Swetha S. M. Bhat* and
Girish Kumar S.
Department of Chemistry and Centre of Excellence in Nanomaterials and Devices, RV College of Engineering, Bengaluru, Karnataka 560059, India. E-mail: swethasm@rvce.edu
First published on 15th August 2025
Photocatalytic reduction of CO2 into value-added products is witnessed as a promising technology to mitigate global warming and energy crisis. Among the illustrious functional semiconductors, bismuth-based materials are feasible for the CO2 reduction reaction owing to their convenient preparation, their narrow band gap and the availability of hybridized energy states in the conduction band. Furthermore, the inherent two dimensional (2D) layered structure of few Bi materials remains an added advantage for various photocatalytic reactions. In this focused review, the interfacial engineering of Bi-based semiconductors achieved by coupling them with distinct photocatalytic materials to form type-I, type-II, Z-scheme or S-scheme heterojunctions is discussed, and their applications in CO2 reduction reactions are emphasized. Further advancements, including co-catalyst loading, defect engineering and designing hierarchical morphology from the perspective of improving charge carrier separation and structural stability are highlighted. The preparation methods and mechanistic pathways for the CO2 reduction reaction are briefly summarized. Finally, the challenges and scope of Bi-based materials to spotlight their applications in energy- and environment-related areas are presented.
The primary step in semiconductor photocatalysis involves the absorption of light energy by catalysts with energy levels greater than or equal to those of their band gaps to generate electron–hole pairs. These excitons migrate towards the surface and interact with the adsorbed species to produce products via redox reactions, which later desorb from the surface of the catalyst.19 As per theoretical calculations and kinetic studies, CO, COOH and COH are common intermediates in the CO2 reduction reaction.20 CO2 photoreduction will proceed only when the energy of the conduction band of the concerned semiconductor is more negative than the reduction potential of the desired products.2 Semiconductors such as BiOX, CdS, g-C3N4, Cu2O, SnO2, TiO2, ZnO, and perovskites are extensively used for the CO2 reduction reaction.21–31 Bismuth-based materials have garnered major interest because they possess band gaps spanning the major portion of the solar spectrum, and their band edge positions are convenient to trigger the desired reduction and oxidation reactions.32–37 The Bi 6s orbital hybridizes with the O 2s orbital to form a highly dispersive energy band structure, which promotes charge carrier transport and also exhibits high oxidizing capabilities. Furthermore, bismuth-based materials offer stable structures with layered spaces to promote the intercalation of foreign ions without changing their structural features. Because of their low toxicity, environmental friendliness and exceptional photostability, the distinct morphologies of bismuth-based materials have been explored for photocatalytic applications.11,38 Different categories of bismuth-based materials such as unitary (Bi), binary (Bi2S3 and Bi2O3), and ternary (stoichiometric and non-stoichiometric bismuth oxy halides, BiVO4, Bi2WO6, and Bi2MoO6) are used for photocatalytic CO2 reduction.36,39–41 However, the massive charge carrier recombination and narrow optical response of single-phase photocatalysts limit their performance under light-illuminated conditions.39,42,43 Several reviews on bismuth-based materials for photocatalytic applications can be found in the literature.37,38,41,44,45 However, the recent trend in the interfacial engineering of these materials is scarcely emphasized. Motivated by these aspects, this review summarizes the interfacial engineering of bismuth-based materials with various semiconductors under different reaction conditions, and their performance in photocatalytic CO2 reduction is outlined.
(i) Type-II heterojunction: when two semiconductors with different band edge potentials and Fermi levels encounter each other, three types of heterojunctions, namely, type-I, type-II and type-III, are formed. Type-I allows for the migration of both the charge carriers from one semiconductor to another due to their straddle bandgap configuration, which enhances charge carrier recombination (Fig. 1). In contrast, type-II is efficient in separating the electron–hole pairs due to the staggered configuration of the electronic band positions. It is also referred to as the cascade electron transfer process, as it promotes interfacial charge carrier separation. However, the potential of electrons and holes decreases, lowering the efficiency of the composite.
(ii) Z-scheme: though type-II heterojunctions can improve the photocatalytic activity, they suffer from the recombination of electrons and holes. This disadvantage can be avoided by the Z-scheme technique.52
The band gap configuration in the Z-scheme is similar to the type-II configuration, but an exception remains in the direction of electron transfer pathways. The low-potential electrons accumulated on one side of the semiconductor recombine with the low-potential holes of another semiconductor, thereby retaining the high-potential electrons and holes for the redox reactions.
Although significant improvement can be observed from Z-scheme photocatalysis, this technique fails to deliver superior photocatalytic performance as the limitation arises from the choice of the semiconductors.53
(iii) S-scheme heterojunction: the step scheme (S-scheme) was introduced by the Jiaguo Yu group.54,55 When reduction and oxidation photocatalysts are in close contact, the electrons in the oxidation photocatalyst recombine with holes present in the valence band of the reduction photocatalyst. Positive and negative charges are formed at reduction and oxidation photocatalysts, respectively, and the direction of the in-built electric field is from the reduction photocatalyst to the oxidation photocatalyst. Thus, electron depletion at the reduction photocatalyst causes the upward band bending, and electron accumulation at the oxidation photocatalyst results in the downward band bending. Driven by the rise of the electric field and Coulombic force of attraction, the electrons in the oxidation photocatalyst recombine with holes in the reduction photocatalyst while preserving the energetic charge carriers for the desired redox reactions.55 The preserved electrons and holes are available for the reaction to take place, making the photocatalytic system more efficient than type-II and Z-scheme heterojunctions. The positions of the conduction and valence bands for different bismuth-based materials clearly signify their ability to participate in the CO2 reduction reactions (Fig. 2).
