Next Article in Journal
Apolipoprotein A-IV Has Bi-Functional Actions in Alcoholic Hepatitis by Regulating Hepatocyte Injury and Immune Cell Infiltration
Previous Article in Journal
Ring-Opening Metathesis Polymerization and Related Olefin Metathesis Reactions in Benzotrifluoride as an Environmentally Advantageous Medium
Previous Article in Special Issue
Mild and Efficient Heterogeneous Hydrogenation of Nitroarenes Facilitated by a Pyrolytically Activated Dinuclear Ni(II)-Ce(III) Diimine Complex
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

High-Throughput Strategies for the Design, Discovery, and Analysis of Bismuth-Based Photocatalysts

by
Surya V. Prabhakar Vattikuti
1,
Jie Zeng
1,
Rajavaram Ramaraghavulu
2,
Jaesool Shim
1,
Alain Mauger
3 and
Christian M. Julien
3,*
1
School of Mechanical Engineering, Yeungnam University, Gyeongsan 38541, Republic of Korea
2
Department of Physics, School of Applied Sciences, REVA University, Bangalore 560064, India
3
Institut de Minéralogie, de Physique des Matériaux et de Cosmochimie (IMPMC), Sorbonne Université, CNRS-UMR 7590, 4 Place Jussieu, 75252 Paris, France
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(1), 663; https://doi.org/10.3390/ijms24010663
Submission received: 24 November 2022 / Revised: 20 December 2022 / Accepted: 27 December 2022 / Published: 30 December 2022

Abstract

:
Bismuth-based nanostructures (BBNs) have attracted extensive research attention due to their tremendous development in the fields of photocatalysis and electro-catalysis. BBNs are considered potential photocatalysts because of their easily tuned electronic properties by changing their chemical composition, surface morphology, crystal structure, and band energies. However, their photocatalytic performance is not satisfactory yet, which limits their use in practical applications. To date, the charge carrier behavior of surface-engineered bismuth-based nanostructured photocatalysts has been under study to harness abundant solar energy for pollutant degradation and water splitting. Therefore, in this review, photocatalytic concepts and surface engineering for improving charge transport and the separation of available photocatalysts are first introduced. Afterward, the different strategies mainly implemented for the improvement of the photocatalytic activity are considered, including different synthetic approaches, the engineering of nanostructures, the influence of phase structure, and the active species produced from heterojunctions. Photocatalytic enhancement via the surface plasmon resonance effect is also examined and the photocatalytic performance of the bismuth-based photocatalytic mechanism is elucidated and discussed in detail, considering the different semiconductor junctions. Based on recent reports, current challenges and future directions for designing and developing bismuth-based nanostructured photocatalysts for enhanced photoactivity and stability are summarized.

1. Introduction

Making the most of renewable resources is a top priority. In this regard, the use of solar energy as a perennial energy source to drive chemical transformation is at the forefront of this movement. In recent years, energy demand and crisis have experienced tremendous growth due to the gradual growth of the global population and industrialization [1]. Primary energy sources come from fossil fuels, including coal, oil, and natural gas. However, one of the main problems associated with the consumption of these non-renewable energy sources is that they will run out soon. Another downside associated with this major energy consumption is the emission of carbon dioxide, which contributes significantly to global warming. Therefore, it is mandatory for humanity to switch to renewable energy sources and reduce greenhouse gas emissions. In this regard, hydrogen (H2) is a completely clean energy source when it is obtained from water-splitting processes and has been reported as an alternative energy source [2,3]. The coming decades will experience sustainable growth in its production and consumption. The photocatalytic process utilizing abundant solar energy as a light source is considered the key to hydrogen production and carbon dioxide conversion [4]. In addition, photocatalysis can also be used to deal with some other environmental problems, including the degradation of organic pollutants, such as rhodamine B (RhB), indigo carmine (IC), methylene blue (MB), methyl orange (MO), brilliant green (BG), bisphenol A (BPA), tetracycline (TC), chlortetracycline (CTC), acetaminophen (APAP), 4-tert-butylphenol (BTBP), ciprofloxacin (CIP), 2,4-dichlorophenol (2,4-DCP), etc., through some complex reactions under mild conditions, such as photocatalytic organic synthesis [5]. Photocatalysis has emerged as a benchmark tool that combines light and catalysts to perform chemical transformations that are elusive using standard synthetic procedures.
In the past decade, rapid development in the synthesis of photocatalyst nanomaterials has taken place [6]. In addition to this, a variety of photocatalysts have been used, which can convert sunlight into chemical energy and transfer this energy to reactive molecules for efficient photo-fuel conversion. Many catalysts have been studied for degrading pollutants or water splitting. Wide-gap semiconductors such as TiO2 can absorb only <5% of solar light. On the other hand, small-gap semiconductors can absorb visible light, but their photocatalytic performance is also poor because of the rapid electron-hole recombination rate. Bismuth compounds have emerged as a family of promising photocatalysts. According to their composition, their band gap can be adjusted to the desired value for visible light absorption. In addition, their internal electrical field induces the separation of the photogenerated charge carriers so that the recombination rate is reduced. Among the different photocatalyst groups, photocatalysts based on nanostructures, nanocomposites, or heterostructures have shown superior photodegradation efficiency compared to their bulk counterparts [7,8]. However, the relevant photocatalytic processes involved in nanostructured photocatalysts are more complex and far from fully understood. For example, identifying the real catalytically active species remains unsolved in most cases and is needed to obtain reproducible and efficient nanostructured or heterostructured photocatalyst materials for practical applications. In fact, the importance of discovering the truly catalytically active species involved in photocatalytic systems allows for a better and more general understanding of the photocatalytic process, which can help improve its efficiency.
In this review article, more attention is paid to the fundamental concepts, mechanisms of the synthetic process, and structural features dependent on the photocatalytic activity of bismuth-based photocatalysts. Their intrinsic photocatalytic properties, namely the charge transfer and separation, excitation formation, and catalytic activity by the formation of nanostructures are described. Their recent trends, including photodegradation, H2O decomposition, CO2 conversion, and important approaches to enhance photocatalytic activity are highlighted. Photocatalytic enhancement via the surface plasmon resonance effect is also examined and the photocatalytic performance of the bismuth-based photocatalytic mechanism is elucidated and discussed in detail, considering the different semiconductor junctions. Finally, the current challenges and future development of bismuth-based photocatalysts are described.

2. Background

Bismuth(III) oxide (Bi2O3) exists in six crystallographic polymorphs; namely, monoclinic α, tetragonal β, body-centered cubic γ, face-centered δ, orthorhombic ε, and triclinic ω [9]. Among them, the α, β, and δ phases show photocatalytic reactivity upon visible light irradiation [10]. Bi2O3 has been used as a heterogeneous photocatalyst capable of catalyzing the degradation of several synthetically important sunlight-driven pollutants. It possesses a narrow bandgap (2.1–2.8 eV) with useful photocatalytic activity. Due to the high-oxidation potential of valence band holes (+3.13 V vs. normal hydrogen electrode (NHE)), its photo-efficiency has been demonstrated in a variety of applications ranging from energy storage and pollutant degradation to bio-compound degradation [11,12,13]. Although Bi2O3 exhibits high efficiency in promoting photooxidation, its conduction band (CB) electrons (+0.33 V vs. NHE) are unable to interact with organic molecules because of the rapid recombination of charge carriers, which hinders its application in reduction processes. However, several studies have shown that the Bi2O3 photocatalytic activity can be improved either by doping or by combining two or more materials or tuning surfaces [14,15,16].
Recently, Bi2O3 has also become a popular photocatalyst for driving the photodegradation of organic dyes. Its photocatalytic activity was investigated for the formation of C-C and C-S bonds [17,18] and atom-transfer radical-addition-type reactions [19]. Bi2O3 is low-cost and non-toxic. Other advantages are its visible-light drive and high availability. Moreover, in some cases, it can replace the use of organo-metal complexes such as the photoredox catalysts Ru(bpy)3Cl2-combining 2,2′-bipyridine and expensive and not-abundant ruthenium [20]. The semi-metallic nature of Bi0 enables its use as a semiconductor photocatalyst, or as a cocatalyst to tune the photocatalytic behavior of the host material [21,22]. Bismuth is a metal element in group V B in the periodic table (Mw = 208.98 g mol−1) with the 6s26p3 electron configuration [23]. The lone-pair distortion of Bi 6s orbitals in Bi-based composite oxides may lead to the overlap of O 2p and Bi 6s orbitals in the valence band, which is beneficial to reduce the bandgap and the mobility of photogenerated charges and improve the photoactivity [24]. At the same time, when the 6s orbital is empty, the Bi5+ valence state also has good absorption of visible light [25]. Bismuth exists as Bi3+ in most common Bi-type photocatalysts, such as complex oxides (BiVO4, Bi2WO6, BiPO4) [26,27,28], sulfides (Bi2S3) [29], and oxyhalides (BiOI, BiOBr, BiOCl) [30,31,32,33]. It is therefore not surprising that a 14-fold increase in the number of reports related to Bi-based photocatalysts has been observed from 2010 to 2022 [34,35].
In order to improve the mineralization rate of organic dyes in wastewater, high charge separation efficiency, long-term stability, suitable band edge positions, and good redox capacity are also required for high-efficiency photocatalysts in addition to suitable bandgaps. Therefore, due to the unsatisfactory photocatalytic activity of single-component photocatalysts, various controllable bismuth-based compounds have been synthesized through morphological structure mediation, the construction of heterostructures or nanostructures, the doping of metal elements, and defect site mediation [36,37,38,39]. In particular, S,F-codoped Bi2WO6 with oxygen vacancies synthesized via hydrothermal calcination and post-sulfurization showed a photocatalytic performance in Cr(VI) reduction of 94.3% and methyl orange degradation of 95.4% in 120 min under visible light [40]. Among defects, oxygen vacancies (OVs) have been shown to improve photocatalytic activity [41,42]. However, the surface oxygen defects are unstable due to easy oxidation during the photocatalytic reaction process [43]. An exception is provided by Bi5O7Br nanotubes synthesized with OVs by combing water-assisted self-assembly and a low-temperature wet chemical approach [44]. In this case, the surface OVs were stable and able to capture and activate N2, reducing it to NH3, in pure water. The properties of many Bi-based photocatalysts and their performance in organic dye degradation and H2 production are described in detail in the following sections.

2.1. Fundamental Mechanism and Main Active Species of Bismuth-Based Photocatalysts

The main mechanism of bismuth-based photocatalysts can be summarized as photon absorption, excitation, and reaction processes. The photocatalytic process can be applied not only to the degradation of dyes but also to the degradation of antibiotics. Specifically, the photocatalytic degradation of antibiotics based on bismuth-based semiconductors is an effective, eco-friendly, and promising method for toxic substances. The predominant mechanisms of the bismuth-based photocatalysts for antibiotic photocatalytic degradation could be summarized as absorbing photons, excitation, and reaction, as shown in Figure 1. The antibiotics, as well as their intermediates, are converted to small-molecule compounds via the oxidation of oxygen species (h+, •O2 or •OH) and eventually decomposed into CO2 and H2O [45].
In detail, when a photocatalyst absorbs photons with energy higher than its bandgap, the valence band (VB) electrons can be excited and jump into the CB. A photohole is inseparable from a photocatalyst and is a vacancy in its crystal lattice. Thus, the photocatalytic process is expressed as follows:
photocatalyst + hv → (photocatalyst + h+) + e,
then, the photogenerated electrons and holes are effectively separated and migrated to the photocatalyst surface. The photo-induced holes directly attack dye molecules, as follows:
h+ + dye → H2O + CO2 + degradation product,
theoretically leading to the significant degradation of these pollutants. Furthermore, when the holes further migrate to the photocatalyst surface, the oxidation pathway starts, accompanied by the oxidation of H2O/OH to generate hydroxyl radicals (•OH), such as:
H2O/OH + h+ → •OH + H+.
Meanwhile, the typical redox potential of the photocatalyst should be higher than •OH/OH (+1.99 eV). Furthermore, hydroxyl radicals have stronger oxidation potential (E0 = 2.8 eV) and lower selectivity than other oxidants during the decomposition of water pollutants [46]. Remarkably, the top of the VB of most Bi-based catalysts is higher than the redox potential of •OH/OH, indicating that hydroxyl radicals are easily generated during Bi-based catalysis. In fact, the reaction pathway between hydroxyl radicals and dye molecules can be summarized as follows: (i) •OH and dye molecules simultaneously adsorb on the catalyst surface and then react spontaneously, (ii) •OH in aqueous solution and adsorbed on the photocatalyst surface reacts with dye molecules, (iii) •OH adsorbed on the catalyst surface reacts with the surrounding dye molecules, and (iv) finally, •OH reacts with the dye molecules in the aqueous solution. Generally, these main pathways are considered for the degradation of dye molecules by bismuth-based photocatalysts [47]. Figure 2 displays the band edge positions of bismuth-based photocatalysts.
Conversely, if the CB potential of the semiconductor is negative compared to the O2/•O2 redox potential (−0.13 V vs. NHE), a reduction pathway can also be observed, in which O2 is reduced by electrons to •O2 (O2 + e → •O2). The excited H2 ions will recombine with electrons and generate thermal energy (H+ + e → energy), which reduces the photodegradation efficiency of the catalyst. Then, dye molecules and their intermediates are converted into small molecular compounds through the oxidation of O2 species (h+, •O2 or •OH) and finally decomposed into CO2 and H2O (dye molecules + radicals (•OH or •O2) → CO2 + H2O + small molecule compound). In contrast, due to their different electronic structures, the effect of various active substances on the degradation of dye molecules differs. In order to direct the preferential active species during the reaction, different scavengers such as MeOH (for •OH), KI (for h+), p-benzoquinone (for •O2 radicals), and AgNO3 (for e) were introduced into the reactor to trap the active species [48]. Apparently, besides superoxide radical (•O2), hydrogen peroxide (H2O2) and hole (h+) play a major role in the photodegradation process of most organic, dye-based, bismuth-based photocatalysts [49]. For example, the main reason for their higher photocatalytic degradation may be that the dye molecules are considered vulnerable to h+ attack [50]. In addition, hydroxyl radicals (•OH) typically react rapidly and non-selectively with most organic pollutants, so they also play a key role in the degradation of dye molecules. It is worth noting that the resulting •O2 is unstable and prone to disproportionation reaction to generate other reactive oxygen species including •OH.

