chronoamperometric curve under light illumination of the PEC cell. (b) LSVs of (i) bare nanoTiO2, (ii) nanoTiO2|nanoNi-Pop(CV), (iii) Pt foil, and (iv) CNT/nanoNi-Pop. A solid line was recorded under light illumination in the deaerated PET lysate, whereas a dashed line was obtained in the dark in the deaerated KOH solution. Reproduced with permission under the terms of a CC BY 4.0 license.106 Copyright 2021, The Author(s), published by Elsevier. (c) Schematic representation of a photovoltaic device (Si solar cell) connected to the nanoCuWO4-based PEC cell for simultaneous glycerol valorisation and H2 production. (d) Overlaid J–V curves of the Si solar cell and nanoCuWO4-based PEC cell either for glycerol oxidation (green) or water oxidation (red). Solid and dash lines represent the J–V curves measured in the light and dark, respectively. (e) Amount of H2 generated by the nanoCuWO4-based PV-PEC cell from glycerol valorisation and water splitting. Reproduced with permission.54 Copyright 2024, John Wiley and Sons. |
As revealed in the last example, a PEC system might still need an external bias supply, although the required external bias is generally much less than in an EC system. A PV cell is typically further connected and stacked with a PEC cell to form an integrated, standalone PV-PEC system. For instance, a PV-PEC system has been applied to water splitting and glycerol valorisation over a CuWO4 photoanode (Fig. 16c–e).54 The PV-PEC cell exhibited a significantly higher operating photocurrent for glycerol valorisation than for water splitting. Product analyses also confirmed that the production rate of H2 in glycerol valorisation was approximately double that of water splitting. This corresponding finding supports that PEC valorisation can generate more valuable chemicals than water splitting and accelerate the H2 production rate.
![]() | ||
Fig. 17 Illustration of the Cu30Pd70|perovskite|Pt PEC system in (a) the two-compartment and (b) artificial leaf. PEC analyses using the two-compartment system or the artificial leaf in a solution containing pre-treated polymeric and real-world substrates in (c) chopped scans and (d) bias-free chronoamperometry. (e) Amount of oxidation product along with the corresponding selectivity generated from different substrates after 10 h of bias-free PEC reactions using the two-compartment system or the artificial leaf. Reproduced with permission under the terms of a CC BY 4.0 license.157 Copyright 2021, The Author(s), published by John Wiley and Sons. (f) Chopped scan and (g) bias-free chronoamperometry of the Ru@TiNS/Ni/perovskite photocathode in a two-electrode integrated system. (h) SAP, the amount of reduction product produced from under AM 1.5 G simulated one-sun condition per hour over an area of 1 cm2, of NH3 and NO2−, and (i) FE of the oxidation products from glycerol using the two-electrode system under bias-free condition. Ru@TiNS/Ni/perovskite and Pt@TiNS electrocatalysts were used as the photocathode and the anode in the two-electrode system, respectively. Reproduced with permission.159 Copyright 2024, Springer Nature. |
CO2RR, coupled with organic waste valorisation, was further demonstrated by the same group using a similar perovskite photocathode-based PEC system. A zero-bias PEC cell consisting of a CO2 reduction catalyst (CO2Rcat)-modified lead-halide perovskite (PVK) photocathode and a Ni foam|Cu27Pd73 microflower anode was used for PEC CO2RR and plastic valorisation.158 Three different CO2Rcat, including cobalt porphyrin (CoPL), Cu91In9 alloy, and tungsten-formate dehydrogenase (FDH) biocatalyst, were incorporated onto the perovskite surface, and their CO2RR performance was investigated. PEC reduction products at the photocathode part were CO (FE of 90%), syngas (CO/H2 = 1/1, FE of 43% (CO) and 49% (H2)), and FA (95%) using PVK|COPL, PVK|Cu91In9, and PVK|FDH photocathodes, respectively. On the other hand, the Cu27Pd73-based anode exhibited high selectivity with an FE exceeding 90% toward GLA from PET plastics. In addition, the PEC cells consisting of Ni foam|Cu27Pd73 anode and either a PVK|Cu91In9 or PVK|FDH photocathode also demonstrated stable photocurrent densities for up to 10 h. Meanwhile, a slight decay in the photocurrent density was observed when PVK|COPL was employed as the photocathode. This system is significant for developing high-value chemicals from the CO2RR platform, as conventional PEC CO2RR is limited by the high thermodynamic barrier associated with the OER.