(iv) Cocatalyst loading: cocatalysts such as noble metals, which include Pt, Ag and Au, non-noble metals, such as Bi, Cu, and Mg, have been exploited to improve the performance of the photocatalytic process.56–58 The activation energy required to convert CO2 to any value-added product can be reduced by the addition of cocatalysts on the semiconductor surface. It is also found from the literature that it increases reaction sites and plays a vital role in determining the selectivity, adsorption of CO2, and photostability of the photocatalyst.59,60
(v) Defect engineering: defects are of paramount importance in tailoring the bandgap and charge carrier dynamics of functional semiconductors. In general, anionic vacancies like oxygen vacancies can narrow the gap region and shift the optical response of the host matrix to the visible region. Few reports also hint that the tailoring the bandgap depends on the density of oxygen vacancies in the host matrix.61,62 Furthermore, oxygen vacancies can trap charge carriers to restrain their recombination with holes and extend the charge carrier lifetime. On the contrary, cationic vacancies can temporarily trap holes. In recent times, the vacancy-rich Bi-based materials are reported to show superior photocatalytic activity compared to their low-defect counterparts.45,63,64 These defects can alter lattice strain and surface defects, which can facilitate the adsorption of oxygen to promote the formation of superoxide radicals in the solution phase.65
It is also noted that the cocatalyst loading on the photocatalyst is a potential approach to improve the adsorption of the intermediates during CO2 reduction, thereby influencing the coupling of the C–C bond. To improve C2 hydrocarbon selectivity and to enhance the photocatalytic reduction of CO2, transition metal cocatalysts with nanostructures were loaded on BiOCl.71,79,80 The selectivity towards CH4 was increased by decorating noble metals, such as Pd, on BiOCl. Pd not only decreased the recombination of the charge carriers but also changed the intermediate formed (from HCO3− to HCOO* and CH3O*; CH3O* will be reduced to CH4) on the surface of BiOCl.80
In situ diffused reflectance infrared Fourier transform (DRIFT) spectroscopy revealed that the key intermediates were bidentate formate and methoxy, which promote the selectivity for CH4 production.80 The photocatalytic activity of ultrathin BiOCl sheets can be altered by introducing magnetically active metal particles. In this context, Li et al. reported that Co-doped BiOCl selectively converted CO2 to CH4 and C2+ products, which was not possible with the pristine BiOCl. The increase in the photocatalytic activity was due to the Co–O bond on the Co–BiOCl.79 The increased CO2 adsorption was primarily responsible for the enhanced CO2 photoreduction. It was also found that the introduction of Co on BiOCl generated spin-polarized charges and changed the intermediate formed on the surface of BiOCl. In order to generate CH4 and C2+ products, the CO* intermediate would be formed via the formation of the COOH* intermediate (Fig. 3a–c). The rate-limiting step for Co–BiOCl was the conversion of CO* to CHO*, whereas for BiOCl, to
was the rate-limiting step.79 This work showed that the spin polarization of the charge on the surface of BiOCl resulted in the formation of a stable intermediate and promoted C–C coupling to yield CH4 and C2+ products. Similarly, Cu was modified on BiOCl nanosheets, which resulted in surface charge distribution. Cu–BiOCl improved the catalytic performance and enhanced the absorption of light.71 Also, it decreased the activation energy barrier, which facilitates the faster consumption of CO2 molecules, resulting in an increased CO yield (Fig. 3d and e). Metal cocatalysts, such as Pt, Pd and Au metals, have been decorated on BiOCl nanostructures by the deposition–precipitation method. The main product obtained with BiOCl was CO, while CH4 production dominated after loading the metal cocatalyst.81
![]() | ||
Fig. 3 Computed Gibbs free energy main reactions for the photocatalytic reduction of CO2 on a) pristine BiOCl and b) Co–BiOCl.79 c) CO2 reduction pathway on the surface of Co-BiOCl. d) Gibbs free energy diagram for CO2 reduction on Cu–BiOCl. e) Scheme for the photoreduction of CO2 to CO. Reproduced with permission from ref. 68. Copyright 2022, American Chemical Society and ref. 79. Copyright 2024, Elsevier. |
Chen et al. demonstrated that Z-scheme In2S3/BiOCl could be prepared by solution combustion synthesis. The nanopetals of BiOCl were decorated on the nanoflowers of In2S3 via solution combustion synthesis. In2S3/BiOCl nanostructures were evenly deposited on the surface of MnO2 nanowires.82 Initially, In2S3 nanoflowers were synthesized by the hydrothermal approach. The synthesized In2S3 nanoflowers were dissolved in thiourea, a bismuth nitrate precursor and concentrated nitric acid and stirred vigorously. To this solution, urea and ammonium chloride were added and stirred constantly for 24 h. The In2S3/MnO2/BiOCl was fabricated by using the suspension of In2S3/BiOCl by the hydrothermal approach. Despite having higher solubility than In2S3, BiOCl grew on the surface of In2S3.82 In2S3/BiOCl deposited uniformly on MnO2 nanowires, which formed an assembly of hierarchical structures. It is noted from this work that the S2− ions play a significant role in controlling the nucleation of In2S3, forming the morphology and controlling the size of the nanostructure. This ternary heterojunction exhibited photocatalytic reduction rates of 51.2 μmol g−1 h−1 for CO2 to CO, 63.2 μmol g−1 h−1 for CO2 to C2H4, and 42.4 μmol g−1 h−1 for CO2 to CH4 conversion.82 The observed synergistic effect was attributed to the Mn 3d electrons, which have narrowed the band gap. The key intermediates of the photocatalytic reduction were CHO* and COOH* for the production of CO, CH4 and C2H4.83 BiOCl with a BiOBr heterojunction was fabricated by the mechanical method, which yielded the type-I heterojunction.84 This heterojunction reduced CO2 to CO with a formation rate of 7.353 μmol g−1 h−1.
The CsPbBr3 perovskite is attracting increasing attention owing to its visible light harvesting capability, and its high conduction band position enhances the reduction capability.85 Various heterojunctions have been constructed based on the CsPbBr3 perovskite because the pristine material suffers from severe recombination of charge carriers. A 2D/2D S-scheme heterojunction comprising BiOCl and CsPbBr3 was fabricated by an efficient self-assembly method. The zeta potential measurements revealed the negative and positive surface charge densities on BiOCl and CsPbBr3, which could adhere to each other via electrostatic forces of attraction. The nanosheets of the perovskite were assembled on the nanosheets of BiOCl, which provided face-to-face contact, and a strong interfacial area existed between them. This heterojunction afforded much improved photocatalytic activity towards CO2 reduction to form CO as the major product and CH4 as a minor product.85 The high content of CsPbBr3 in the composite reduced the CO production, while the yield of CH4 remained almost the same irrespective of its content.85
Chen et al. synthesized BiOCl by the hydrothermal method, and the oxygen vacancy was introduced by heating at 60 °C in a vacuum oven.86 The Z-scheme heterostructure was fabricated by vigorously stirring powders of BiOCl with oxygen vacancies and 2D g-C3N4 nanosheets in an ultrasonic bath. It was found that the heterojunction could be constructed effectively if the two individual components were stirred in an ultrasonic bath rather than on a magnetic stirrer.86 The authors observed that the photocatalytic reduction of CO2 to CO on the interfacial oxygen-vacancy induced g-C3N4/BiOCl heterostructure was 1.6 times higher than that on the pristine g-C3N4/BiOCl.