2.2. Significance of Nanostructure or Heterostructure or Heterojunction or Nanointerfaces

In fact, a practical approach to improve the photo-response in photoactive materials is the formation of nanointerfaces (heterojunctions) by coupling with other semiconducting materials or metals. The formation of heterojunctions has been widely used in visible-light-responsive photocatalytic dye degradation and H2O splitting in batteries [51,52,53]. It is also worth noting that heterojunctions in nanomaterials can be mainly classified into three different types: (1) type I is a straddling gap, where the VB and CB energies of the cocatalyst are higher and lower than those of the photocatalyst; (2) type II is a staggered gap, the VB and CB of the cocatalyst are higher than that of the catalyst; (3) Z-scheme has the same band structure as Type-II, but with different charges of the acceptor/donor pair carrier transfer pathway, which may enhance redox capacity. The Z-scheme configuration can further be differentiated into three types; namely, direct Z-scheme (mediator free), solid-mediator, and redox pair mediator types (see Figure 3 [54]).
Indeed, the band gap and Fermi level can be tuned at their interfaces, providing charge separation and facilitating alternative paths for excited electrons to prevent charge recombination. For example, Shan and co-workers [55] proposed a band alignment of α-Bi2O3/BiOCl (001) core-shell heterojunctions based on the shifted positions of the CB and VB to facilitate the accumulation of photoinduced electrons at their interfaces. Volnistem et al. [56] constructed mechanistically synthesized BiFeO3/Fe3O4 nanostructures and used them to degrade MB dyes under visible light. This study shows that the nanointerface promotes the ferrous Fe2+ ions of Fe3O4 to enhance the catalytic efficiency compared to the bulk. The direct Fenton-like method is another effective method to degrade dyes using Fe2+ ions and H2O2. The presence of Fe2+ ions combined with the photo-Fenton process can enhance the decomposition of H2O2 into oxidative radicals, thereby increasing the degradation rate. In another study, Liu et al. [57] reported that the hydrothermally synthesized Bi4Ti3O12/BiOI nanostructures degrade BPA. The results showed that the degradation was 12 times faster than that of Bi4Ti3O12 crystallites, which was attributed to the internal electric field in the ferroelectric domains under the external electric field. The improved internal electric field in ferroelectric catalysts can facilitate the separation and transfer of charge carriers, driving more carriers to the surface of the photocatalytic material, thereby enhancing its photocatalytic efficiency.
BiOX/CuFe2O4 (X = Br, Cl, and I) nanostructured p–n junctions were constructed by hydrothermal and coprecipitation methods [58]. Such a nanostructure induces a built-in electric field at its interface, which facilitates the transfer of pattern changes, indicating a significantly enhanced visible-light-driven photoactivity without the use of any cocatalyst. The BiOI or BiOBr/CuFe2O4 nanostructure demonstrates the conventional type I and type II charge-transfer mechanism, which can effectively reduce the charge-transfer resistance compared with the bulk structure. Importantly, the direct Z-type mechanism of BiOCl/CuFe2O4 nanostructures has formed tight interfacial contacts, resulting in a 5.7-fold increase in H2 release compared to bare BiOCl and improved catalytic efficiency by a factor of two compared to type II BiOI/CuFe2O4 nanostructures. The study also shows that the low resistance of the Nyquist plot confirms the superiority of the direct Z-scheme in promoting charge separation and transfer and increasing carrier density. Furthermore, by designing the band-edge potential, the BiOX/CuFe2O4 heterostructure achieves optimal space charge layer width and redox potential, which reduces the fast recombination rate. This work delivers a model for designing highly engineered BiOX-based nanostructures with tuned band edges for efficient photocatalytic activity. Therefore, the construction of nanostructures or heterostructures is of great significance for improving photocatalytic efficiency.

2.3. Surface Plasmon Resonance Effects in Bismuth-Based Photocatalysts

The surface plasmon resonance (SPR) effect of noble metals such as gold and silver is currently used to enhance the visible photocatalytic activity of semiconductor photocatalysts [59]. This mechanism is attributed to the huge local electric field enhancement observed at the surface of metallic nanoparticles (NPs) due to the interaction with the electric and magnetic fields of light. The excitation of electron-hole (e-h+) pairs is boosted in the catalyst with the enhanced near-field of NPs, improving the photocatalytic activity. Although, the SPR effect of Bi nanospheres has been used to stimulate the excitation of photo-generated e-h+ pairs in Bi-based semiconductor photocatalysts by the deposition of Bi on their surface. Improvement via the SPR effect of visible photoreactivity has been demonstrated for BiOBr [60], Bi4MoO9 [61], BiPO4 [62], and Bi2WO6 [63].

3. Synthesis Strategies of Bismuth-Based Photocatalysts

Various strategies have been proposed for the preparation of bismuth-based photocatalysts with desirable structure and morphology [64]. Synthesis methods of bismuth-based photocatalysts include room-temperature solid-state milling [65], high-temperature solid-state reaction [66], the precipitation method [67], the hydro/solvothermal technique [68,69], the ion-exchange route [70], the microwave-assisted method [71], and the microemulsion-based route [72]. Lu et al. [73] reported the synthesis of β-Bi2O3 microrods of ~1 µm in width by a solution crystallization technique at 70 °C, without further calcination treatment. The metastable tetragonal β-Bi2O3 crystalline powders with orange color transformed into yellow monoclinic α-Bi2O3 crystals after 60 min of reaction. The morphology of a pumpkin was changed from the microrod-like structure of β-Bi2O3 crystals to the large rhomboid structure of α-Bi2O3 crystals. This synthesis method reflects the size control process of metastable β-Bi2O3. By simply adjusting the experimental parameters such as NaOH concentration, stirring, and reaction temperature, metastable β-Bi2O3 crystals can be stably stored in the reaction system for different lengths of time. The pumpkin β-Bi2O3 nanostructures exhibited good photocatalytic performance for the degradation of RhB dyes under visible light irradiation.

3.1. Synthesis Strategies of Bismuth Oxides

One-dimensional (1D) Bi2O3, including nanotubes [74], nanowires [75], nanosheets [76], and nanorods [77], holds promise for photocatalytic activity. Tien et al. [78] reported the synthesis of α-Bi2O3 nanowires with a diameter of 500 nm and a length of up to 20 μm by catalyst-driven gas-phase transport, and the growth direction was (010). The growth mechanism of α-Bi2O3 nanowires is explained as a two-step growth model, which considers the formation of crystal planes catalyzed by gold and the growth of α-Bi2O3 nanowires during bismuth catalysis. It is revealed that the formation and growth mechanism of α-Bi2O3 nanowires is influenced by Au nanoparticles. In detail, the formation of nanowires is shown in Figure 4. In steps 1 and 2, the Au-catalyzed growth of the precursor vapor is adsorbed onto Au nanoparticles, where facets are formed between the Au/Bi interface under an oxidizing environment. Once a facet is formed, it can serve as a nucleation site for further one-dimensional growth. At this stage, growth has shifted into different growth patterns (steps 3 and 4). Since the growth occurs at the dual-surface interface, the growth is driven by a dual catalytic mechanism. By comparing nucleation on heterogeneous solid surfaces (gold nanoparticles) and self-nucleation (sapphire surfaces), nucleation on gold catalysts interacting with nuclei will have lower free energy than self-nucleation. Therefore, the degree of supersaturation required for self-nucleation is much higher than for heterogeneous nucleation. Finally, the formation of crystal planes provides the nucleation and growth of α-Bi2O3 nanowires, which facilitates a simple strategy to control the nucleation and structural characteristics of α-Bi2O3 nanowires.
Xiao et al. [79] reported the solvothermal synthesis of β-Bi2O3 nanospheres followed by a calcination process. In detail, monodispersed bismuth nanospheres were formed by a solvothermal process with D-fructose as the main reducing agent, followed by calcination in air to transform into β-Bi2O3 nanostructures, revealing that the D-fructose concentration significantly affects the structural β-characteristics of Bi2O3 nanospheres. The growth mechanism of β-Bi2O3 nanospheres involves the in situ reduction of Bi(III)-ethylene glycol complex spheres as self-sacrificial agents, followed by the in situ oxidation of bismuth nanospheres by contact with oxygen. The β-Bi2O3 nanospheres exhibited APAP degradation efficiency that was 79 times higher than that of TiO2 powder (Degussa P25), which was attributed to the suitable energy band structure, high oxidation potential, and good dispersion of β-Bi2O3 nanospheres. Higher photoactivity is supported by the experimental determination of reactive oxygen species during photocatalysis. However, secondary pollutants may still occur during the photocatalytic process, requiring an in-depth analysis of the treated water.

3.2. Synthesis Strategies of Bismuth Ferrites

BiFeO3, with a rhombohedral twisted perovskite structure and a narrow bandgap visible light response of 2.2 eV, is an attractive candidate for its fascinating application in novel photocatalyst materials [80]. BiFeO3 is a ferroelectric material, and the intrinsic internal electric field reduces the recombination of photoinduced charge carriers and increases the degradation rate. The low decomposition temperature of bismuth salts and the change in ionic valence make it difficult to prepare impurity-free BiFeO3 by conventional solid-state reactions at the evaluated temperature. Other strategies such as the soft sol-gel method, co-precipitation method, and solvent/hydrothermal method have been identified as promising strategies to prepare BiFeO3 nanoparticles with desirable morphology [81,82,83]. The hydrothermal synthesis of BiFeO3 without high-temperature calcination can better control the purity and morphology of the material by controlling the reaction conditions. Due to the influence of the number of reaction sites and the size of the bandgap energy, the shape and size control of the particles plays a very important role in the photocatalytic activity. Many different morphologies of BiFeO3 particles have been reported, including wires, tubes, submicron spindles, and rod-like particles [84,85], including synthetic steps and controlled processes of self-assembly or organization of BiFeO3 with regular geometry. However, the continuous regulation of the shape or size of microscale BiFeO3 and its effect on photocatalytic activity is an important topic of core research, which is of great significance for understanding the catalytic mechanism and developing dual-semiconductor photocatalysts.
Ferrite bismuth materials, including perovskite (BiFeO3), mullite (Bi2Fe4O9), and sillenite (Bi25FeO40), exhibit outstanding magnetic, electronic, and dielectric properties. Among them, mullite-based Bi2Fe4O9 is a competitive candidate photocatalyst to drive visible light-catalyzed oxidation reactions due to the band gap energy of 1.9–2.1 eV, with standard multi-band semiconductor properties [86]. However, the catalytic efficiency of Bi2Fe4O9 is relatively low due to the fast recombination of photogenerated electron-hole pairs [87]. The photosensitivity can be improved by the separation of electrons and holes in Bi2Fe4O9 with silver halides (AgX, X = Br, I, and Cl). Unfortunately, the strong photosensitivity of silver halides leads to the reduction of Ag+ to Ag0 under light irradiation, which reduces their stability and lifetime, thus limiting their photocatalytic applications. In general, AgBr is a popular high-efficiency photocatalyst with a band gap of 2.6 eV. Ma et al. [88] developed a one-dimensional magnetically separable Bi2Fe4O9/C@AgBr nanostructured photocatalyst that can degrade 97.4% of MB within 60 min. The high catalytic performance is mainly due to the efficient charge separation and migration in the Bi2Fe4O9/C@AgBr nanostructures. In addition, carbon also promotes the chemical protection of nanostructures and improves the conductivity and stability of catalysts.

4. Recent Developments in Bi-Based Photocatalysts

4.1. Bi-Oxide Nanostructures

Bandgaps of Bi2O3 polymorphs range in the order of δ-Bi2O3 (3.0 eV) > α-Bi2O3 (2.8 eV) > β-Bi2O3 (2.1 eV) > γ-Bi2O3 (1.64 eV) [89]. Since γ-Bi2O3 has a narrow band gap, it can efficiently utilize light in the visible region of the solar spectrum. However, the photocatalytic efficiency of bare Bi2O3 is still unsatisfactory for dye degradation unless it is integrated or doped with other semiconducting compounds, especially for the δ-phase because of its wide bandgap value. In particular, the α-Bi2O3 type is a stable phase over a wide temperature range, while β-, γ-, and δ-Bi2O3 are metastable at 25 °C. For this reason, many strategies have been applied to enhance the photoactivity of Bi2O3. Barno et al. [90] pioneered the hydrothermally prepared heterojunction features of BiVO4/MnV2O6 photocatalysts for the photodegradation of RhB and MB dyes, showing that the BiVO4/MnV2O6 heterojunction photocatalyst achieved a degradation rate of 96% in 35 min and MB dye degradation efficiency reached 98% within 6 min under visible light exposure. This study shows that superoxide anion radical is the main responding species during dye degradation. Furthermore, the BiVO4/MnV2O6 heterojunction photocatalyst exhibits excellent 4-nitrophenol reduction in the presence of NaBH4, and 4-aminophenol is produced without intermediate by-products, thanks to the heterojunction properties and its suitable band alignment. By controlled hydrothermal synthesis, allowing the control of the ratio of {010} to {110} facets on BiVO4, which respectively serve as reductive and oxidative sites, Guan et al. obtained decahedral BiVO4 single crystals with superior photocatalytic water oxidation achieving efficient water splitting [91]. Tian et al. [92] reported thermochemically prepared β-Bi2O3/Mn3O4 nanostructures for the photodegradation of RhB, BPA, and MB and the removal of nitric oxide (NO). The optimized β-Bi2O3/Mn3O4-2 wt.% photocatalyst exhibits excellent photocatalytic activity for pollutant (RhB, MB, and BPA) degradation and NO removal. This efficiency was ascribed to tight contacts between the β-Bi2O3 and Mn3O4 at their interface, which possesses a type-II heterojunction photocatalytic mechanism. This mechanism facilitates the rapid separation of photo-induced charge carriers, resulting in excellent photocatalytic activity.
He et al. [93] pioneered solvothermally synthesized 3D flower-like β-Bi2O3/Bi12O17Cl2 nanostructures for the degradation of PTBP under visible light. The nanostructures are formed by reducing Bi(III) to nano-metallic bismuth, followed by the thermal treatment of bismuth with oxygen and bismuth oxide chloride hydroxide in the presence of air. The synthesized β-Bi2O3/Bi12O17Cl2 nanostructures have a good energy band structure, and a close-contact heterojunction is formed between the synthesized β-Bi2O3 and Bi12O17Cl2, with a high specific surface area and a hierarchical micro-nanostructure, thereby decomposing PTBP under visible light with excellent photo-mineralization efficiency. Compared with the as-synthesized Bi12O17Cl2, the optimally synthesized β-Bi2O3/Bi12O17Cl2 nanostructure exhibited 12-fold higher photocatalytic activity, which was attributed to the direct hole and superoxide radical oxidation rather than oxidation by hydroxyl radicals. Due to the presence of heterojunction features, the visible light absorption range is enhanced at the origin of their remarkable photoactivity under visible light illumination. Sun et al. [94] reported water thermal synthesis of α-/γ-Bi2O3 nanostructures for RhB degradation under visible light. The key parameters of the hydrothermal process are holding time, temperature, additive dosage, and pH conditions. Compared with α-Bi2O3 and γ-Bi2O3 nanostructures, α-/γ-Bi2O3 nanostructures exhibited higher RhB photodegradation activity, which was attributed to the synergistic effect of the homojunction. Gardy et al. [95] reported a solid reactive heat treatment made of α-/β-Bi2O3 nanopowders to degrade a mixed dye of RhB and IC under UV and visible light irradiation. It can be observed that the α-/β-Bi2O3 mixed phase produced 20% β-Bi2O3 phase after annealing at 550 °C, while the α-Bi2O3 heterojunction was formed after annealing at 650 °C. α-/β-Bi2O3 photocatalysts exhibit better efficient charge separation and activity via α-/β-Bi2O3 transfer, indicating that α-/β-Bi2O3 heterojunctions are more efficient than commercial α- and β-Bi2O3 materials separately.