PEC nitrate reduction (NO3RR) to form ammonia (NH3) has recently attracted considerable attention, as it generates green NH3 from nitrate-contaminated solutions. Furthermore, NH3 is a promising H2 carrier due to its safer and more efficient transportation. Tayyebi et al. recently demonstrated a bias-free PEC cell for co-generating NH3 and glyceric acid from NO3− and glycerol, respectively.159 The PEC cell was composed of a Ru-loaded titanate nanosheet loaded on a Ni-modified Pb-perovskite (Ru@TiNS/Ni/perovskite) photocathode and a Pt-decorated TiNS electrocatalyst, and could be operated bias-free. The bias-free PEC full cell achieved a stable photocurrent density of 21.2 ± 0.7 mA cm−2 over 24 h, which was attributed to the effective Ni coating that acted as a protecting layer over the perovskite photoelectrode, along with strong bonding between the Pt and TiNS at the anode (Fig. 17f and g). The reduction product was NH3, with a high yield of 1744.9 μg NH3 cm−2 h−1, corresponding to an FE of 99.5% ± 0.8%. Meanwhile, the major oxidation products were glyceric acid and lactic acid, which has a combined FE of 98.1% ± 2.4% (Fig. 17h and i).
![]() | ||
Fig. 18 (a) Schematic representation of the two-photoelectrode tandem PEC cell consisting of NiOOH/α-Fe2O3 paired with Bi/GaN/Si for FA production from biomass and CO2. (b) LSV curves of the tandem PEC cell with (BOR) and without (OER) the addition of biomass. (c) Chronoamperometric curve and corresponding productivity and FEs of FA from biomass and CO2 using the tandem PEC cell integrated with a Si solar cell under AM 1.5 G one-sun illumination. Reproduced with permission under the terms of a CC BY 4.0 license.18 Copyright 2023, The Author(s), published by Springer Nature. (d) Schematic of the two-photoelectrode tandem PEC cell for EGOR coupled with HER. PEC cell consists of a BiVO4-based photoanode (Mo:BiVO4/NiCo-LDH) and a Cu2O-based photocathode (Au/Cu2O/Ga2O3/TiO2/RuOx) and can be operated without bias. (e) LSV curves and (f) unbiased stability investigation along with the corresponding FE of FA and H2 of the BiVO4-based photoanode and the Cu2O-based photocathode placed behind the BiVO4-based photoanode under simulated AM 1.5 G illumination in buffer solution containing EG at pH 9.0. Reproduced with permission.111 Copyright 2024, John Wiley and Sons. |
A Cu2O photocathode paired with a BiVO4 photoanode in a two-electrode tandem configuration (Cu2O/BiVO4) has been demonstrated as a state-of-the-art PEC system for solar water splitting.160 The two-electrode Cu2O/BiVO4 tandem cell, with the assistance of a co-catalyst, generated a photocurrent of 1 mA cm−2 under one-sun illumination. However, the Cu2O/BiVO4 tandem cell exhibited significant instability for water splitting, losing the majority of its photocurrent within 1 min. Most recently, Kang et al. synthesized and demonstrated BiVO4/NiCo-LDH photoanodes for EG oxidation reaction (EGOR), subsequently coupled with Cu2O photocathode for H2 production (Fig. 18d–f).111 In the photoanode, NiCo-LDH significantly improved the photostability of BiVO4 by accepting the photogenerated holes from BiVO4 and thereby facilitating the formation of Ni3+/Co3+ active sites for EGOR. Furthermore, they also found that Ni(OH)x in NiCo-LDH controlled the selectivity toward EG oxidation into FA, whereas Co(OH)x enhanced the photostability by creating a protective layer and passivating surface defects to promote hole transfer to the electrolyte. The synergistic effect between BiVO4 and NiCo-LDH resulted in a high photocurrent density of 2.3 mA cm−2 at 1.23 VRHE, stable operation up to 200 min, and FEFA above 85% in 0.1 M KOH, which outperformed BiVO4, BiVO4/Ni(OH)x, and BiVO4/Co(OH)x. Additionally, an unassisted PEC device for EGOR and HER by coupling the Mo:BiVO4/NiCo-LDH photoanode with a Cu2O photocathode resulted in a maximum photocurrent density of 2.3 mA cm−2 and stable operation up to 2 h with FEFA of ∼60% and FEH2 of ∼100%. However, degradation began at 4 h, with photocurrent density decreasing to 1.7 mA cm−2 (74% of the initial photocurrent density), reducing FEFA by ∼40%, and finally stabilising at ∼30% after 8.3 h. This decline might be attributed to unfavourable product generation from the peroxidation of EG into CO2.