Ternary composites of BiOCl with dual Z-schemes of BiOCl/g-C3N4/Ag2CrO4 have also been synthesized and found to give CO and CH4 with yields of 30.20 and 81.21 μmol g−1, respectively.87 It is possible to make a ternary heterojunction with C3N4 and BiOCl using the Ag2CrO4 photocatalyst. The dual Z-scheme heterojunction photocatalyst was synthesized by a simple method. BiOCl/C3N4 was prepared by taking the Bi, C and N precursors and heating solvothermally using ethylene glycol as the solvent for 12 h at 140 °C. However, the BiOCl/C3N4/Ag2CrO4 was synthesized by stirring the powders of BiOCl/C3N4 and precursors of Ag+ and Cr6+ ions for 2 h. When the ternary heterojunction was formed, the morphologies of BiOCl, C3N4 and Ag2CrO4 were retained. The irregular shapes of Ag2CrO4 were uniformly decorated on the microspheres of BiOCl and the nanosheets of C3N4. It was observed that C3N4 provided charge transfer pathways to form dual Z-scheme heterojunctions.
MgIn2S4 is another chalcogenide that exhibits a promising band gap and photostability and is extensively used for the removal of liquid pollutants and for the production of hydrogen.88 Zhang et al. fabricated an S-scheme heterojunction of Bi–BiOCl/MgIn2S4 by the hydrothermal technique and NaBH4 reduction technique.89 The microspheres of BiOCl with an average diameter of 1 μm were obtained by the hydrothermal method, and Bi self-doped BiOCl was synthesized by treating BiOCl microspheres with NaBH4, followed by vacuum drying. Bi–BiOCl/MgIn2S4 was synthesized by the in situ hydrothermal method, and the obtained product exhibited ultrathin sheets of BiOCl dispersed on the surface of MgIn2S4 marigold flowers. It was noted that NaBH4, which was used as a reducing agent, influenced the morphology of BiOCl and caused the rough surfaces and loose internal structures.89 The formed S-scheme heterojunction exhibited the conversion of CO2 to CH4 with a yield of 25.72 μmol g−1.
BiOBr exhibits an appropriate band structure, making it suitable for photocatalytic applications. However, the performance of BiOBr needs band-structure modification, as it promotes electron–hole recombination. The g-C3N4 has a band gap that falls in the visible light wavelength, and therefore, coupling it with BiOBr would help utilize a larger fraction of solar light.90 The hollow microspherical morphology of BiOBr was completely destroyed when grown on the surface of g-C3N4 sheets via the wet-chemical approach. The composite selectively produced CH3OH from CO2 at the optimized content of g-C3N4.90 Liu et al. synthesized BiOBr/C3N4 by ultrasonicating the synthesized BiOBr and C3N4 for 6 h.91 BiOBr was synthesised by the solvothermal technique, and the oxygen vacancy was introduced by adding the surfactant along with precursors. Furthermore, oxygen vacancies could be introduced to BiOBr by adding the surfactant during solvothermal synthesis, which is not possible during hydrothermal synthesis. The g-C3N4 heterojunction was constructed with BiOBr possessing surface oxygen vacancies. The morphology has been compared in Fig. 4.91 Along with the heterojunction interface, the presence of surface oxygen vacancies promoted the charge separation, and hence, the photocatalytic performance was enhanced.
![]() | ||
Fig. 4 Transmission electron microscopy (TEM) images of a) pure C3N4 (CN), b) CN–BiOBr and BiOBr (marked in the lower right corner), c) bare BiOBr–OV, and d) the CN–BiOBr–OV catalyst. e) and f) High resolution (HR)-TEM images of the CN–BiOBr–OV heterojunction with the lattice spacings, and g) elemental maps of C, N, O, Br, and Bi in the CN BiOBr–OV photocatalyst. Reproduced with permission from ref. 91. Copyright 2020, Wiley-VCH. |
Xi et al. also synthesized BiOBr/Bi2S3 and conducted a detailed analysis on the growth mechanism of Bi2S3 nanoarrays on BiOBr nanoplates.92 The facet-sensitive growth of Bi2S3 can be used as an effective strategy to construct a heterojunction with an increased number of active sites using thioacetamide as the sulfur source. It can also be observed that the structurally a very close matching of the lattice parameters between c-axis of Bi2S3 and a or b axis of BiOBr facilitates the selective growth of Bi2S3 facets.92 The established S-scheme heterointerface between BiOBr and Bi2S3 produced 20.32 μmol g−1 CO from CO2 photoreduction.92 The large area exposure of both semiconductors helped in exploiting the maximum amount of visible light. Directional growth of mesh like Bi2S3 nanostructure as grown on the surface of BiOBr nanoplates via topotactic transformation process found to be effective in enhancing the photocatalytic activity.92
Vertically aligned Bi2S3 nanowalls helped in the minimal shielding of the light falling on the surface of the BiOBr nanoplate substrate, which extended the optical response to 1000 nm (Fig. 5). The authors demonstrated that the vacancy-rich BiOBr substrate enhanced the adsorption of CO2 on the surface.92 Due to the weak Bi-O linkage, it was easy to produce oxygen vacancies in BiOBr. Because oxygen vacancies created trap centres for photogenerated charge carriers, it enhanced the separation efficiency. The rational design of oxygen-vacancy rich BiOBr heterojunctions would be beneficial in elevating the photocatalytic CO2 reduction.
![]() | ||
Fig. 5 Schematic illustrating the proposed mechanism for photocatalytic CO2 reduction over the BiOBr/Bi2S3 heterojunction. Reproduced with permission from ref. 92. Copyright 2022, Royal Society of Chemistry. |
Ma et al. suggested that instead of constructing conventional heterostructures, co-sharing atoms can be an effective strategy for interfacial engineering along with a suitable band structure. This work stresses the atom-level interfacial contact of the heterostructures.93 BiOBr was coupled with a Bi2S3 layered structure to deliver a full-spectrum responsive S-scheme photocatalyst. BiOBr/Bi2S3 was fabricated by an in situ ion exchange method, which yielded Bi2S3 nanorods on BiOBr nanosheets. It was observed that the morphology of BiOBr nanosheets changed after reacting with sulfur atoms during the hydrothermal reaction, which indicates that the occurrence of the ion exchange reaction led to the formation of the heterojunction. It was also noted that as the sulfur concentration increases, the density of Bi2S3 nanorods on BiOBr nanosheets. The BiOBr/Bi2S3 heterostructure yielded 20.32 μmol g−1 CO.