4.2. Bismuth Vanadate (BiVO4)

BiVO4 has attracted extensive attention due to its remarkable structural, optical, and chemical properties, photocorrosion resistance, and good activity for the photocatalytic degradation of organic pollutants. The crystal structures of BiVO4 are monoclinic, orthorhombic, and tetragonal, among which the monoclinic with a band gap of about 2.4 eV has good photocatalytic activity compared with the other two forms. The phase transition from tetragonal to monoclinic occurs irreversibly at 500 °C. The basic building blocks are developed from the VO4 tetrahedron and the BiO8 dodecahedron. In addition, Bi and V atoms are alternately arranged along the crystallographic axis, which makes monoclinic BiVO4 exhibit the properties of a layered structure. However, BiVO4 has limited use as a photocatalyst because of its fast recombination rate of photogenerated carriers due to its band edge position. Furthermore, the photocatalytic efficiency of BiVO4 is much lower than expected due to its lower surface area and lower carrier separation and transfer ability. Therefore, topography control, cocatalysts and selective deposition, and coupling to other semiconductors to build nanostructures are needed. For example, Liu et al. [96] developed an in situ transformation of as-prepared BiVO4 with the help of NaOH to form BiVO4/Bi25VO40 nanostructures through a dissolution–recrystallization process, in which the monoclinic decahedron of BiVO4 was first etched with an alkaline solution on the preferential crystal planes (010) and converted to cubic Bi25VO40. In this study, the authors successfully controlled the concentration conditions of the alkaline solution to precisely tune the phase composition of the heterojunction by reducing BiVO4 and increasing Bi25VO40 in the nanostructures. Advancing from the in situ switching strategy and combined band structure, the type II tight heterojunction formed tight interfacial contacts. This led to fast charge transfer with the spatial separation of carriers, which considerably enhanced the photocatalytic degradation of tetracycline hydrochloride (TCHC) under visible light. In the obtained BiVO4/Bi25VO40 nanostructures, the role of Bi25VO40 is crucial in the photoactivity; the atypical bismuth-rich phase bismuth vanadate consists of the same elements as BiVO4 with a narrower bandgap of 2.1 eV, resulting in a wider range of visible light absorption. Since BiVO4 and Bi25VO40 have suitable energy band positions and approximate crystal structures, the rational coupling of BiVO4 and Bi25VO40 through an in situ synthesis process has been shown to yield close-contact heterojunctions with suitable band energies, leading to an enhanced charge-carrier transmission rate. In another study, Duan et al. [97] reported hydrothermally synthesized BiVO4/rGO nanostructures with the assistance of ethylenediaminetetraacetic acid disodium salt instead of nitrate, which facilitates the formation of a fully acidic environment to prevent the hydrolysis of Bi3+, and applied them as a photocatalyst for the degradation of RhB. The BiVO4/rGO nanostructured photocatalyst exhibited 98.3% degradation in 180 min under visible light. When photons land on the BiVO4/rGO surface, the electrons in the VB of BiVO4 are excited to the CB by leaving a hole on the VB. Since rGO acts as an electron acceptor with good electrical conductivity, the photogenerated electrons in the CB can move to rGO, which speeds up the separation efficiency and thus enhances photoactivity.
El-Hakam et al. [98] reported the ultrasound-assisted introduction of mesoporous SiO2 (i.e., m-SiO2) on BiVO4 nanoparticles to control the size of BiVO4 nanoparticles to 2.4–5.1 nm on m-SiO2 to form BiVO4/m-SiO2 nanostructures. This BiVO4/m-SiO2 nanostructure was used to degrade MB and BG dyes as a function of m-SiO2 in the nanostructure. Compared with bare BiVO4, the BiVO4/m-SiO2 nanostructured photocatalyst exhibits remarkable photoactivity, which is attributed to the synergistic effect between m-SiO2 and BiVO4, which enhances the separation of charge carriers. The effects of operating parameters such as dye concentration, m-SiO2 content, reaction time, and temperature were closely related to the photocatalytic activity. The nanostructure with a 10 wt.% m-SiO2/BiVO4 sample exhibited the highest photocatalytic activity. Since the silica in the nanostructure is in close contact with the BiVO4 nanoparticles, the photoelectron conversion of BiVO4 is improved by reducing the recombination charge carriers based on the suitable band positions of BiVO4, which is found to be reusable.

4.3. Silver-Bismuth Photocatalysts

Advanced oxidation methods are considered promising methods to solve environmental problems by releasing free radicals, which have strong oxidative power. Furthermore, chemical oxidation and photocatalysis are two common tools for removing pollutants from wastewater. Compared with semiconducting oxides, perovskite-group silver-bismuth-based photocatalysts have different crystal structures, which provide a wide range of degrees for tuning their physicochemical properties. Recent studies have shown that silver bismuth, through defect engineering, can effectively enhance its photoactivity due to its physical adsorption and chemical oxidation [99]. Therefore, it significantly improves the mineralization ability of organic dyes. Typically, silver bismuth is synthesized from AgNO3, which replaces the Na ions of NaBiO3 in a hydrothermal reaction. The silver-bismuth photocatalyst has chemical oxidation abilities mainly due to the release of lattice oxygen in bismuthate, which is partially converted into active oxygen. When these reactive oxygen species come into contact with dye molecules, large amounts of reactive oxygen species are released. However, due to the irreversible transformation of lattice oxygen into chemisorbed oxygen, accompanied by the transformation from Bi(V) to Bi(III), the photocatalytic performance of silver bismuth single compounds decreases to varying degrees [100]. Many studies have been carried out to promote the release of reactive oxygen species, thereby enhancing the catalytic ability of silver bismuth. Silver bismuth was converted into α-/β-Bi2O3/Ag2O nanostructures by a simple calcination process, during which the morphology was transformed from nanosheets to porous nanosheets and ravines, for the degradation of TC under visible light [99]. After the deactivation of silver bismuth, the lattice oxygen is transformed into chemically and physically adsorbed oxygen, and numerous carbon species can be adsorbed onto the surface of the material. Bi species in silver bismuth are transformed to β-Bi2O3, and all Ag species are converted to AgO2. Furthermore, with increasing temperature, the β-Bi2O3 phase transforms into α-Bi2O3 identified by the color of the sample, and morphological changes may occur as described above. β-Bi2O3/Ag2O activated at 290 °C exhibited the best degradation efficiency of 78% within 2 h, and its reaction rate constant was 3.7 times higher than that of silver bismuth due to the low recombination probability and strong photoresponsivity. Figure 5 shows the possible photocatalytic mechanism of silver-bismuth-based photocatalysts.