On the other hand, stability is a crucial parameter for assessing the performance and practicality of a PEC system. Nonetheless, stability continues to be neglected in PEC organic waste valorisation, although 100 h operation has been demonstrated in the studies of Mo:BiVO4(001) for PEC glycerol valorisation115 and NiOOH/α-Fe2O3 for sawdust-derived sugar solution.18 Strategies such as doping and heterojunction formation, widely applied in the current PEC valorisation studies, offer only modest enhancements to stability (Table 3). Other strategies to improve the stability of PEC systems, such as surface modification with a stable passivation layer, are suggested to be emphasised in future research. However, several requirements must be met to design a stable and efficient passivation layer for a photoelectrode. For example, the passivation layer should be thin enough to avoid charge transfer inhibition and parasitic light absorption,161 and the passivation process must be compatible with the photoelectrode.
In contrast to water splitting, which has an endless supply of substrate, i.e., water, the amount of substrate is limited by the solubility of organic waste in water for organic valorisation reaction. The effects of reactant consumption and mass transfer on product selectivity in organics valorisation must be carefully considered. Current research mainly performs the PEC valorisation reaction in a batch system. However, continuous reactant consumption and product accumulation in a batch system result in the unfavourable overoxidation of the target products. Examples can be easily found in the reported research.6,54 In addition, product accumulation might lead to altered solution pH and, thereby, affect product selectivity. Shifting the batch design to a flow reactor powered by a peristaltic pump, which can provide a continuous reactant supply, is necessary to the maintain stable production of the target product. Alternatively, the microfluidic effect could be applied to the reactor design to enhance mass transfer and, thereby, improve selectivity. As shown in Fig. 3, other system parameters beyond photoelectrodes, such as substrate concentration, solution pH, temperature, applied potential, and substrate absorption model, also contribute to the performance of a PEC organic valorisation reaction. Research into the optimisation and investigation of its mechanism is also essential to maximise the efficiency of a PEC system. Additionally, a techno-economic analysis helps assess the technical feasibility of PEC organic waste valorisation.
In conclusion, PEC valorisation of organic waste could be a promising route for not only waste mitigation but also the cogeneration of value-added chemicals and solar fuels. However, research into cost-effective pretreatment processes and the maximisation of the operational efficiency of robust PEC systems in a manner that allows for scalability is required for a commercially viable technology. To meet this demand, research directions are suggested to focus on investigating cost-effective catalysts for polymer waste depolymerisation in neutral or near-neutral conditions, integrating rational design strategies for the existing state-of-the-art photoelectrodes, developing compatible co-catalysts that can target C2+ products, and exploring novel photoelectrodes that meet high PEC performance and stability requirements. Additionally, integrating PEC oxidation with PEC reduction or cathodic reaction for value-added chemical production and advanced PEC reactor design could facilitate a circular carbon economy while addressing the energy crisis and environmental pollution.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5sc03146j |
This journal is © The Royal Society of Chemistry 2025 |