Bi2WO6 has a layered structure comprising fluorite-like [Bi2O2]2+ and perovskite-like [WO4]2−. Owing to the strong oxidising capability of its valence band, Bi2WO6 is extensively studied for photocatalytic applications. However, the performance is limited by the severe recombination of the photogenerated charge carriers. Therefore, a composite of Bi2WO6 and BiOBr would enhance the photocatalytic conversion of CO2 reduction.94 It is known from the literature that surface oxygen vacancies can easily be created on Bi2WO6 and BiOBr, which can provide an increased number of adsorption sites for the reactants and trap states for the photogenerated charge carriers.68,95 Apart from oxides and sulfides, layered double hydroxides (LDHs) have also been used for making composites with BiOBr. CoAl-LDH has high selectivity towards CH4, whereas NiAl-LDH yields CO.96 As CoAl-LDH has suitable band positions with BiOBr, facile electron–hole pair separation is possible. The heterojunction was fabricated by the hydrothermal method. The optimal amount of ultrathin CoAl-LDH loaded on BiOBr improved the selectivity towards CH4, a value-added chemical.96 In most of the cases, pristine BiOBr yields CO via photocatalytic conversion. In order to obtain value-added products such as CH4, the p-type semiconductor Cu2O was heterojunctioned with BiOBr.97 Recent studies have found that using metallic Bi is beneficial for enhancing the photogenerated charge separation.14,98,99 Many reports claim that Bi–BiOBr shows enhanced performance towards CO2 photoreduction.100 Abundant oxygen vacancies can be successfully synthesized by the in situ partial ion-exchange method on the heterojunction materials.74 More importantly, metallic Bi self-doped Bi2SiO5 synthesized by the hydrothermal method can provide a facile situation for S-scheme heterojunctions and thereby increase the redox capability in the S-scheme BiOBr-(001)/Bi2SiO5/Bi heterojunction, which gives a CO production rate of 234.05 μmol g−1 h−1.74
AgBr is another excellent photo-responsive material, but it suffers from instability due to the reduction of silver ions to metallic silver upon exposure to light. The low electron–hole recombination and extended visible light response were achieved in BiOBr/AgBr.101 Density functional theory and experimental results demonstrated that an electron depletion region is produced on BiOBr, whereas electrons accumulate on AgBr, resulting in an S-scheme heterojunction to produce CO and CH4. As the intensity of the light increases, the production of CO and CH4 increases linearly, proving that the reaction follows first-order kinetics.101 BiOBrxI1−x was coupled with BiPO4 to form a p–n junction, which was synthesized by the solvothermal method.102 This heterojunction exhibited increased photocatalytic activity compared to BiPO4.102 The 0D/2D CsPbBr3/BiOBr grown under different conditions selectively produced CO rather than CH4.103,104 The oxygen vacancies in BiOBr modulated the adsorption and activation of CO2, and the highly negative conduction band position of quantum dots (QDs) promoted the reduction reaction pathways. Recently, the coupling of BiOBr with p-type NiO was reported to be active for CO2 reduction to yield CO and CH4.105 It was proposed that the adsorption of CO2 on the composite surface formed carbonate species, which were converted to formic acid by capturing protons and electrons. The formic acid was successively transformed to formaldehyde, methanol and methane by continuous hydrogenation processes. Alternatively, CO2 was directly transformed into CO by hydrogenation and dehydration.105 The conduction band of NiO is mainly derived from the orbitals of Ni, which enables the accumulation of electrons at the Ni sites of NiO, which primarily serve as adsorption centres for CO2.105
Among BiOX materials, BiOI exhibits the narrowest band gap (∼1.6 to 1.9 eV), which enhances the absorption of light in the visible region. BiOI consists of alternative layers of [Bi2O2] and iodine held by van der Waals forces of attraction. BiOI, an indirect band gap semiconductor, was heterojunctioned with a perovskite material to improve the interfacial charge separation. The direct band gap semiconductor cesium formamidinium lead halide perovskite (Cs1−xFAxPbBr3) was coupled with BiOI nanosheets.106 It is worth noting that the heterojunction was created by stirring the solution of BiOI and perovskite nanocrystals owing to the low surface energy of perovskite nanocrystals and strong interactions between Bi cations and perovskite Br anion. The perovskite with a cuboid structure was well decorated on the ultrathin nanosheets of BiOI, as confirmed by SEM and TEM images. The authors demonstrated the kinetic models, as shown in Fig. 6, and they claim that BiOI had more surface defects than the perovskite Cs1−xFAxPbBr3 (CF).106 However, when BiOI and CF were heterojunctioned, the blue side of the photoinduced absorption band (PIA) allowed transition with Bi3+ doping in CF. An internal electric field (IEF) was created due to the electron transfer between BiOI and CF, which facilitates the recombination of electrons from the conduction band of BiOI and the holes from the valence band of the CF perovskite to form an S-scheme heterojunction, as depicted in Fig. 6.
![]() | ||
Fig. 6 (a) Depiction of kinetic models for (a) pristine BiOI and Cs1−xFAxPbBr3 (CF) and (b) hybrid Cs1−xFAxPbBr3![]() ![]() ![]() ![]() |
The formation of the S-scheme was confirmed by X-ray photoelectron spectroscopy. Femtosecond transient absorption spectroscopy was utilized to investigate the ultra-rapid dynamics of charge carriers.