4.4. Bismuth Oxide Silicate-Based Photocatalysts

The compounds with molecular formula Bi2XO20 (X = Si, Ti, Ge, Pb, etc.) are called bismuth sillenites and are considered promising materials for developing low-temperature co-fired ceramic technology [101]. Mainly bismuth silicates, such as Bi2SiO5, Bi12SiO20, and Bi4Si3O12, have received increasing attention, in particular Bi2SiO5, as an alternative to conventional lead-based ferroelectric materials with a phase transition temperature of 673 K. Bi2SiO5 crystallizes with an orthorhombic structure (space group Cmc21) with lattice constants a = 15.19 Å, b = 5.68 Å, c = 5.314 Å and Z = 4 [102]. As the general formula of the Aurivillius-like structure is (Bi2O2)[Am−1(B)mO3m+1], Bi2SiO5 with m = 1 is composed of BiO4 pyramids in the [Bi2O2]2+ layers and [SiO3]2− layers, as shown in Figure 6a [103]. First-principle calculations suggest that the polarization of Bi2SiO5 originates from the SiO3 layer rather than the Bi2O2 layer [104].
Figure 6b–e confirm the concept of fragments for the experimental extraction of single dipole units derived from BiO and SiO3 clusters. The boundaries of a segment can be determined by local minima around the segment. Therefore, the fragments satisfy the charge neutrality of Bi2SiO5, and their partial electrical polarization is estimated by considering the volume of the unit cells. Bi2SiO5 crystals have special properties, namely dielectric properties, thermoelectric properties, and nonlinear optical properties, and have ferroelectric properties due to their non-centrosymmetric structure, and the band gap of 3.54 eV is narrower than that of BiPO4 (3.85 eV) [105]. Actually, the inductive effect of PO43− benefits the separation of the photoinduced electron-hole pairs, but the large energy gap of BiPO4 implies that this compound is a good photocatalyst only in UV light, which accounts for 4% of solar irradiation. Therefore, the formation of nanostructures with a smaller gap material such as Bi2SiO5 is a means of avoiding this drawback.
Co-precipitated hydrothermally synthesized Bi2SiO5/BiPO4 nanostructures were constructed and applied to degrade phenol and MB dyes under UV-light irradiation [106]. This work revealed the extension of the photoresponse range of BiPO4 by coupling with Bi2SiO5 and forming a type-II heterojunction. Bi2SiO5/BiPO4 nanostructures exhibit significantly enhanced photoactivity against phenols and dyes, being 4.36-fold and 1.13-fold higher with respect to Bi2SiO5. This improvement was attributed to the significantly enhanced charge separation ability, expanded absorbance, and good crystallinity through the heterojunction. A synergistic effect was observed with medium Brunauer–Emmett–Teller specific surface area. The energy (vs. NHE) of the bottom of the CB of BiPO4 (−0.65 eV) is more negative than that of Bi2SiO5 (0.05 eV), while the top of the VB of BiPO4 is at 3.2 eV against 3.59 eV for Bi2SiO5. Therefore, the two components, BiPO4 and BiSiO5, have matched charge potentials, which can facilitate the flow of charge carriers through their interfaces. Furthermore, it is reasonable to design a type-II heterojunction as a novel and robust photocatalytically active system by coupling BiPO4 with a narrower bandgap semiconductor, Bi2SiO5. Zou et al. [107] reported a one-pot solvothermal synthesis of Bi2SiO5/Bi4MoO9 nanostructures for the degradation of CIP under UV-light irradiation. The results show that the Bi2SiO5/Bi4MoO9 nanostructures exhibit higher photoactivity than Bi2SiO5 and Bi4MoO9; such a heterostructure not only suppresses the recombination of photoexcited charge carriers but also enhances light absorption. In addition, the effects of initial CIP concentration and coexisting ions on the photodegradation process of Bi2SiO5/Bi4MoO9 nanostructures were also confirmed. Figure 7 pictures the density-functional theory results of Bi2SiO5 and Bi4MoO9, indicating that the VB and CB of Bi2SiO5 and Bi4MoO9 are the same k-space, indicating that the intrinsic optical transition properties of Bi2SiO5 and Bi4MoO9 are direct transitions. The estimated theoretical band gaps of Bi2SiO5 and Bi4MoO9 are 3.69 and 2.86 eV, respectively, which are in good agreement with the experimental results. The total and partial electronic state densities of Bi2SiO5 and Bi4MoO9 are shown in Figure 7c,d. The VB top of Bi2SiO5 is mainly composed of O 2p and Si 3p orbitals, while the CB bottom is mainly composed of Bi 6p orbitals (Figure 7c). In the case of Bi4MoO9, the VB top is mainly composed of O 2p states, while the CB bottom is contributed by Mo 4d states.
In general, the sol-gel route has potential advantages over traditional solid-state synthesis methods because it allows precise control over composition, coating deposition, and uniformity. Veber et al. [108] proposed a synthetic procedure for bismuth silicate, consisting of bismuth nitrate pentahydrate, dried in a vacuum oven at 65 °C for 96 h to remove water contamination. The dehydrated bismuth nitrate was then dissolved in acetic acid and placed in a magnetic stirrer for 2 h. Another solution was prepared with tetraethoxysilane (Si(OC2H5)4, TEOS) and 2-ethoxyethanol with continuous stirring for 30 min. 2-ethoxyethanol can be used as a solvent for TEOS. After the two solutions were stirred separately, they were mixed and placed under magnetic stirring for 3 h and adjusted to pH 4 with ethanolamine. Di et al. [109] reported Bi2SiO5 nanosheets modified by carbon quantum dots (CQDs) with a diameter of 3 nm and applied them to the photocatalytic degradation of RhB under UV-light irradiation. CQDs-modified Bi2SiO5 nanosheets were shown to accelerate the charge transfer between the interiors of Bi2SiO5 nanosheets and promote the separation of surface charge carriers. The active species that enhance the photocatalytic activity are hydroxyl radicals and superoxide radicals, as evidenced by electron spin resonance analysis. Under UV-light irradiation, electrons are transferred from the VB to the CB of Bi2SiO5. Electrons migrating from the surface of Bi2SiO5 are transferred to the CQDs through the interface between the CQDs and Bi2SiO5. Electrons on the CQDs reduce the adsorbed O2 to •O2, while holes on the VB oxidize OH to OH. The generated reactive oxygen species play a key role in the subsequent photocatalytic degradation process. Yang et al. [110] developed nanostructures of Bi2SiO5/g-C3N4 using a controllable hydrothermal method. The synthesized Bi2SiO5/g-C3N4 nanostructures were applied to the degradation of crystal violet (CV) dyes under visible-light irradiation, and the reaction rate constant was 0.1257 h−1, which was five times and three times higher than that of Bi2SiO5 and g-C3N4, respectively. From electron spin resonance and scavenger-test results, it was revealed that •O2 active species played a major role in the degradation of CV dyes, while other primary reactive oxygen species such as •OH, h+ and 1O2 played a secondary role (where 1O2 is the first excited state of molecular oxygen (O2), known as singlet oxygen). Wu et al. [111] developed Bi2SiO5-SiO2, and Bi12SiO20-SiO2 photonic crystal films were prepared by spin-coating Bi2SiO5 or Bi12SiO20 on SiO2 photonic crystals and used as photocatalysts for the degradation of RhB dyes under UV-light irradiation. The photon localization of SiO2 photonic crystal plays a key role in improving the light absorption of bismuth silicate. This study provides a simple approach to improve the light-harvesting efficiency of photocatalysts and expand the application of photonic crystals. However, the thickness of the bismuth silicate film exhibits dual photocatalytic activity; with the thickening of bismuth silicate, the light absorption of bismuth silicate increases, but the photon localization weakens.
Building heterojunctions or nanostructures to prolong the lifetime of electron/hole pairs is a very important strategy to endow them with excellent photoactivity. However, developing nanostructures of bismuth silicate with the same composition, but forming different crystal structures, and with suitable band gaps, remains challenging. Jia et al. [112] developed different crystal structures using a one-pot hydrothermal synthesis method without the addition of other inorganic materials. However, the dose of cetyltrimethylammonium bromide (CTAB) is the key to modulating the formation of bismuth silicate crystal phases with assembled nanostructures and their surface states. When the concentration of CTAB was 1.5–2 mmol, Bi2SiO5 nanoparticles were anchored on Bi12SiO20 or Bi4Si3O12 nanosheets. The obtained two kinds of bismuth silicate nanostructures, Bi2SiO5/Bi12SiO20, have rod-like structures, and Bi2SiO2/Bi4Si3O12 have flower-like nanostructures. Owing to these two nanostructures, the optimized bismuth silicate material exhibits high photoactivity and remarkable cycling stability. Specifically, the degradation rate of RhB under visible light can reach as fast as 15 min with a reaction rate constant of 0.34 min−1, which is 189 times faster than other reports. This one-pot synthesis strategy for developing single-component nanostructures has significant implications for designing other novel photocatalysts based on their natural multivalent states, or various crystals such as Mn-, Fe-, and V-based nanostructures. Liu et al. [113] reported novel nanostructures of Bi4O5Br2/Bi24O31Br10/Bi2SiO5 developed by in situ ion exchange. The successful formation of nanostructures between bismuth bromide and Bi2SiO5 can be attributed to their structural similarity, thermodynamic tolerance, and high lattice matching. The novel nanostructured photocatalysts have well-aligned span bands at their closely contacted interfaces and exhibit remarkable photoactivity for phenol degradation under visible light. This ternary nanostructure exhibits about 2.5 times higher photoactivity against phenol than bulk BiOBr. The detailed photocatalytic mechanism of Bi4O5Br2/Bi24O31Br10/Bi2SiO5 nanostructures shows that in this ternary nanostructure (Figure 8), Bi4O5Br2 and Bi24O31Br10 have narrower band gaps than BiOBr, so they can absorb more long-wavelength light and improve the light utilization rate. For example, Bi4O5Br2 nanosheets with vertically aligned facets exhibited ∼6 times greater visible-light photodegradation efficiency against BPA than that of BiOBr nanosheets [114].
Based on the electrical structure, the band potentials of Bi4O5Br2, Bi24O31Br10, and Bi2SiO5 are compatible, forming a heterojunction with well-aligned cross-bands when they are in close contact. The photogenerated electrons in the CB of Bi24O31Br10 easily migrate to the CB of Bi4O5Br. Meanwhile, the holes formed in the VB of Bi4O5Br2 are easily transferred to the VB of Bi24O31Br10 and occur also in the VB of Bi2SiO5. Thus, long-lived reactive photo-charges can be generated, allowing for improved charge separation at their interfaces. The nanostructure-enhanced photocatalytic activity can be attributed to (i) improved photo-utilization due to the presence of multi-components with narrower band gaps, (ii) significantly improved charge separation capability due to the well-aligned cross-band structure, and (iii) having a large specific surface area due to their layered features, which can generate abundant active sites for catalytic reactions. This study may help to design novel nanostructured photocatalysts with higher photoactivity.
The photoreactivity of Bi/Bi2WO6 was found to steadily increase from 12.3% to 53.1% with increases in the number of Bi nanospheres from 0 to 10 wt% due to the SPR effect of Bi nanospheres on the Bi2WO6 photocatalyst [63]. After the modification of Bi2WO6 microspheres with Bi nanospheres, the photo-generated carriers can transfer from the CB of Bi2WO6 microspheres to Bi nanospheres, retarding the recombination. In addition, the near-field enhancement produced on Bi nanospheres by the SPR effect can significantly enhance the energy of electrons, which then consequently boosts the separation and migration of photo-generated carriers. Then, highly concentrated reactive oxygen species such as •O2 and •OH radicals are produced in Bi/Bi2WO6 to oxidize NO (Figure 9).
Zhang et al. [115] developed Bi-induced Bi2O2CO3/Bi12SiO20 (i.e., Bi/Bi2O2CO3/Bi12SiO20) nanostructures grown in situ on Bi2O2CO3 on Bi12SiO20 by the oil bath method and applied them to the degradation of RhB and TC dyes. RhB and TC degradation by Bi/Bi2O2CO3/Bi12SiO20 nanostructures under simulated light are 12 times and 3.3 times higher than those of RhB and TC, respectively, which is attributed to the synergy effect of Bi heterojunction and the SPR effect. The trapping test results showed that •O2 played a key role in the photodegradation process. The band gaps of Bi2O2CO3 and Bi12SiO20 materials are 2.57 and 3.2 eV, respectively. The estimated CB for Bi12SiO20 is −0.65 V vs. NHE, which is more negative compared to Bi2O2CO3 (−0.59 V vs. NHE). Electrons in Bi12SiO20 can migrate to the CB of Bi2O2CO3. Moreover, since the redox potential of •O2 is −0.33 V with respect to NHE, the photoinduced e accumulation on the CB of Bi2O2CO3 further generates a large amount of O2-degrading dyes. Meanwhile, the VB (2.61 eV) of Bi2O2CO3 is still more negative than that of Bi12SiO20, and the h+ on the VB of Bi2O2CO3 can be transferred to Bi12SiO20, thereby directly degrading the dye. Therefore, the nanostructure composed of Bi12SiO20 and Bi2O2CO3 effectively promotes carrier separation and enhances photoactivity. In addition, the SPR effect of Bi on the surface of Bi2O2CO3/Bi12SiO20 nanostructures not only broadens the light absorption and improves the light utilization efficiency, but also increases the surface electron excitation and interfacial electron transfer rate. The local electromagnetic field induced by the SPR effect of metallic Bi also promotes the migration and separation of charge carriers in the Bi2O2CO3/Bi12SiO20 nanostructures. The Fermi level of Bi (−0.17 eV vs. NHE) is more negative than the CB of Bi12SiO20, •O2 can also be generated on Bi, and e from Bi12SiO20 can be moved to metallic Bi to facilitate carrier separation and enhance photoactivity. The synergy effect and SPR effect of Bi/Bi2O2CO3/Bi12SiO20 nanostructures are caused by metallic Bi, making Bi/Bi2O2CO3/Bi12SiO20 exhibit good photoactivity.

4.5. Bismuth and Bismuth-Rich Oxyhalides

Recently, a promising photocatalyst of the bismuth family, bismuth oxyhalide (BiOX, X = I, Cl, and Br), has been shown to induce more efficient charge separation due to its unique layered structure with an internal electrostatic field perpendicular to each layer, thus exhibiting significant photoactive performance [116]. Among them, BiOI has the smallest bandgap and strong absorption in the visible-light region. BiOI is a p-type semiconductor with a narrow bandgap of 1.8 eV, enabling it to absorb and utilize visible light. Therefore, it exhibits good photoactivity under sunlight exposure. Other forms of BiOI materials including the Bi4O5I2, Bi7O9I3, β-Bi5O7I, and α-Bi5O7I types have been widely reported [117]. However, the bandgap energy of these compounds is higher than that of BiOI, although lower than that of Bi2O3 [118]. Therefore, these materials are used as visible-light-induced photocatalysts. Interestingly, the structural and compositional features of bismuth iodide strongly affect its optical power, oxidative power, electronic properties, and other physicochemical properties, providing opportunities to obtain novel nanostructure photocatalysts for the efficient degradation of pollutants with different characteristics. Xiao et al. [119] reported that high-purity Bi4O5I2 with a hierarchical nanoflake structure can be easily obtained by reacting Bi3+, I, and OH under solvothermal conditions at pH values of 6–10. The as-prepared Bi4O5I2 nanoflakes have a band gap of 2.17 eV, a CB edge potential more negative than the superoxide radical reduction potential, and a specific surface area of about 39 m2g−1. Bi4O5I2 nanoflakes exhibited excellent photoactivity and mineralization efficiency for the degradation of PTBP under visible light, and the reaction rate was 6.8 times higher than that of BiOI microspheres. More importantly, the as-prepared Bi4O5I2 nanoflakes remain stable during the photoreduction process and can be reused.
Bi4O5I2/Bi4O5I2 nanostructures were synthesized by the electrostatic self-assembly method and used to degrade RhB under visible light irradiation [120]. In detail, two different Bi4O5I2 compounds were synthesized using an ionic iodine source, namely [Hmin]I (1-hexyl-3-methylimidazolium iodide) and a KI source. The Bi4O5I2(KI) was negatively charged, while [Hmin]I was positively charged, resulting in an electrostatic attraction between Bi4O5I2(KI) and Bi4O5I2([Hmin]I) to form a final product in the form of Bi4O5I2([Hmin]I) nanosheets introduced in the Bi4O5I2(KI) bulk. This final Bi4O5I2/Bi4O5I2 product exhibited higher photoactivity with an RhB degradation rate of up to 98.34%, which is higher than the physical mixture material (80.4%), which means that the heterojunction between Bi4O5I2([Hmin]I) and Bi4O5I2(KI) was a chemical force rather than a simple physical connection. Within the Bi4O5I2/Bi4O5I2 nanostructure, the photoinduced transfer of h+ and e, the h+ accumulated in the VB of Bi4O5I2(KI) (1.45 eV vs. NHE), was insufficient to oxidize H2O or OH to •OH since E0 (•OH/OH) (2.38 eV vs. NHE). At the same time, the e accumulated in the CB of Bi4O5I2 ([Hmin]I) (0.54 eV vs. NHE) was more positive than the E0(O2/•O2) (−0.04 eV vs. NHE), which means that no O2 was reduced to •O2. Therefore, the RhB adsorbed on the nanostructured material is mainly degraded by the h+ oxidative degradation accumulated in the VB of Bi4O5I2(KI). In addition, Bi4O5I2 ultra-thin nanosheets synthesized via processing the molecular precursor were found to photo-reduce CO2 into CO selectively with a photocatalytic activity of 19.82 μmol h−1 g−1 [121]. Finally, Yin et al. demonstrated that Bi4O5I2 has good catalytic activity in the degradation of not only MB and RhB but also methyl orange [122].
BiOCl has attracted great interest due to its remarkable photoactivity under UV irradiation [123], and a recent review has focused on this compound [124]. The high photoactivity of BiOCl can be attributed to its unique layered structure, i.e., [Bi2O2]2+ lamellae are interleaved by the double lamellae of Cl atoms, with an internal electrostatic field perpendicular to each layer [125]. This structural feature can effectively promote the transfer of electrons and holes generated inside the crystal face, promote charge separation, and improve quantum yield [126]. However, the broad band gap of BiOCl is 3.1–3.6 eV, which, due to its preparation method and morphology, can only absorb ultraviolet light similarly to TiO2, which also limits the effective utilization of solar energy. Therefore, to utilize the high quantum efficiency of BiOCl under visible light irradiation, an efficient approach is combining BiOCl with narrow bandgap semiconductors to form a nanostructure/heterojunction. For example, Li et al. [127] developed BiOCl/Bi24O31Cl10 nanostructures by an ionic liquid self-association method. BiOCl/Bi24O31Cl10 nanostructures were obtained by heating and burning the ionic liquid, which can also be used as the main fuel. The BiOCl/Bi24O31Cl10 nanostructure was applied for the degradation of MO and RhB, which was attributed to the narrow bandgap of 2.3 eV, which enabled efficient electron transfer from the CB of Bi24O31Cl10 to BiOCl and improved the separation efficiency of charge carriers. The BiOCl/Bi24O31Cl10 nanostructure containing 60.4% BiOCl and 39.6% Bi24O31Cl10 exhibited the highest photocatalytic performance. Among these dyes, BiOCl/Bi24O31Cl10 nanostructures exhibited remarkable adsorption activity for cationic dyes of RhB due to their negative surface charges. Furthermore, the main active species responsible for the efficient degradation of pollutants are holes and superoxide radicals involved in the photocatalytic process. Both Bi4O5Cl2 and Bi4O5l2 proved to be efficient for photocatalytic water splitting for hydrogen evolution, but Bi4O5Cl2 is the best for H2 production [128].
BiOCl and related compounds are best prepared with an exposed (001) surface to obtain the best activity, owing to the effect of the self-induced electric field by this polar surface [129]. In addition, surface OVs play an important role. The OVs extend the visible-light adsorption range from 200 nm to 800 nm because of the formation of a localized state [130], and they significantly improve the UV-light harvesting ability [131]. In this work, Dong et al. synthesized BiOCl with OVs by solvothermal-induced hot ethylene glycol reduction at 160 °C for 12 h, followed by mixing with H2O2, drying, and subsequent treatment at 300 °C in an O2 atmosphere for 4 h. They demonstrated that the solid solution with OVs increased the removal ratio of toluene from 52.5% to 64% and that of NO from 33.2% to 43.5% in air under 360 nm UV-light irradiation for 15 min. They also showed that the main reason for the important improvement of the toluene degradation comes from the shortening of the toluene degradation pathway via the surface OVs, which enables the production of radicals with high oxidation capability for the accelerated chain scission of the ring-opening intermediates. Zhao et al. showed that the photocatalytic degradation of dyes with Bifocal OVs is greatly enhanced when H2O2 is added, because the activation of H2O2 increases the production of •O2¯, (H2O2 + h+ → •O2 + 2H+) [130]. Again, the localized state introduced by the OVs is important here, because the electrons in the localized state are transferred to the CB by interband excitation. Then electrons are trapped by O2 and generate •O2¯. Besides OVs, co-doping is another way to create a localized level in the band gap of BiOCl to expand the light absorption region. Wang et al. prepared co-doped BiOCl nanosheets using a simple hydrothermal route. They exhibited outstanding photocatalytic performance in degrading BPA under visible light irradiation with a degradation rate of 3.5 times higher than that of pristine BiOCl [132].
Jia et al. used diatomite as a solid dispersant to immobilize BiOCl microspheres. BiOCl/60% diatomite presented 94% removal efficiency for CIP under simulated solar light within 10 min irradiation, and also presented a 42.9% total organic carbon (TOC) removal after 240 min, and good reusability [133].
Alansi et al. [134] synthesized the OVs rich in BiOCl0.8Br0.2 flower-like materials under direct sunlight exposure within 10 min and noted that they changed color from white to black (i.e., UV-BiOCl0.8Br0.2), which the authors applied for RhB degradation under visible light. When pristine BiOCl0.8Br0.2 was prepared using low-frequency UV radiation, pristine BiOCl0.8Br0.2 nanoflowers with a highly exposed facet (001) on the surface were obtained. The (001) facet of pristine BiOCl0.8Br0.2 has a tight structure, in which the high-density oxygen atoms are exposed with long, weak Bi-O bonds, facilitating the escape of oxygen atoms from the surface, creating OVs behind them. Due to the abundance of OVs of UV-BiOCl0.8Br0.2, the nanoflowers enhanced photocatalytic activity because the vacancies serve as electron capture centers. The proposed photocatalytic mechanism is as follows: the presence of Br in the BiOCl0.8Br0.2 material significantly increased its surface area and decreased the Bi-O bond energy on BiOCl, which in turn provides the formation of OVs resulting in wider visible light absorption and a fast charge-transfer rate. Water molecules are rapidly adsorbed on the OV sites of the aqueous solution, and adsorbing on the (001) BiOCl0.8Br0.2 surface leads to the formation of a layer of hydroxyl groups. Hydroxyl groups increase the length of the Bi-O bond and thus reduce its energy, which provides for the escape of oxygen atoms from the surface upon exposure to an energy source, leaving OVs behind them. Therefore, by exposing pristine BiOCl0.8Br0.2 to low-energy irradiation, such as the UV component of natural sunlight irradiation, the hydroxyl readily exits and allows the regeneration of OVs on the surface, which changes color from white to black. The weakening of the Bi-O bond in the presence of Br has also been used to synthesize BiOBr ultrathin nanosheets with abundant surface Bi vacancies (VBi-BiOBr) by reactive ionic liquid ([C16mim]Br)-assisted synthesis at room temperature [135]. With the advantages of optimized CO2 adsorption, activation, and CO desorption, VBi-BiOBr UNs can deliver a 3.8-times-improved CO formation rate relative to BiOBr nanosheets, with a selective CO generation rate of 20.1 μmol g–1 h–1 in pure water. Another example of CO2 conversion is given by BiOBr atomic layers with OVs obtained by ultra-sonication exfoliation followed by UV irradiation. The visible-light-driven conversion rate of CO2 to CO was increased to 87.4 µmol g−1 h−1, much higher than the value obtained in the absence of VOs [136].
Wang et al. fabricated OV-rich sulfur-doped BiOBr nanosheets through a facile one-step solvothermal method [137]. The synergistic effect between S doping and oxygen vacancy led to superior photoactivity for non-dye organic contaminants. In particular, under visible light irradiation, the optimal BB-5S sample exhibited 98% degradation efficiency of 4-chlorophenol within 120 min.
The solid solution BiOClxBr1−x exists in the whole range 0 ≤ x ≤ 1, so that the ratio of Cl and Br can be tuned to decrease the band gap and thus improve the photocatalytic activity. In particular, Yang et al. prepared this solid solution through a glycol-assisted hydrothermal process [138]. The degradation rate of methyl orange (in aqueous solution) reached a maximum value at x = 0.5.
In BiO(ClBr)(1−x)/2Ix solid solutions, the introduction of I has two advantages. First, the ionic radius of I is larger than that of Cl or Br, so that the introduction of I leads to a lattice dilatation in the c axis, which reduces the energy barrier at the interface of different crystal planes and reduces the recombination rate of the photo-electrons and holes. Second, the CB edge is mainly composed of Bi 6p and is thus not significantly dependent on x. On the other hand, the valence band edge is mainly composed of the p states of the oxygen and halogens, so it is modified, and actually increases with x. Consequently, the band gap decreases with the introduction of I and can be varied from 2.88 to 1.82 eV depending on x. As a result, BiO(ClBr)(1−x)/2Ix showed improved photoactivity for the degradation of 2-propanol to acetone and CO2, under visible light [139].
Dong et al. synthesized four-layered bismuth oxyhalides BiOX and BiOXO3 (X = Br, I). They found that the order of the photocatalytic performance for water splitting (including the carrier’s lifetime, photocurrent density, and H2 evolution rate) is BiOBrO3 > BiOI > BiOIO3 > BiOBr, emphasizing the role of the polar electric field [140]. The remarkable performance of BiOBrO3 is due to the inhibition of the recombination of the charge carriers by the internal polar electric field along the (001) direction. On another hand, the built-in electric field has no impact on the recombination rate in bulk BiOX, owing to the mirror symmetry. However, the recombination is hindered in BiOI by the surface polar electric field, which breaks the mirror symmetry. This polar behavior of BiOXO3 is due to the difference between the crystallographic structure of BiOX and BiOXO3. They all crystallize in the sillenite structure, but the double X layers in BiOX are replaced by double [XO3] layers in BiOXO3. The XO3 units have a trigonal pyramidal structure that generates electric dipoles. For the same reason, BiOIO3 is a polar material, with the c-axis as the polar axis, which explains its high photocatalytic activity [141,142]. Chen et al. synthesized BiOIO3 single crystal nanoribbons along the (001) direction to take advantage of the strong polarity of the IO3 units. This polarity acted collaboratively with surface oxygen vacancies to boost CO2 reduction [143]. Huang et al. found that the activity of BiOIO3 for photocatalytic water splitting can be increased by V5+ ion doping into IO3 pyramidal units. The •O2 and •OH evolution rates of BiOI0.926V0.074O3 increased by ∼3.5- and ∼95.5-fold, respectively, with respect to BiOIO3 [144]. Another example of the beneficial effect of the polarization-induced electric field is the modification of porous BiVO4 microtubules with inorganic acids. The generation of free hydroxyl radicals by the ionization of hydroxyl groups in the modified inorganic acid increased the intensity of the surface electric field, enhancing their reactivity toward CTC degradation [145].

5. Type of Photocatalytic Mechanism of Bismuth-Based Photocatalysts

5.1. p–n Junction

The synthesis of nanostructures with highly reactive exposed faces and p–n junctions is of great interest for semiconductor photocatalysis. The construction of nanostructured semiconductor junctions has been very active recently because of their perfect effect in promoting the separation of photogenerated charge carriers and enhancing photocatalytic reactions. In general, nanostructured catalysts containing p–n junctions with direct contact between p-type and n-type semiconductors have drawn much devotion due to their large potential gradients, and the built-in electronic field established at their junction level can induce efficient charge transfer and separation. The main effective strategy to enhance photocatalytic activity is crystal-facet engineering. We have already mentioned the case of BiOX/CuFe2O4 [58]. Another example is provided by Cai et al. [146], who reported nanostructures of β-Bi2O3/Bi2O2CO3 and α-Bi2O3 prepared by a rational calcination process of Bi2O2CO3, used as photocatalysts for MB degradation under visible light irradiation. A p–n junction was successfully created by the proposed synthetic procedure. The β-Bi2O3/Bi2O2CO3 (at 300 °C) nanostructure reduces recombination by promoting the separation of photogenerated electrons and holes, showing higher MB degradation efficiency than Bi2O2CO3 and α-Bi2O3. When β-Bi2O3 is in contact with Bi2O2CO3, the CB potential of Bi2O2CO3 is more positive than that of β-Bi2O3 with the adjustment of the Fermi level. Therefore, electrons generated on β-Bi2O3 CB can be transferred to Bi2O2CO3 by the electric field formed inside. Therefore, the formation of a p–n heterojunction of β-Bi2O3/Bi2O2CO3 can effectively separate electron-hole pairs and suppress the undesired recombination of electrons and holes. The separated electron-hole pairs are then freely transferred to the surface to react with the adsorbed dye molecules, thereby enhancing the photocatalytic activity of the nanostructures.
Huang et al. [147] pioneered the in situ-constructed BiOI/Bi12O17Cl2 nanostructures consisting of BiOI nanosheets grown vertically on the surface of the Bi12O17Cl2 plate, forming a unique front-coupling nanostructure that enables high exposure of the (001) facet reaction exposed surface of BiOI. The photocatalytic behavior of various industrial pollutants such as 2,4-DCP, RhB, BPA, and antibiotics (TCHC) was tested on BiOI/Bi12O17Cl2 nanostructures. The BiOI/Bi12O17Cl2 nanostructures not only exhibited significantly enhanced photoactivity but also exhibited strong non-selective photooxidation ability under visible light irradiation. The BiOI/Bi12O17Cl2 nanostructures exhibit the benefits of facilitating the separation and transfer of charge carriers, which originate from the BiOI (001) active facet and p–n junction responsible for high photoactivity. The highly promoted photoactivity of BiOI/ Bi12O17Cl2 nanostructures is mainly credited to the following aspects: (i) Bi12O17Cl2 can serve as an excellent substrate to support and uniformly distribute BiOI nanosheets, which helps to increase the specific surface area for enhanced absorption and reaction sites. Nevertheless, the enhanced level of the surface area is lower than the enhanced level of photoactivity. (ii) The main advantage of BiOI/Bi12O17Cl2 nanostructures is heterojunction formation, which plays an important role, due to the front-side surface coupling assembly of BiOI/Bi12O17Cl2 nanostructures enabling the (001) crystal planes to be more exposed. Due to the strong light absorption ability, a large number of photogenerated carriers will appear under visible light irradiation. (iii) Driven by the strong self-built electric field from the (001) active surface of BiOI, these induced electrons and holes flow from the interior of the BiOI nanosheets, densely migrate to the surface, and then accumulate on opposite surfaces, such as the top and bottom surfaces, respectively (Figure 10).
Therefore, the photoinduced electron-hole pairs of BiOI and BiOI/Bi12O17Cl2 also contribute to high photoactivity. p-type BiOI and n-type Bi12O17Cl2 can form a stable p–n junction. The CB and VB of BiOI before contact are lower than those of Bi12O17Cl2. After the p–n structure is built, the energy level of Bi12O17Cl2 decreases, while the energy level of BiOI increases until BiOI and Bi12O17Cl2 reach the Fermi level equilibrium. The bottom of the CB of BiOI and the top of the VB can quickly migrate to the bottom of Bi12O17Cl2, while the holes generated by the VB of Bi12O17Cl2 are transferred to the VB of BiOI. Therefore, the electron-hole pairs are effectively separated at the p–n junction of BiOI/Bi12O17Cl2 nanostructures, and the holes accumulated in the Bi12O17Cl2 VB directly oxidize the pollutants. The electrons accumulated at the CB of BiOI are further converted into O2 with a strong oxidizing ability, which subsequently induces the decomposition of various pollutants. This work delivered a new avenue for us to design novel nanostructured photoactive materials with integrated p–n junctions and numerous active exposed facets. We have already noted above that BiPO4 has a too-large band gap to have good photocatalytic properties in visible light. BiPO4 is n-type. A solution is then to modify the surface of BiPO4 with BiOBrxI1−x, which is p-type, to form a p–n heterojunction, and optimize the band gap by the choice of x. The 5% BiPO4–BiOBr0.75I0.25 heterojunction showed the highest photocatalytic activity in the reduction of CO2 [148]. After 4 h of visible light irradiation (λ > 420 nm), the yield of CO and CH4 reached 24.9 and 9.4 μmol g−1, respectively.
Tang et al. [149] reported the construction of BiOI/tetrapod-like ZnO whiskers (T-ZnOw). The p–n junction photocatalysts with different Bi/Zn molar ratios were prepared by the in situ precipitation of BiOI on T-ZnOw templates and applied to degrade RhB and oxytetracycline (OTC) under visible-light irradiation. Compared with other samples, the 1:10 nanostructured photocatalysts with different Bi/Zn molar ratios exhibited the highest photoactivity, namely 97.1% RhB and 88% OTC. This is endorsed by the large specific surface area and efficient separation of charge carriers caused by the formation of p–n heterojunctions between T-ZnOw and BiOI. Figure 11a shows the energy bands of BiOI and T-ZnOw before the formation of the BiOI/T-ZnOw nanojunction. BiOI and T-ZnOW have nested energy levels that are not conducive to the transfer of system-generated charge carriers. Since the BiOI/T-ZnOw nanostructure has higher photoactivity and photocurrent response than T-ZnOw and BiOI, it can be inferred that a p–n junction is formed. T-ZnOw is an n-type semiconductor with a Fermi level close to the CB, while BiOI is a p-type semiconductor with a Fermi level close to the VB. When BiOI and T-ZnOw are in close contact, a p–n junction is formed (Figure 11b).
The electrons are transferred from the T-ZnOw near the p–n junction, and at the same time, the holes are transferred from BiOI to T-ZnOw, causing the positive charge region in T-ZnOw to reach equilibrium with BiOI, and the built-in electric field direction from T-ZnOw to BiOI is constructed at the same time. The band positions of BiOI and T-ZnOw shift up and down together with the Fermi level, and photoactivity occurs as follows: (i) BiOI is excited under visible light, causing electrons to move from VB to CB. Then, since the CB of BiOI is more negative compared to that of T-ZnOw, electrons are easily moved to the CB of T-ZNOw from the CB of BiOI. Furthermore, the built-in electric field can provide the migration of photogenerated electrons from BiO to T-ZnOw. The holes remain in the VB of BiOI, which enables the efficient separation of electrons and holes in BiOI. Therefore, efficient charge carrier separation and more electrons and holes can participate in the photocatalytic process, thereby enhancing the photoactivity. The electrons in the CB of T-ZnOw can react with dissolved O2 to generate O2, and then generate OH from O2 through a reduction process. The holes in the VB of BiOI directly participate in the oxidation of OTC. Therefore, •O2, •OH, and h+ jointly participate in the degradation of OTC.
Nie et al. [150] developed a p–n junction-derived flower-like CeO2-δ (n-type) coupled to β-Bi2O3 of p-type (i.e., β-Bi2O3/CeO2-δ) nanostructures via the thermal decomposition of Bi/Ce precursors (Figure 12). This β-Bi2O3/CeO2-δ nanostructure was used for NO removal under visible light.
The excellent photoactivity of β-Bi2O3/CeO2-δ nanostructures is attributed to the synergistic effect of oxygen vacancies and p–n junctions. The associated oxygen vacancies not only improve the utilization of visible light and facilitate the separation of electron-hole pairs, but also enhance the adsorption of NO and the activation of O2. In fact, the synergistic effect of the p–n heterojunction through the p–n junction favors the interfacial migration of charge carriers, and oxygen vacancies can induce more active radicals. The nanoflower-like β-Bi2O3/CeO2-δ nanostructures exhibit excellent photoactivity, which can completely remove NO and inhibit NO2 production. The authors verified the reaction products by in situ fast Fourier infrared analysis, showing that the main product of nitrate is formed during the photocatalytic process.