In another work, BiOI nanosheets and In2O3 were synthesized individually by the solvothermal technique.107 Similar to the above work, a heterojunction was formed by mixing the as-synthesized semiconductors and heating solvothermally to obtain the heterojunction of BiOI/In2O3. The spherical morphology of In2O3 was decorated on the nanosheets. However, it was observed that the morphology was non-uniform throughout the heterojunction material. It can be inferred that the heterojunction fabricated via the solvothermal technique using the individual semiconductor materials resulted in the aggregation of In2O3 on BiOI nanosheets.107 The BiOI/C3N4 constructed on hydrophobic carbon fiber paper via the electrophoretic deposition process exhibited significant selectivity for CO production during the CO2 reduction reaction.108 Initially, pre-formed g-C3N4 sheets and an iodine powder were dispersed in the acetone solvent, wherein protons were released upon the reaction of iodine with acetone. Later, BiOI/carbon fibers and Pt electrodes were used as the cathode and anode, respectively, which was followed by the application of an external potential.108 The BiOI-nanosheets were vertically grown on the surface of carbon fibers, and g-C3N4 sheets completely wrapped the fibers to form 2D/2D contacts between the semiconducting surfaces.108
Li et al. carried out interfacial engineering with the formation of the p–n junction by compositing BiOI and Zn2TiO4.109 The presence of oxygen vacancies sped up the CO2 reduction process.110 They extended the lifetime of the charge carrier and also acted as activation sites to enhance the adsorption of CO2. Wang et al. fabricated a 2D/2D Bi2MoO6/BiOI S-scheme heterojunction to enhance the photoreduction of CO2.110 BiOI nanosheets were grown on few-layered Bi2MoO6 nanosheets using the solvothermal technique. The S-scheme mechanism was investigated by utilizing time-resolved photoluminescence spectroscopy, work function and charge density difference analyses. The nanosheets of Bi2MoO6 on BiOI nanospheres demonstrated that the van der Waals heterojunction is effective in accelerating the photogenerated charge carriers. The authors established that the charge transfer takes place from Bi2MoO6 to BiOI in addition to large-area van der Waals heterojunctions and S-scheme heterojunctions.110
Hongyu Fu et al. synthesized the heterostructure of Cu2O and carbon-loaded BiOI using two-step methods.111 The yield of methanol and ethanol from CO2 was 722.8 μmol g−1 and 264.46 μmol g−1 for 8 h respectively. The photocatalytic performance of C–BiOI was significantly increased compared to pristine BiOI. The heterostructure of Cu2O and carbon-loaded BiOI further improved the photocatalytic activity. Cu2O is a p-type semiconductor with a band gap of 2.2 eV. Nanoparticles of Cu2O were decorated onto the nanosheets of C–BiOI by the chemical deposition method. The introduction of Cu2O remarkably increased the photogenerated charge carriers and the surface area of the heterostructure, which provided more active sites for CO2 reduction.111
BiOI-based 2D–2D heterojunction materials have been designed in order to elevate the photocatalytic performance. The band structures of C3N4 and BiOI are favourable to form S-heterojunctions, which can boost charge separation.71 Li et al. proposed S-scheme charge transport, which increases the redox capability of both BiOI and C3N4 and converts CO2 to CO.112 The formation of the S-scheme heterojunction facilitates the electrons to move from C3N4 to BiOI, creating a depletion region in C3N4, whereas electron accumulation occurs in BiOI. In order to develop wide-wavelength-responding photocatalysts, the authors interfaced In2O3 with BiOI, because In2O3 exhibits a band gap of 2.8 eV, which falls in the visible region. It is also known to possess good electronic conductivity and is resistant to corrosion. The type-II heterojunction was fabricated by the solvothermal method.107 In2O3/BiOI produced CO and CH4.72
2D-BiVO4 has also been heterojunctioned with perovskites 2D-CsPbBr3 to achieve high efficiency in CO formation without any co-catalysts or sacrificial agents (Fig. 7).122 The preparation step involved the heterogeneous nucleation of CsPbBr3-sheets on the surface of BiVO4 via the annealing step in an inert atmosphere, which outlines the stability of BiVO4 as the substrate surface. The presence of oxygen vacancies in BiVO4 achieved a gradient Fermi level shift and enlarged the Fermi level gap between the semiconductors. This process was accompanied by an enhanced interfacial electric field in the heterojunction, which consecutively promoted the charge carrier separation process. It is interesting to note that pure CsPbBr3 had a quasi-square morphology with random sizes, while uniform nanosheets were grown on the BiVO4 substrate surface, which further emphasize the pivotal role of the substrate surface in altering the morphological features during the nucleation process.122
![]() | ||
Fig. 7 Schematic of the synthesis, structural modification and band modulation of the BiVO4/CsPbBr3 heterojunction. Reprinted with permission from ref. 122. Copyright 2022, Elsevier. |
Wang et al. further explored the synthesis of the Bi2S3/BiVO4 heterostructure, which was obtained from the in situ selective ion exchange method. The authors loaded the cocatalyst MnOx on the heterojunction photocatalysts to improve the selectivity.60 In this work, the photooxidation deposition method was adopted to decorate the MnOx facet selectively. BiVO4 was synthesized by the hydrothermal method, and Bi2S3 nanosheets were preferentially grown on (110) of BiVO4. Bi2S3 was grown on (110) of BiVO4 by sulfurization. As the concentration of S2− ions increased, the surface of BiVO4 turned rough as VO3− ions were replaced by S2− because the solubility of Bi2S3 was more than BiVO4. It is worth noting that if the reaction time was less than 90 minutes, the crystallinity of Bi2S3 was poor and structurally unstable. Therefore, effective interfacial contact would be unsuccessful in the short duration of the reaction.60 However, as the time duration increased to 120 minutes, the reaction would promote non-selective epitaxial growth. The sponge-like structure of MnOx grew selectively on the (110) facets of BiVO4 rather than on the (010) facets of Bi2S3 nanosheets.60 The photocatalytic reduction of CO2 to CH3OH was observed in this heterostructure with a production rate of 20 ± 2.33 μmol g−1 h−1.
The type-II heterojunction of Bi4Ti3O4 with BiVO4 was fabricated through a simple electrospinning technique and a solvothermal method.123 Bi4Ti3O4 with BiVO4 formed a type-II heterojunction, which yielded various intermediates to give CO and CH4OH (Fig. 8). A direct S-scheme heterojunction was formed between BiVO4 and TiO2, which changed the charge separation. Platinum decoration was carried out to enhance the CO2 photoreduction. Platinum decoration on the surface of these catalysts helped in converting CO2 to CH4 by mitigating electron–hole pair recombination.124
![]() | ||
Fig. 8 (a) CH3OH and CO evolutions on BiVO4/Bi4Ti3O12 with different Bi4Ti3O12 contents and (b) photocatalytic activity stability of the BiVO4/10% Bi4Ti3O12 sample. Reprinted with permission from ref. 123. Copyright 2020, Elsevier. |
An S-scheme heterojunction was fabricated by growing BiVO4 on the surface of a metal–organic framework PCN-224 (Cu).125 This photocatalyst exhibited 100% CO selectivity. Liu et al. decorated Co–Pi, a cocatalyst, on BiVO4/SnO2 and coupled it with an Au cathode to achieve better conversion efficiency for CO2.126 The Z-schematic pathway for charge carriers was constructed using SrTiO3:
Rh and BiVO4.127 In this work, a powder of the photocatalyst was used without adjusting the pH. It is worth noting here that the conversion of CO2 to CO was progressed without any addition of additives.
NiO, with a narrow band gap of ∼2.5 eV, which can efficiently absorb visible light, was composited with Bi2WO6, which gave an S-scheme charge transfer. The heterojunction yielded CO and CH4 with 46.9 and 12.1 μmol g−1, respectively, with the CO* intermediate.135 Lulu Zhao et al. attempted to improve the selectivity of the catalyst towards CO from CO2 by substituting Bi3+ and Ag+ ions via the liquid-phase exchange process.136 The incorporation of Ag+ not only improved the adsorption of CO2 and H2O but also enhanced the separation of the charge carriers.136 Ultrathin Bi2WO6 nanosheets with Ag+ and Bi3+ exhibited the conversion of CO2 to CO with a yield of 116 μmol g−1 and with a selectivity of 95.7% after 6 h of the reaction without using any sacrificial agents. In situ DRIFT suggested that the intermediate formed was COOH*, which is then converted to thermodynamically stable CO*. It was possible to obtain methane via CO2 photoreduction on Bi2WO6 by modifying the surface with surface plasmon resonance. Precisely controlling the oxygen vacancy in Bi–O–Bi and W–O–W led to high selectivity towards CH4. The vacancy created an energy level that is very close to the conduction band, which helped in improving the life span of the charged carriers. To further improve the charge transport pathway, Bi2WO6 was covalently heterostructured with Bi2O3. The Bi2O3 nanosheets present on the Bi2WO6 surface rendered more active sites, which enhanced the adsorption of CO2 on the surface.