5.2. n–n Junction

Su et al. [151] developed a post-calcination process for hydrothermal synthesis. One-dimensional, rod-like BiOI/Ag2Mo2O7 nanostructures can reduce the photocatalytic activity of RhB and TC by 70- and 16-fold compared with Ag2Mo2O7, which is attributed to the efficient separation of photogenerated charges. The reason is to form an n–n junction between BiOI and Ag2Mo2O7. From the free radical trapping test results, it can be concluded that •O2 and h+ species play a major role in photoactivity. In addition, this nanostructure has a photodegradable TC solution, which is basically harmless to Escherichia coli. Figure 13a,b depict the possible photocatalytic mechanism of BiOI/Ag2Mo2O7 nanostructures, showing the energy band positions of BiOI and Ag2Mo2O7 before and after contact. When BiOI is combined with Ag2Mo2O7, an n–n junction is formed at the contact interface. Since the Fermi level of BiOI is lower than that of Ag2Mo2O7, the electrons in BiOI can be transferred to Ag2Mo2O7, thus generating a positive consumption layer on one side of BiOI and a negative consumption layer on the other side of Ag2Mo2O7. When the Fermi levels of the two components reach equilibrium, an internal electric field from BiOI to Ag2Mo2O7 will form at the interface. Under illumination, the photogenerated electrons with strong yield in BiOI CB can reduce O2 to O2•− (O2/O2•−) (−0.33 eV vs. NHE), and the photogenerated holes in Ag2Mo2O7 VB have strong oxidizing ability, which can direct the oxidation of contaminants to non-toxic products. Therefore, the n–n junction can facilitate the separation of photogenerated charges, thereby enhancing the photocatalytic activity.

5.3. Z-Scheme

Li et al. [152] constructed ternary nanostructures composed of AgBr anchored on BiOI/g-C3N4 nanostructures and applied them to degrade MO (20 mg L−1) under visible light irradiation. The MO degradation rate of AgBr/BiOI/g-C3N4 nanostructures on the nanostructured catalyst reached 93.41% within 120 min, which is attributed to the double Z-type heterojunction between AgBr, BiOI, and g-C3N4, which has a strong Ag electron capture effect (Figure 14). It was concluded that the main active species was •O2, and h+ also played a role. A double Z-type electron transfer mechanism is formed between AgBr, BiOI, and g-C3N4. Under illumination, the electrons accumulated on AgBr can easily react with the attached Ag+ to form metallic Ag, so the AgBr/BiOI/g-C3N4 system is transformed into Ag/AgBr/BiOI/g-C3N4. Metallic Ag has good electron-trapping ability; it can capture electrons to generate active •O2 from the CB of AgBr for degrading MO molecules in solution. However, due to the small doping connection of BiOI and AgBr, the degradation effect is limited. Finally, for the potential of oxidation to •OH (•OH/H2O = +1.99 eV and •OH/OH = +2.4 eV vs. NHE), the g-C3N4 VB is lower, but it is higher for BiOI and AgBr. Therefore, the amount of •OH produced is small, and it is concluded from the scavenger experiments that •OH can hardly degrade the pollutants during the photocatalytic reaction.
Graphitic carbon nitride (g-C3N4) has photocatalytic activity for BPA degradation [153,154]. g-C3N4 is well matched with Bi2WO6 composites. In particular, the Z-scheme g-C3N4-Zn/Bi2WO6 synthesized by a two-step solvothermal method followed by a calcination process, using 2.0 g of dicyanamide as the precursor for g-C3N4, photodegraded 93% of the BPA within 120 min [155]. Even better results were obtained with Bi2WO6/g-C3N4/black phosphorus quantum dots (BPQDs) composites fabricated by the hydrothermal reaction of Bi2WO6 and g-C3N4 and a succedent BPQDs modification [156]. This composite with a direct dual Z-Scheme configuration showed photocatalytic activity for BPA degradation in visible light (95.6%, at 20 mg L−1 in 120 min), higher than that of Bi2WO6 (63.7%), g-C3N4 (25.0%), BPQDs (8.5%), and Bi2WO6/g-C3N4 (79.6%), respectively.
Deng et al. [157] fabricated a Z-scheme black BiOCl-Bi-Bi2O3/rGO heterojunction, where rGO and metallic Bi serve as charge-transfer channels between black BiOCl and Bi2O3. The black BiOCl-Bi-Bi2O3/rGO0.4 shows the highest visible-light photocatalytic activity with almost complete degradation of 2-nitrophenol, owing to the proper bandgap match between black BiOCl and Bi2O3, multiple charge-transfer channels via Bi-bridge and rGO, and efficient charge separation.
Recently, Ag/SnO2-x/Bi4O5I2 showed high efficiency in degradation and antibacterial properties, owing to the Z-scheme of this ternary composite. The optimum sample of 3% Ag/SnO2−x/Bi4O5I2 can degrade 80% TC in 120 min, inactivate Escherichia coli (E. coil) in 15 min, and Staphylococcus aureus (S. aureus) in 20 min under LED light [158].
Zhang et al. [159] pioneered microwave-hydrothermally synthesized Bi2SiO5/Bi2SiO20 nanostructured photocatalysts by bismuth nitrate and nano-SiO2 as precursors and applied for the degradation of RhB and MB dyes under UV-light irradiation. The results show that the photocatalytic activities of the Bi2SiO5/Bi2SiO20 nanostructures of RhB and MB dyes are 3-fold and 4.3-fold higher than that of Bi12SiO20, which are attributed to their large specific surface area, smaller particle morphology, and good crystallinity through their heterogeneity. The heterojunction facilitates an efficient charge separation capability. The trapping test results show that superoxide radicals and holes play a major role in the photoactivity. The Bi2SiO5/Bi2SiO20 nanostructured photocatalyst has Z-type photocatalytic activity. However, the oxidizing power of the photogenerated holes at the Bi2SiO20 VB is not sufficient to oxidize H2O to •OH because its potential is shallower than that of •OH/H2O (2.8 eV vs. NHE). The CBs of Bi2SiO5 and Bi2SiO20 are not sufficiently negative compared to the standard reduction potential O2/O2•− (−0.33 eV vs. NHE), indicating that electrons cannot be captured by O2 in solution to form reduced •O2. According to the trapping test and ESR results, h+ and •O2 play important roles in photodegradation; therefore, another common possible mechanism is a Z-type heterojunction. The photogenerated electrons on Bi2SiO5 transfer from the CB of the photogenerated holes to the VB and can oxidize H2O to •OH, while the photogenerated electrons on the CB of Bi12SiO20 cannot reduce O2 to •O2. This finding contradicts the test results, indicating that the Z-scheme system also cannot explain the degradation mechanism of Bi2SiO5/Bi12SiO20 nanostructures. Surface oxygen vacancies are considered shallow donors for semiconductor photocatalysts and can serve as adsorption and reaction sites, as shown in Figure 15a. For example, oxygen vacancies can dynamically capture directly excited electrons from the CB and can directly activate O2 to form reactive oxygen species (O2). In fact, Bi12SiO20 and γ-Bi2O3 have similar crystal structures, in which 80% of tetrahedral sites are occupied by Bi3+ and 20% of vacancies (Si) in the γ-Bi2O3 crystal structure. Oxygen vacancies in Bi12SiO20 are mainly concentrated in tetrahedral silicate groups (SiBiO4) [160]. Therefore, the photocatalytic mechanism of Bi2SiO5/Bi12SiO20 nanostructures can have multiple charge-transfer channels, as shown in Figure 15b. The major contributions are: (i) the existence of surface oxygen vacancies conducive to the enrichment of O2; (ii) the photo-induced generation of holes and electrons can control the direction of charge transfer from Bi12SiO20 to Bi2SiO5 and also control the electron transfer between O2 and Bi2SiO5/Bi12SiO20 nanostructures, thereby improving carrier separation in photocatalysis; and (iii) surface oxygen vacancies continuously capture and release electrons to generate new active species, acting as d-electron carriers, which then donate electrons to the anti-bonding orbital of O2, reducing it to •O2. The relevant reactions involving photocatalytic removal of dyes are as follows:
Bi2SiO5/Bi12SiO20 + hν → (Bi2SiO5/Bi12SiO20 + h+) + e
O 2 + e OV O 2
h+(Bi2SiO5) → h+(Bi12SiO20)
dye + •O2 → degradation products
dye + h+ → degradation products
However, it remains questionable why •O2can still be generated when both Bi2SiO5 and Bi12SiO20 have negative CB potentials compared to the O2/•O2 reduction potential. Finally, the dye is degraded by the interaction of O2 and h+. This is the unique reaction mechanism of the photocatalytic process proposed in their study.
Other members of the layered bismuth oxide family were also successfully used for the construction of Z-scheme heterojunctions with high photocatalytic activity; in particular, Bi2MoO6. This is a member of the Aurivillius family, a potential candidate as an excellent photocatalyst and solar-energy-conversion material for water splitting and the degradation of organic compounds under visible-light irradiation [161,162]. We guide the reader to the excellent review on the recent advances in the photocatalytic degradation of organic pollutants using Z-scheme Bi2MoO6-based heterojunctions [54].

5.4. S-Scheme

The S-scheme effect promotes the interface charge transfer and can be used to improve photocatalytic activity. Lu et al. synthesized a Bi2O3/Bi2SiO5 p–n heterojunction photocatalysts [163]. The p–n heterojunction was formed by increasing the amount of nano-SiO2 precursor, which transformed α-Bi2O3 into β-Bi2O3. Dou et al. [164] developed Bi2O3-related oxygen vacancies coupled with Bi2SiO5 microspheres to self-assemble to form OVs-Bi2O3/Bi2SiO5 heterojunctions via a simple one-pot solvothermal process. The OVs-Bi2O3/Bi2SiO5 nanostructures consist of one-micron-sized microspheres for the degradation of MO dyes under visible light. The synergistic effect of Bi2O3 and Bi2SiO5 greatly improved the removal rate of MO, and the carrier separation and transfer of the OVs-Bi2O3/Bi2SiO5 nanostructure were associated with a ladder mechanism, which endowed the OVs-Bi2O3/Bi2SiO5 nanostructure with higher photoactivity as compared to bare Bi2O3. After the combination of Bi2O3 and Bi2SiO5, due to the interfacial electric field gradient in the OVs-Bi2O3/Bi2SiO5 nanostructure, the Fermi energies of the two materials are arranged into a new energy band structure at their interface (Figure 16). The photogenerated electrons on the CB of Bi2SiO5 are transferred to the CB of Bi2O3. The relatively useless photogenerated electrons on Bi2SiO5 CB can recombine with relatively useless holes on the Bi2O3 VB. The holes on the Bi2SiO5 VB can oxidize H2O/OH to form OH radicals. Therefore, MO can degrade electrons and holes in different spatial regions through •O2, •OH, or h+ oxidation pathways.
Note that S-scheme heterojunctions have been also made between a Bi-based compound and another compound. An S-scheme heterojunction was formed by depositing Bi2O3 nanoplates on TiO2 nano-fibers [165]. This Bi2O3/TiO2 heterojunction demonstrated good activity to remove phenol under visible light. The S-scheme heterojunction formed by the combination of Bi2WO6 and a metal–organic framework (NH2-MIL-125(Ti)) displayed enhanced photocatalytic activity for the removal of RhB and TC under visible light irradiation [166]. Li et al. [167] fabricated a black phosphorus/BiOBr S-scheme heterojunction by a convenient liquid-phase ultrasound combined with a solvothermal method. The photocatalytic performance of this heterojunction for the TC degradation, oxygen evolution, and H2O2 production rate of Sol-10BP/BiOBr was 7.8, 7.0, and 2.6 times that of pure BiOBr, respectively. Xie et al. fabricated an S-type g-C3N4/Bi/BiVO4 photocatalyst with the aid of a facile substrate-directed liquid phase deposition route [168]. In addition to the S-scheme effect promoting the interface charge transfer, this structure utilized the SPR effect of bismuth. This effect accelerates the separation of the photo-generated carriers [169], already evidenced in heterojunctions of Bi and BiOCl [21,170,171]. Moreover, the excellent NO removal efficiency observed with nanoparticles of Bi on g-C3N4 was achieved for the optimized size of the Bi nanoparticles (12 nm) [172]. Owing to these synergetic effects, the g-C3N4/Bi/BiVO4 exhibited superior performance toward artificial carbon cycling.
A recent outstanding example of the efficiency of the S-scheme is the Bi4Ti3O12/ZnIn2S4 S-scheme heterojunction, which demonstrated an outstanding hydrogen production efficiency of 19.8 mmol h−1 g−1 under visible light irradiation [173].

6. Other Strategies for Enhance the Photocatalytic Activity

6.1. Doped-Bismuth-Based Photocatalyst

Chen et al. [174] reported the synthesis of C-N-doped β-Bi2O3 nanosheets by solvothermal calcination using poly(aniline-co-pyrrole) as C and N sources and applied them to degrade 17α-ethynylestradiol. The photodegradation rate of 17α-ethynylestradiol of C-N doped β-Bi2O3 nanosheets was 98.86% within 20 min under visible light irradiation, which was attributed to the high specific surface area and hydrophilicity of carbon and C and N doping and the post-induced narrow bandgap in β-Bi2O3. Shahid et al. [175] reported simple wet-chemically derived gadolinium (Gd)-doped BiFeO3 nanoparticles grafted onto rGO using an ultrasonic strategy, and formed Gd-doped BiFeO3/rGO nanostructures, which were applied by solar irradiation for the degradation of MB dye. Compared with Gd-doped BiFeO3 and bare BiFeO3, Gd-doped BiFeO3/rGO nanostructures exhibited superior photocatalytic activity. The Gd-doped BiFeO3/rGO catalyst removed 87% of MB dyes in 120 min with a rate constant of 0.016 min−1, while Gd-doped BiFeO3 and bare BiFeO3 degraded only 66% (0.008 min−1) and 55% (0.003 min−1) of the MB dye, due to the synergistic effect of Gd-doping and rGO inclusion, resulting in a red-shift in light absorption. Due to the nanoscale features of the structures, the nanostructures fabricated by this synthesis process can suppress charge carrier recombination and charge-transfer resistance by enhancing electronic conductivity and diffusion properties. Shamin et al. [176] developed a low-temperature hydrothermal synthesis of 10% Gd, Cr-doped Bi25FeO40 for the degradation of RhB and MB. Gd-Cr-doped Bi25FeO40 exhibits a low band gap of 1.76 eV and higher photocatalytic degradation performance for RhB and MB, which is attributed to the phase distribution, regular power-like morphology, reduced electron-hole recombination, and lower bandgap. Therefore, Gd-Cr-doped Bi25FeO40 has been shown to be a compelling energy-saving and low-cost strategy for the preparation of sillenite-phase bismuth ferrite as a promising photocatalyst.

6.2. Ligand Modification Strategy

Tien et al. [177] reported that Bi12O17Cl2 nanowires were synthesized by the chlorination of α-Bi2O3 at 400 °C and consisted of tetragonal structures with a length of 15 μm and a diameter of 400 nm. Bi12O17Cl2 nanowires were prepared by a chlorination method. Typically, α-Bi2O3 nanowires are directly reacted with HCl and converted into Bi12O17Cl2 nanowires, as follows:
6α-Bi2O3 +2HCl → Bi12O17Cl2 + H2O
Bi12O17Cl2 nanowires have red emission at 746 nm and strong green emission at 568 nm at room temperature, which is a hallmark of visible-light-emitting materials for photocatalytic applications, because the synthesized Bi12O17Cl2 nanowires exhibit a narrow bandgap of 2.28 eV.

7. Future Prospects and Expectations

Regarding the huge number of research works published in the last years, Bi-based materials have received notable attention as possible active solar photocatalysts and demonstrate high photocatalytic activity when used for the degradation of environmental pollutants. Bismuth-based photocatalysts have proved to be a promising class of materials for a variety of energy- and environment-related applications due to their unique, layered structures, excellent physicochemical properties, and tunable electronic structures. However, a single component of bismuth-based materials often suffers from several inherent disadvantages, including low light-harvesting efficiency, few active sites, and the recombination of charge carriers. To overcome these shortcomings, efforts have been devoted to optimizing the photoactivity of bismuth materials by coupling them with metallic or semiconducting materials to alter their band energies, including rational structural design, compositional tuning, electronic structure tailoring, and interfacial engineering. Furthermore, Bi-based photocatalysts with rationally high catalytic yield still require further alteration to enhance their photocatalytic activity. To satisfy this objective, it was recently revealed that microstructures synthesized in a controlled manner using an appropriate bromide source exhibit improved photocatalytic performance induced by the formation of hierarchical 3D flower-like open petal structures [178]. Controllable morphological features of microparticles have always been an important research topic in material synthesis, which allows us not only to perceive unique features but also to obtain desirable physicochemical properties. For instance, several workers demonstrated the effect of solvent on the morphological characteristics of BiOBr by a solvothermal method and showed that the viscosity of the solvent causes morphological changes in the formation mechanism of photocatalytic materials [179,180,181].
Regardless of the good progress in bismuth-based photocatalysts, the nature of the active sites in these nanostructures remains unclear. It is highly desirable to understand the associated photocatalytic mechanism at the nanoscale level during multiple photocatalytic applications. More attention should be paid to the determination and quantification of active sites by calculation and direct experimental analysis. Further in situ characterization and calculations close to realistic conditions are needed to gain an atomic-level view of the relationship between active sites and photoactivity, which would be a merit for better design and adaptation of bismuth-based photocatalysts. As can be envisaged in this short overview, one of the main approaches to the development of highly efficient photocatalytic material passes through the fabrication of complex heterostructures. Table 1 summarizes recent studies on nanostructured Bi-based photocatalysts for pollutant degradation [56,164,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196,197]. The excellent photocatalytic performance of Bi-based heterojunctions for the degradation of cationic pollutants under visible-light irradiation is superior to that of single sheets, which is ascribed to the efficient charge separation and transfer across the phase junction.
Recently, Bi et al. [184] constructed a Bi2WO6/ZnIn2S4 phase junction with a Z-scheme structure showing high photocatalytic activity due to the rapid transfer of carriers, which inhibits the recombination of e and h+. Thus, the phase-junction approach is opening new avenues for the development of efficient photocatalysts for water purification and energy conversion. Highly active S-scheme heterojunctions show outstanding photocatalytic activity, in which both •O2 radical attack, H+ direct oxidation, and OH oxidation are the processes implicated in the removal of pollutants. Ligand modification based on chlorination is another successful strategy to enhance the photocatalytic activity of BBNs [177].
There is a separate class of bismuth-based photocatalysts—alkaline earth metal (Mg, Ca, Sr, Ba) bismuthates (BiO3 or Bi2O62–)—with bismuth in its pentavalent state [198]. Magnesium bismuthate with the composition MgBi2O6 is a degenerate semiconductor with a bandgap of only 1.8 eV. Shtarev et al. [199] examined the sillenite structure MgBi12O20 to probe the effect of the cationic composition of this magnesium bismuthate on its photocatalytic properties.
The photocatalytic activity of Ca3Bi8O15 [200] was estimated from the decomposition of various pollutants, e.g., MO (6.1 × 10–5 mol L−1), RhB (3.0 × 10–4 mol L−1), and 4-CP (3.0 × 10−4 mol L−1) in aqueous media irradiated with visible light (420 nm < λ < 800 nm) at room temperature. The photocatalytic activity of SrBi4O7 assessed by Yang and coworkers [201] was examined through the decomposition of MG in aqueous media (initial concentrations, 5–50 mg L−1) under visible light irradiation, subsequently homogenizing the suspension in the dark to ensure adsorption–desorption equilibrium. Optimal conditions appeared to be 5 mg L−1 of MG and 1.5 g L−1 of the SrBi4O7 bismuthate. Under such conditions, the irradiation of the SrBi4O7/MG aqueous suspension for 3 h caused about 98% degradation and 90% mineralization.
Table 1. Recent studies on nanostructured Bi-based photocatalysts for pollutant degradation.
Table 1. Recent studies on nanostructured Bi-based photocatalysts for pollutant degradation.
CatalystDosage
(g L−1)
PollutantDye Concentration (mg L−1)Light ConditionsDegradation Time and EfficiencyRef.
CNFs/g-C3N4/BiOBr3.0TC20300 W, Xe lamp, λ > 400 nm120 min/86.1%[182]
Bi2MoO6/CQDs/Bi2S30.3TC20300 W, Xe lamp, λ > 420 nm120 min/92.5%[183]
Bi2WO6/ZnIn2S40.2MO10300 W, Xe lamp, λ > 420 nm60 min/97.5%[184]
OVs-Bi2O3/Bi2SiO51.0MO-500 W, Xe lamp, λ > 420 nm7 h/71.8%[164]
BiVO4/Bi2S3-Cr(VI)-500 W, Xe lamp, 420 nm filters-[185]
BiVO4/Bi2O2CO3-RhB---[186]
Bi25FeO40/Bi2Fe4O90.1RhB10--[187]
Bi2WO6/BiOBr-RhB-300 W, Xe lamp, λ > 420 nm-[188]
Bi2WO6/ZnFe2O40.05RhB50-300 min[189]
BiFeO3/TiO20.5MB-300 W, Xe lamp120 min/96%[190]
BiFeO3/carbon0.01MB-300 W, Xe lamp54 min/54%[191]
BiFeO3-GdFeO30.01MB-Sunlight120 min/56%[192]
BiFeO3/Fe3O40.02MB-500 W, halogen lamp40 min/100%[56]
MOF-BiFeO30.02MB-300 W, Xe lamp120 min/93%[193]
CuO–BiVO40.01MB-150 W, Xe lamp150 min/100%[194]
Bi2MoO6–ZnSnO30.01MB-300 W, Xe lamp60 min/95%[195]
BiOBr/Bi2O30.01MB-300 W, Xe lamp50 min/87%[196]
BiFeO3/Bi2WO60.06MB-500 W, halogen lamp54 min/75%[197]

8. Concluding Remarks

In summary, this review article further emphasizes the development of and recent advances in bismuth-based photocatalysts and discusses various approaches to improve their photocatalytic performance and associated photocatalytic mechanisms by modifying the band energies and electronic structures of nanostructures or heterojunctions. Particular attention has been devoted to the application of BBNs in the degradation of organic pollutants (i.e., organic dyes and pesticides) and water treatment due to their photocatalytic activity for the formation of C-C and C-S bonds and atom-transfer radical-addition-type reactions. Hazardous dyes in wastewater treated by BBN photocatalysts include methylene blue, rhodamine B, 2,4,6-trichlorophenol, methyl orange, acid orange, acetaminophen, carcinogenic reactive black 5, carbamazepine, malachite green, benzene in aqueous solution, phenol, bisphenol A, and antibiotics. BiOX (X = Cl, Br, and I) nanoparticles and up-conversion phosphors/BiOBr composites are also efficient catalysts for the degradation of NOx gas. Overall, the fundamental study of the synthesis, characterization, adsorption, and photocatalytic applications of some popular bismuth oxide-based materials have been examined and may be helpful for the development of other metal oxide compounds toward dye degradation under solar illumination and for other environmental and energy-related applications. The introduction of crystal plane tailoring, the creation of porous or hollow structures, and the construction of nanostructured features can enhance the photoactivity of bismuth materials. Various strategies have been applied to further enhance their photoactivity, including double-rich approaches, interfacial engineering, metal doping and SPR effects, and the internal coupling of cocatalysts to host materials for various environmental applications. Although efforts have been made to tune bismuth-based materials and optimize the high performance of their photocatalytic activity, their potential has not been fully exploited. Thus, there are several fundamental insights into the formation of phases and their effects on photocatalytic processes. Current research on the interfacial engineering of bismuth-based materials shows that due to the rich physicochemical structural features of nanostructures that bulk materials do not possess, changing their electronic structures and band energies has a great impact on their photocatalytic efficiency. Therefore, more attention needs to be paid to controlling the preparation of bismuth-based photocatalysts using facile synthetic methods. Although efforts have been made to introduce defects and form heterojunctions to optimize photocatalytic activity, the defect types or doping types present in bismuth materials are almost always oxygen vacancies. These defects can induce additional electronic states and affect the electron transfer rate in nanostructures, implying that an increase in the density of states generally increases the photoactivity of the nanostructures. More efficient methods are needed to construct different types of defects in nanostructures and gain insight into the relationship between defect type and number and photoactivity. Therefore, it is highly desirable to achieve high-efficiency photoactivity through further interfacial engineering of bismuth-based materials.

Author Contributions

Writing—original draft preparation, S.V.P.V., J.Z., R.R., J.S. and C.M.J.; writing—review and editing, S.V.P.V., A.M. and C.M.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was sponsored by the NRF-Korea (2020R1A2B5B01002744, and 2020R1A2C1012439, 2020R1A4A1019227).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

1D One-dimensional
2,4-DCP2,4-dichlorophenol
4-CP4-chlorophenol
APAPAcetaminophen
BBNsBismuth-based nanostructures
BGBrilliant green
BPABisphenol A
CB Conduction band
CIPCiprofloxacin
CQDsCarbon quantum dots
CTABCetyltrimethylammonium bromide
CTCChlortetracycline
CVCrystal violet
ICIndigo carmine
LEDLight-emitting diode
MBMethylene blue
MOMethyl orange
NHENormal hydrogen electrode
NONitric oxide
OTCOxytetracycline
OVsOxygen vacancies
PTBP4-tert-butylphenol
rGOReduced graphene oxide
RhBRhodamine B
SPRSurface plasmon resonance
TCTetracycline
TCHCTetracycline hydrochloride
TEOSTetraethoxysilane
TOCTotal organic carbon
UVUltraviolet
VBValence band