Graphdiyne (GDY) is explored as a potential material for solar energy harvesting since its first synthesis.137,138 An ultrathin heterojunction of GDY and Bi2WO6 was prepared by a simple hydrothermal method because GDY is an excellent material for efficient charge transfer.139 Also, the increased surface area helped improve the CO2 adsorption on the photocatalyst. GDY/Bi2WO6 yielded 2.13 and 0.23 μmol h−1 g−1 of CH3OH and CH4, respectively. The reduction of CO2 yielded CH3OH and CH4 (Fig. 9).139
![]() | ||
Fig. 9 (a) Photocatalytic CO2 reduction performance of samples. (b) Cycling tests of the GDY/Bi2WO6 heterojunction (BiGDY). Reproduced with permission from ref. 139. Copyright 2021, American Chemical Society. |
Further, to enhance the light absorption ability of the material, the hierarchical hollow structured heterojunction of Bi2WO6/TiO2 was constructed by the in situ synthetic technique. This heterojunction interface generated the Bi(3−x)+ active site, which could suppress the charge carrier recombination very effectively. In producing CO via CO2 photoreduction, this heterojunction was better than that of pristine Bi2WO6 and Bi2WO6/TiO2.27 Similarly, Bi2WO6 was heterojunctioned with other semiconductors such as InVO4, La2Ti2O7, and ZnV2O6 either to improve the light absorption ability or to suppress the recombination rate of electron–hole pairs.140–142
Qiaoya Tang et al. constructed an S-scheme heterojunction of C3N4/Bi2WO6 using the electrostatic self-assembly method.143 The authors claimed that the formed ultrathin 2D/2D C3N4/Bi2WO6 heterostructure offers abundant reaction sites, which promote efficient charge transfer. Yunpeng Liu et al. introduced hydrophobic and hydrophilic ends in the diphasic photocatalyst to solve the competitive adsorption issue of CO2 and H2O. This amphipathic diphasic hydrophobic and hydrophilic photocatalyst not only inhibited the recombination of charge carriers but also could enrich CO2 and H2O on hydrophobic and hydrophilic surfaces, respectively, as depicted in Fig. 10.144
![]() | ||
Fig. 10 (a) Schematic of the hydrophobic Bi2WO6–hydrophilic C3N4 photocatalyst and the distribution of CO2 and H2O. (b) Schematic of the preparation of a phospholipid-mimicking photocatalyst. (c) TEM image and (d) AFM image of C3N4. (e) TEM image and AFM image (inset) of Bi2WO6–CN. (f and g) HR-TEM images of Bi2WO6–C3N4. (h) EDS elemental mapping of Bi2WO6–C3N4. Reproduced with permission from ref. 144. Copyright 2024, American Chemical Society. |
In an effort to improve the yield of CH4 in CO2 photoreduction, Yan-Yang Li et al. modified Bi2WO6 with chloride ions to study the influence of protons produced during the water oxidation on CH4 generation.145 The density functional theory confirmed the mappings of Bi2WO6–C3N4. The chloride ions present on the surface of Bi2WO6 nanosheets not only promoted water oxidation but also favoured the formation of the CHO* intermediate, which facilitated the formation of CH4. Due to the presence of chloride ions on the surface, the rate of the oxidation half-reaction goes smoothly, which indirectly promotes the reduction half-reaction (CO2 reduction).145 As 2D/2D nanosheets provide a large specific surface area and rich active sites, Yong Jiang et al. fabricated the heterojunction of CsPbBr3/Bi2WO6.146 This enabled Z-scheme charge transfer and reduced charge carrier recombination to obtain the yields of 1582.0 μmol g−1 for CO and 8602 μmol g−1 of CH4.146
Black phosphorous (BP) is a 2D material that is considered a potential candidate for photocatalytic applications because it provides a short diffusion length for the charge carriers. Because BP has a large specific area, it offers ample reaction sites for CO2 adsorption. Therefore, BP has been used for constructing heterojunctions with Bi2WO6 because it is believed to accelerate the multistep electron process and dimerization of the C–C bond. Minghui Zhang et al. fabricated 2D/2D BP/Bi2WO6 by the electrostatic assembly method.147 Organic oxidation to value-added products was coupled with CO2 reduction, which lowered the thermodynamic barrier for water oxidation. In this regard, benzylamine, which is a derivative of the renewable biomass, was used for organic oxidation to give imines as a high-value product. The adsorption study revealed that Bi2WO6 with BP exhibited more active sites for CO2 adsorption. In addition to this, benzylamine provided alkaline media, which increased the solubility of CO2 in the solution. In situ Fourier transform infrared (FTIR) suggested that CO* and were the intermediates for the conversion of CO2 to C2H5OH.136 To harvest visible light absorption and to improve charge transfer, g-C3N4 was introduced on Bi2WO6. The photocatalytic performance was further enhanced by incorporating reduced graphene oxide (rGO).143,148 In another report, ultrathin 2D/2D Bi2WO6/g-C3N4 offered abundant contact interfaces with more accessible reaction sites that exhibited higher selectivity for CO generation compared to CH4. The higher content of Bi2WO6 hampered the performance of the composite, as it could shield the incident photons striking the catalyst surface and hinder the bandgap excitation process. Furthermore, the composite retained its performance even after four consecutive cycles.143 These 2D/2D heterojunctions with face-to-face interfaces could lower the intrinsic resistance and boost the charge carrier separation process.149
The deposition of QDs like Cs3Bi2Br9 on the Bi2WO6-nanosheets could selectively promote the CO2 reduction reactions to form CO under visible light.134 This unique 0D/2D (dot-to-face) geometry and lead-free bismuth halide perovskite semiconductor remained as an added advantage for this composite. The larger surface area of 2D sheets enabled the finer distribution of QDs on their surface, which not only resulted in a stable heterojunction but also lowered the distance for the charge carrier migration process. The preparation method was flexible as it only involved the dispersion of positively charged QDs on the negatively charged Bi2WO6 sheets in isopropanol suspension through the electrostatic self-assembly approach. This strategy benefits from the prospect of retaining the pristine morphology of the concerned semiconductors during the heterojunction formation. The ESR analysis revealed the generation of both hydroxyl and superoxide radicals, which further confirmed the formation of the S-scheme heterojunction between them.134
Dai et al. grew Bi2MoO6 QDs on rGO scaffolds using an economically viable hydrothermal method.152 Due to the high surface area of rGO, the dispersion of Bi2MoO6 QDs was enhanced. The electron migration from Bi2MoO6 QDs to rGO was facile because rGO exhibited excellent electrical conductivity. The photocatalytic mechanism is illustrated in Fig. 11.
![]() | ||
Fig. 11 Unit cell model of Bi2MoO6 and electron transfer. Reproduced with permission. Reprinted with permission from ref. 152. Copyright 2020, American Chemical Society. |
Yu et al. fabricated a hierarchical heterostructure of 2D nanosheets of Bi2MoO6 grown on 1D In2S3 using the solvothermal technique. The hollow nanotubes of the In2S3 nanostructure were designed using the MIL-68 precursor. The surface oxygen vacancy was created by using N,N,N′,N′-tetramethylethylenediamine (TMEDA). However, the morphology of Bi2MoO6 nanosheets was distorted to give a particle-like morphology when TMEDA was used to introduce surface oxygen vacancies.153 The presence of the interfacial Mo–S bond in Bi2MoO6/ZnIn2S4 boosted the production of CO from CO2 reduction reactions.151 The in situ loading of ZnIn2S4 nanoflakes on the porous microspheres of Bi2MoO6 under the solvothermal conditions offered plentiful catalytic sites for CO2 adsorption and activation processes. Furthermore, the hierarchical structures enhanced the visible light absorption and possessed good photostability and reusability. The generation of superoxide and hydroxyl radicals, as confirmed from the ESR analysis, validated the formation of the S-scheme heterojunction between them.151 The oxygen-vacancy engineered Bi2MoO6−x/MoS2 was active for CO2 reduction to form CO, wherein oxygen vacancies were obtained from pristine Bi2MoO6 via a simple annealing step. It is interesting to note that the oxygen vacancies did not disrupt the microsphere-like morphology of the host matrix, and the density of oxygen vacancies increased upon composite formation with MoS2-nanosheets. The presence of these oxygen vacancies lowered the bandgap of the host matrix and acted as effective electron sink to delay the recombination of charge carriers and also reduced the work function of the host matrix, which indirectly decreased the energy required for the electron transfer process.154
Maryam Ahmadi et al. heterojunctioned Bi2MoO6 with the benchmark photocatalyst TiO2 to reduce the rate of recombination of photogenerated electron–hole pairs.155 The synergistic effect was achieved by compositing Bi2MoO6 with TiO2, as the band gap of both materials achieved a staggered configuration. Bi2MoO6 nanosheets were decorated with TiO2 nanobelts by a simple solvothermal method. As cocatalysts acted as electron scavengers, Pt–Cu was decorated on the composite by the reduction method. It was noticed that the photoresponse of the photocatalyst increased, and 34.6 μmol g−1 of methane was obtained for Bi2MoO6/TiO2. Once the Pt–Cu were loaded on the composite, methane production was increased further as the cocatalyst helped trap the electrons and reduce the recombination rate.155
Other metal oxide semiconductor materials have also been heterojunctioned such as a spinel compound ZnFe2O4, which produced CO and CH4.156 The S-scheme pathway helped in a slight improvement of the CO2 reduction; the catalyst needs to be improved to achieve better photocatalytic performance. The 2D/2D Bi2MoO6/Zn3V2O8, which formed an S-scheme, produced CO and CH4.157
Nanosheets, belts, particles and spheres of Bi2MoO6 have been synthesized, but reports on the QDs of Bi2MoO6 are uncommon.158–160 Weili Dai et al. anchored QDs of Bi2MoO6 on 2D rGO sheets, which act as an electron reservoir, helping boost CO2 reduction. QDs Bi2MoO6/2D rGO produced 84.8 μmol g−1 of methanol and 57.5 μmol g−1 of ethanol, which are the highest among bismuth-based photocatalysts in the reduction of CO2.152 Further, the QDs of Bi2MoO6 were deposited on pomelo-peel derived carbon (CPP) to improve the photocatalytic property of the catalyst. The CPP possesses a high specific surface area and has excellent thermal conductivity. As the photothermal temperature increases, it not only enhances the migration rate of the electrons and holes but also promotes the adsorption of CO2. Mingnv Guo et al. synthesized 0D/3D CPP-A/Bi2MoO6 to achieve a photothermal synergy to effectively utilize the solar spectrum.161 As the temperature increases up to 100 °C, the migration ability is increased, and a C–O–Bi bridge is formed, which enhances the electron transfer at the interface. This work provides a solution for the bottleneck problem of poor charge carrier transport in photothermal catalytic activity.161
By combining both n-type semiconductors In2S3 and Bi2MoO6, an S-scheme was achieved to design an efficient photocatalyst. 2D nanosheets of Bi2MoO6 grown on 1D nanotubes of In2S3 reduced the carrier diffusion length. The Z-scheme heterojunctions such as Bi2MoO6/CdS and Bi2MoO6/CeO2 have also been constructed to improve the charge carrier transport.162,163 An S-scheme heterostructure can also be achieved from ZnIn2S4 decorated on Bi2MoO6, where the Mo–S bond at the interface helps in improving the photocatalytic activity.151 S-scheme heterostructures can also be achieved from chalcogenides, especially with MnS. Diethylenetriamine (DETA) ammonia treated MnS hollow sphere heterostructured with plasmonic Bi enhanced CO2 reduction rate of Bi2MoO6.164 The presence of basic functional groups on DETA reduced the activation energy of acidic CO2, whereas the surface plasmon effect from Bi produced hot electrons and decreased the recombination rate of charge carriers. This combined effect helped boost the CO2 photocatalytic reduction reaction.
Table 1 illustrates the different interfacial engineering of bismuth-based materials with various other compounds and their respective CO2 conversion products. The synthesis method and charge carrier pathway followed by the heterojunction are also mentioned. It is interesting to note that most of the bismuth-based materials are fabricated by wet chemical approaches such as hydrothermal and solvothermal methods. However, these techniques pose several challenges, as the materials' purity and morphology are sensitive to the synthesis temperature and reaction time. Therefore, the reproducibility of the structural–morphological features would be difficult for the large-scale synthesis.165,166
Bismuth-based material | Preparation method | Nature of the heterojunction | Products: production rate in μmol g−1 h−1 | Ref. |
---|---|---|---|---|
Note: readers are requested to follow the respective references for more information. | ||||
BiOCl/BiOBr | Hydrothermal | Type-I | CO: 7.353 | 84 |
Temperature – 160 °C | ||||
Time – 12 h | ||||
Bi2WO6/TiO2 | Solvothermal | Type-II | CO: 43.7 | 27 |
BiOCl/Bi2WO6 | Hydrothermal | Type-II | CO: 6.63 | 76 |
Temperature – 120 °C | ||||
Time – 12 h | ||||
BiOBr/Bi2S3 | Hydrothermal | Type-II | CO: 103![]() |
92 |
Temperature – 180 °C | ||||
Time – 2 h | ||||
BiOBr/CoAl LDH | Hydrothermal | Type-II | CO: 4.096 | 96 |
Temperature – 160 °C | CH4: 4.174 | |||
Time – 12 h | ||||
BiOI/In2O3 | Hydrothermal method | Type-II | CO: 11.98 | 107 |
Temperature – 180 °C | CH4: 5.69 | |||
Time – 12 h | ||||
BiVO4/Bi4Ti3O12 | Hydrothermal | Type-II | CH3OH: 16.6 | 123 |
Temperature – 160 °C | CO: 13.29 | |||
Time – 21 h | ||||
BiVO4/Bi2S3/MnO3 | Selective epitaxial growth | Z-scheme | CH3OH: 20 ± 2.33 | 60 |
Temperature – 180 °C | ||||
Time – 90 min | ||||
BiOBr/HNb3O8 | Self-assembly ultrasonic dispersion for 30 min | Z-scheme | CO: 164.6 | 167 |
g-C3N4/BiOI/RGO | Hydrothermal method | Z-scheme | CO: 21.85 | 168 |
Temperature – 160 °C | ||||
Time – 6 h | ||||
Co-MOF/Bi2MoO6 | Solvothermal, followed by the in situ growth method | Z-scheme | CO: 19.76 | 169 |
Temperature – 165 °C | CH4: 8.24 | |||
Time – 16 h | ||||
BiVO4/Cu2O/Bi | Solvothermal | Z-scheme | CH4: 1.8 | 170 |
Temperature – 120 °C | CO: 8.4 | |||
Time – 6 h | ||||
In2O3/Bi2S3 | Hydrothermal, followed by ultrasonication | Z-scheme | CO: 2.67 | 171 |
Temperature – 180 °C | ||||
Time – 24 h | ||||
Quinacridone/BiVO4 | Self-assembly method | Z-scheme | CO: 407 | 172 |
Stirring for 1 h | CH4: 29 | |||
BiOCl/C3N4 | Hydrothermal | Z-scheme | CO: 45.33 | 173 |
Temperature – 120 °C | ||||
Time – 12 h | ||||
BiOBr-(001)/Bi2SiO5/Bi | Solvothermal method | S-scheme | CO: 234.05 | 74 |
Temperature – 190 °C | ||||
Time – 12 h | ||||
BiOBr/Bi2WO6 | Hydrothermal | S-scheme | CO: 55.17 | 94 |
Temperature – 160 °C | ||||
Time – 6 h | ||||
BiOBr/Cu2O | Hydrothermal | S-scheme | CH4: 22.78 | 97 |
Temperature – 160 °C | ||||
Time – 12 h | ||||
BiOBr/CsPbBr3 | Self-assembly process | S-scheme | CO: 104.4 | 103 |
Overnight magnetic stirring in the dark | CH4: 10.0 | |||
BiVO4/CsPbBr3 | In situ colloidal growth method | S-scheme | CO: 103.5 | 122 |
Cs3Bi2Br9/porous BiOCl | Dipping method | S-scheme | CO: 25.5 | 174 |
Magnetic stirring for 1 h | ||||
BiVO4/Cu–Bi | Solvothermal | Oxygen vacancy | CO: 2.96 | 56 |
Temperature – 120 °C | ||||
Time – 6 h | ||||
BiOCl/Br–OV | Solvothermal | Defects | CO: 7.37 | 175 |
Temperature – 160 °C | ||||
Time – 10 h |
Other bismuth containing materials, such as Bi12O17Br2/g-C3N4, Bi12O17Cl2/g-C3N4, Bi3NbO7/g-C3N4, SrBi4Ti4O15/Bi2O3, Cs2CuBr4/Bi2O3, BiOIO3/CdS and Bi/Bi4V2O11, are explored for CO2 photocatalytic conversion.176–183 In particular, the presence of interfacial Bi–N bonds and the nitrogen vacancy in Bi12O17Br2/g-C3N4 achieved simultaneous tetracycline degradation and CO2 reduction reactions.177 In fact, the photocatalytic reduction of CO2 was improved in the presence of tetracycline compared to rhodamine B and phenol molecules, which was attributed to the superior electron-donating ability of TC. The flower-like Bi12O17Br2 were decorated on the defective g-C3N4 surface via the wet chemical approach. The presence of defects was important to drive the coupled photocatalytic reactions. It was proposed that *CO2 was transformed to *COOH through a hydrogenation step, which upon further protonation yielded *CO, which was feasible for the desorption from the composite surface. This work is a classical illustration of win-to-win strategy to achieve both environmental and energy related issues.177 Alternatively, Bi12O17Cl2/g-C3N4 promoted the formation of CH4 via two-electron and two-proton reaction pathways.181 On the contrary, Bi3NbO7/g-C3N4 promoted the formation of CH4 as the reduction pathway proceeded through the formation of CH3O* and CHO* fragments.178 Lee et al. reported that CdS-nanorods/BiOIO3-nanosheets (1D/2D) selectively generated CO and minor amounts of H2 during the CO2 reduction reactions, and the activity was retained even after four consecutive cycles.182 The line interfacial contact between the integrated materials promoted the charge carrier lifetime with minimal recombination pathways.184 A significant improvement in photocatalytic performance was noticed, yet these materials must be explored further to enhance the conversion of CO2 to value-added products.
This review summarizes the interfacial engineering carried out for bismuth-based materials in recent years for CO2 reduction. A large number of reported works have demonstrated that interfacial engineering helps effectively enhance the charge separation, light absorption and adsorption of CO2 on the surface of the photocatalyst. However, the conversion efficiency is far from the efficiency required to implement in practical applications. Most of the interfacially engineered bismuth-based photocatalysts exhibit a rate of reduction of CO2 less than 100 μmol g−1. It is also noticed from the previous literature that the product obtained from the photocatalytic reduction reaction is carbon monoxide. Though bismuth based materials show promising results in photoelectrochemical water splitting, the photocatalytic activity towards CO2 reduction is insignificant even after interfacial engineering with several other competent materials. Therefore, attention must be paid to enhance the photocatalytic activity of bismuth-based materials to meet the practical requirements for photocatalytic CO2 reduction.
This journal is © The Royal Society of Chemistry 2025 |