References

  1. Li, K.; Li, B. Impacts of urbanization and industrialization on energy consumption/CO2 emissions: Does the level of development matter? Renew. Sustain. Energy Rev. 2015, 52, 1107–1122. [Google Scholar] [CrossRef]
  2. Andersson, J.; Grönkvist, S. Large-scale storage of hydrogen. Int. J. Hydrogen Energy 2019, 44, 11901–11919. [Google Scholar] [CrossRef]
  3. International Renewable Energy Agency (IRENA). Hydrogen: A Renewable Energy Perspective. Report Prepared for the 2nd Hydrogen Energy Ministerial Meeting in Tokyo, Japan (September 2019). Available online: https://www.irena.org/-/media/Files/IRENA/Agency/Publication/2019/Sep/IRENA_Hydrogen_2019.pdf (accessed on 10 September 2019).
  4. Colón, G. Towards the hydrogen production by photocatalysis. Appl. Catal. A Gen. 2016, 518, 48–59. [Google Scholar] [CrossRef]
  5. Reischauer, S.; Pieber, B. Emerging concepts in photocatalytic organic synthesis. iScience 2021, 24, 102209. [Google Scholar] [CrossRef]
  6. Sudha, D.; Sivakumar, P. review on the photocatalytic activity of various composite catalysts. Chem. Eng. Proc. 2015, 97, 112–133. [Google Scholar] [CrossRef]
  7. Xiao, M.; Wang, Z.; Lyu, M.; Luo, B.; Wang, S.; Liu, G.; Cheng, H.-M.; Wang, L. Hollow nanostructures for photocatalysis: Advantages and challenges. Adv. Mater. 2019, 31, 1801369. [Google Scholar] [CrossRef] [PubMed]
  8. Liu, J.; Ma, N.; Wu, W.; He, Q. Recent progress on photocatalytic heterostructures with full solar spectral responses. Chem. Eng. J. 2020, 393, 124719. [Google Scholar] [CrossRef]
  9. Pereira, A.L.J.; Errandonea, D.; Beltrán, A.; Gracia, L.; Gomis, O.; Sans, J.A.; Garcia-Domene, B.; Miquel-Veyrat, A.; Manjón, F.J.; Muñoz, A.; et al. Structural study of α-Bi2O3 under pressure. J. Phys. Condens. Matter 2013, 25, 475402. [Google Scholar] [CrossRef] [Green Version]
  10. Zhou, L.; Wang, W.; Xu, H.; Sun, S.; Shang, M. Bi₂O3 hierarchical nanostructures: Controllable synthesis, growth mechanism, and their application in photocatalysis. Chem. Eur. J. 2009, 15, 1776–1782. [Google Scholar] [CrossRef]
  11. Ke, J.; Liu, J.; Sun, H.; Zhang, H.; Duan, X.; Liang, P.; Li, X.; Tade, M.O.; Liu, S.; Wang, S. Facile assembly of Bi2O3/Bi2S3/MoS2 n-p heterojunction with layered n-Bi2O3 and p-MoS2 for enhanced photocatalytic water oxidation and pollutant degradation. Appl. Catal. B Environ. 2017, 200, 47–55. [Google Scholar] [CrossRef]
  12. Jiang, L.; Yuan, X.; Zeng, G.; Liang, J.; Chen, X.; Yu, H.; Wang, H.; Wu, Z.; Zhang, J.; Xionga, T. In-situ synthesis of direct solid-state dual Z-scheme WO3/g-C3N4/Bi2O3 photocatalyst for the degradation of refractory pollutant. Appl. Catal. B Environ. 2018, 227, 376–385. [Google Scholar] [CrossRef]
  13. Riente, P.; Foianchini, M.; Llanrs, P.; Pericas, M.A.; Noël, T. Shedding light on the nature of the catalytically active species in photocatalytic reaction using Bi2O3 semiconductor. Nat. Commun. 2021, 12, 625. [Google Scholar] [CrossRef] [PubMed]
  14. Khairnar, S.D.; Kulkarni, A.N.; Shinde, S.G.; Marathe, S.D.; Marathe, Y.V.; Dhole, S.D.; Shrivastava, V.S. Synthesis and characterization of 2-D La-doped Bi2O3 for photocatalytic degradation of organic dye and pesticide. J. Photochem. Photobiol. 2021, 6, 100030. [Google Scholar] [CrossRef]
  15. Najafian, H.; Manteghi, F.; Beshkar, F.; Niasari, M.S. Enhanced photocatalytic activity of a novel NiO/Bi2O3/Bi3ClO4 nanocomposite for the degradation of azo dye pollutants under visible light irradiation. Separ. Purif. Technol. 2019, 209, 6–17. [Google Scholar] [CrossRef]
  16. Muruganandham, M.; Amutha, R.; Lee, G.-J.; Hsieh, S.-H.; Wu, J.J.; Sillanpää, M. Facile fabrication of tunable Bi2O3 self-assembly and its visible light photocatalytic activity. J. Phys. Chem. C 2012, 116, 12906–12915. [Google Scholar] [CrossRef]
  17. Riente, P.; Matas, A.A.; Albero, J.; Palomares, E.; Pericàs, M.A. Light-driven organocatalysis using inexpensive, nontoxic Bi2O3 as the photocatalyst. Angew. Chem. Int. Ed. 2014, 53, 9613–9616. [Google Scholar] [CrossRef]
  18. Fadeyi, O.O.; Mousseau, J.J.; Feng, Y.; Allais, C.; Nuhant, P.; Chen, M.Z.; Pierce, B.; Robinson, R. Visible-light-driven photocatalytic initiation of radical thiol-ene reactions using bismuth oxide. Org. Lett. 2015, 17, 5756–5759. [Google Scholar] [CrossRef]
  19. Hakobyan, K.; Gegenhuber, T.; McErlean, C.S.P.; Müllner, M. Visible-light driven MADIX polymerisation via a reusable, low-cost, and non-toxic bismuth oxide photocatalyst. Angew. Chem. Int. Ed. 2019, 58, 1828–1832. [Google Scholar] [CrossRef]
  20. Nicewicz, D.A.; MacMillan, D.W.C. Merging photoredox catalysis with organocatalysis: The direct asymmetric alkylation of aldehydes. Science 2008, 322, 77–80. [Google Scholar] [CrossRef] [Green Version]
  21. Dong, F.; Xiong, T.; Sun, Y.; Zhao, Z.; Zhou, Y.; Feng, X.; Wu, Z. A semimetal bismuth element as a direct plasmonic photocatalyst. Chem. Commun. 2014, 50, 10386–10389. [Google Scholar] [CrossRef]
  22. Sun, Y.; Zhao, Z.; Zhang, W.; Gao, C.; Zhang, Y.; Dong, F. Plasmonic Bi metal as cocatalyst and photocatalyst: The case of Bi/(BiO)2CO3 and Bi particles. J. Coll. Interface Sci. 2017, 485, 1–10. [Google Scholar] [CrossRef] [PubMed]
  23. Boutinaud, P. Revisiting the spectroscopy of the Bi3+ ion in oxide compounds. Inorg. Chem. 2013, 52, 6028–6038. [Google Scholar] [CrossRef] [PubMed]
  24. Bainglass, E.; Walsh, A.; Huda, M.N. BiSbWO6: Properties of a mixed 5s/6s lone-pair-electron system. Chem. Phys. 2021, 544, 111117. [Google Scholar] [CrossRef]
  25. Liu, Y.; Zhu, G.; Gao, J.; Zhu, R.; Hojamberdiev, M.; Wang, C.; Wei, X.; Liu, P. A novel synergy of Er3+/Fe3+ co-doped porous Bi5O7I microspheres with enhanced photocatalytic activity under visible-light irradiation. Appl. Catal. B Environ. 2017, 205, 421–432. [Google Scholar] [CrossRef]
  26. Ganeshbabu, M.; Kannan, N.; Sundara Venkatesh, P.; Paulraj, G.; Jeganathan, K.; Mubarak Ali, D. Synthesis and characterization of BiVO4 nanoparticles for environmental applications. RSC Adv. 2020, 10, 18315–18322. [Google Scholar] [CrossRef] [PubMed]
  27. Huang, C.; Chen, L.; Li, H.; Mu, Y.; Yang, Z. Synthesis and application of Bi2WO6 for the photocatalytic degradation of two typical fluoroquinolones under visible light irradiation. RSC Adv. 2019, 9, 27768–27779. [Google Scholar] [CrossRef] [Green Version]
  28. Kumar, R.; Raizada, P.; Parwaz Khan, A.A.; Nguyen, V.-H.; Le, Q.V.; Ghotekar, S.; Selvasembian, R.; Gandhi, V.; Singh, A.; Singh, P. Recent progress in emerging BiPO4-based photocatalysts: Synthesis, properties, modification strategies, and photocatalytic applications. J. Mater. Sci. Technol. 2022, 108, 208–225. [Google Scholar] [CrossRef]
  29. Zhao, G.; Zheng, Y.; He, Z.; Lu, Z.; Wang, L.; Li, C.; Jiao, F.; Deng, C. Synthesis of Bi2S3 microsphere and its efficient photocatalytic activity under visible-light irradiation. Trans. Nonferrous Met. Soc. China 2018, 28, 2002–2010. [Google Scholar] [CrossRef]
  30. Arumugam, M.; Choi, M.Y. Recent progress on bismuth oxyiodide (BiOI) photocatalyst for environmental remediation. J. Ind. Eng. Chem. 2020, 81, 237–268. [Google Scholar] [CrossRef]
  31. Imam, S.S.; Adnan, R.; Mohd Kaus, N.H. The photocatalytic potential of BiOBr for wastewater treatment: A mini-review. J. Environ. Chem. Eng. 2021, 9, 105404. [Google Scholar] [CrossRef]
  32. Wang, Q.; Hui, J.; Huang, Y.; Ding, Y.; Cai, Y.; Yin, S.; Li, Z.; Su, B. The preparation of BiOCl photocatalyst and its performance of photodegradation on dyes. Mater. Sci. Semicond. Process. 2014, 17, 87–93. [Google Scholar] [CrossRef]
  33. Ai, L.H.; Zeng, Y.; Jiang, J. Hierarchical porous BiOI architectures: Facile microwave nonaqueous synthesis, characterization and application in the removal of Congo red from aqueous solution. Chem. Eng. J. 2014, 235, 331–339. [Google Scholar] [CrossRef]
  34. He, R.; Xu, D.; Cheng, B.; Yu, J.; Ho, W. Review on nanoscale Bi-based photocatalysts. Nanoscale Horiz. 2018, 3, 464–504. [Google Scholar] [CrossRef] [PubMed]
  35. Song, S.; Xing, Z.; Zhao, H.; Li, Z.; Zhou. Recent advances in bismuth-based photocatalysts: Environment and energy applications. Green Energy Environ. 2022, in press. [Google Scholar] [CrossRef]
  36. Lu, Z.; Zhang, X.; Li, G.; Cao, Y.; Shao, Y.; Li, D. Highly efficient Bi2O2CO3/BiOCl photocatalyst based on heterojunction with enhanced dye-sensitization under visible light. Appl. Catal. B Environ. 2016, 187, 301–309. [Google Scholar]
  37. Qiu, F.; Li, W.; Wang, F.; Li, H.; Liu, X.; Ren, C. Preparation of novel p-n heterojunction Bi2O2CO3/BiOBr photocatalysts with enhanced visible light photocatalytic activity. Coll. Surf. A 2017, 517, 25–32. [Google Scholar] [CrossRef]
  38. Huang, Y.; Fan, W.; Long, B.; Li, H.; Zhao, F.; Liu, Z.; Tong, Y.; Ji, H. Visible light Bi2S3/Bi2O3/Bi2O2CO3 photocatalyst for effective degradation of organic pollutions. Appl. Catal. B Environ. 2016, 185, 68–76. [Google Scholar] [CrossRef]
  39. Ding, J.; Dai, Z.; Qin, F.; Zhao, H.; Zhao, S.; Chen, R. Z-scheme BiO1-xBr/Bi2O2CO3 photocatalyst with rich oxygen vacancy as electron mediator for highly efficient degradation of antibiotics. Appl. Catal. B Environ. 2017, 205, 281–291. [Google Scholar] [CrossRef]
  40. Peng, D.Y.; Zeng, H.Y.; Xiong, J.; Liu, F.-Y.; Wang, L.-H.; Xu, S.; Yang, Z.-L.; Liu, S.-G. Tuning oxygen vacancy in Bi2WO6 by heteroatom doping for enhanced photooxidation-reduction properties. J. Colloid Interface Sci. Part B 2023, 629, 133–146. [Google Scholar] [CrossRef]
  41. Ding, J.; Dai, Z.; Tian, F.; Zhou, B.; Zhao, B.; Zhao, H.P.; Chen, Z.Q.; Liu, Y.L.; Chen, R. Generation of VBrV’’’BiV••O defect clusters for 1O2 production for molecular oxygen activation in photocatalysis. J. Mater. Chem. A 2017, 5, 23909–23918. [Google Scholar] [CrossRef]
  42. Zhao, S.; Dai, Z.; Guo, W.; Chen, F.; Liu, Y.; Chen, R. Highly selective oxidation of glycerol over Bi/Bi3.64Mo0.36O6.55 heterostructure: Dual reaction pathways induced by photogenerated 1O2 and holes. Appl. Catal. B Environ. 2019, 244, 206–214. [Google Scholar] [CrossRef]
  43. Bai, S.; Li, X.Y.; Kong, Q.; Long, R.; Wang, C.M.; Jiang, J.; Xiong, Y.J. Toward enhanced photocatalytic oxygen evolution: Synergetic utilization of Plasmonic effect and Schottky junction via interfacing facet selection. Adv. Mater. 2015, 27, 3444–3452. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, S.; Hai, X.; Ding, X.; Chang, K.; Xiang, Y.; Meng, X.; Yang, Z.; Chen, H.; Ye, J. Light-switchable oxygen vacancies in ultrafine Bi5O7Br nanotubes for boosting solar-driven nitrogen fixation in pure water. Adv. Mater. 2017, 29, 1701774. [Google Scholar] [CrossRef] [PubMed]
  45. Qin, K.; Zhao, Q.; Yu, H.; Xia, X.; Li, J.; He, S.; Wei, L.; An, T. A review of bismuth-based photocatalysts for antibiotic degradation: Insight into the photocatalytic degradation performance, pathways and relevant mechanisms. Environ. Res. 2021, 199, 111360. [Google Scholar] [CrossRef]
  46. Asghar, A.; Raman, A.A.A.; Wan Daud, W.M.A. Advanced oxidation processes for in-situ production of hydrogen peroxide/hydroxyl radical for textile wastewater treatment: A review. J. Clean. Prod. 2015, 87, 826–838. [Google Scholar] [CrossRef] [Green Version]
  47. Subhiksha, V.; Kokilavani, S.; Sudheer Khan, S. Recent advances in degradation of organic pollutant in aqueous solutions using bismuth based photocatalysts: A review. Chemosphere 2022, 290, 133228. [Google Scholar] [CrossRef]
  48. Huang, C.-H.; Wu, T.; Huang, C.-W.; Lai, C.-Y.; Wu, M.-Y.; Lin, Y.-W. Enhanced photocatalytic performance of BiVO4 in aqueous AgNO3 solution under visible light irradiation. Appl. Surf. Sci. 2017, 399, 10–19. [Google Scholar] [CrossRef]
  49. Wang, S.; Huang, H.; Zhang, Y. A novel layered bismuth-based photocatalytic material LiBi3O4Cl2 with –OH and h+ as the active species for efficient photodegradation applications. Solid State Sci. 2016, 62, 43–49. [Google Scholar] [CrossRef]
  50. He, S.; Shen, M.; Wu, E.; Yin, R.; Zhu, M.; Zeng, L. Molecular structure on the detoxification of fluorinated liquid crystal monomers with reactive oxidation species in the photocatalytic process. Environ. Sci. Ecothechnol. 2022, 9, 100141. [Google Scholar] [CrossRef]
  51. Low, J.; Yu, J.; Jaroniec, M.; Wageh, S.; Al-Ghamdi, A.A. Heterojunction photocatalysts. Adv. Mater. 2017, 29, 1601694. [Google Scholar] [CrossRef]
  52. Hong, Y.; Li, C.; Meng, Y.; Huang, C.; Shi, W. In situ synthesis of a nanoplate-like Bi-based heterojunction for photocatalytic degradation of ciprofloxacin. Mater. Sci. Eng. B 2017, 224, 69–77. [Google Scholar] [CrossRef]
  53. Shafiq, F.; Tahir, M.B.; Hussain, A.; Sagir, M.; Rehman, J.; Kebaili, I.; Alrobei, H.; Alzaid, M. The construction of a highly efficient p-n heterojunction Bi2O3/BiVO4 for hydrogen evolution through solar water splitting. Int. J. Hydrogen Energy 2022, 47, 4594–4600. [Google Scholar] [CrossRef]
  54. Stelo, F.; Kublik, N.; Ullah, S.; Wender, H. Recent advances in Bi2MoO6 based Z-scheme heterojunctions for photocatalytic degradation of pollutants. J. Alloys Compd. 2020, 829, 154591. [Google Scholar] [CrossRef]
  55. Shan, L.; Wang, G.; Liu, L.; Wu, Z. Band alignment and enhanced photocatalytic activation for α-Bi2O3/BiOCl (001) core–shell heterojunction. J. Mol. Catal. A Chem. 2015, 406, 145–151. [Google Scholar] [CrossRef]
  56. Volnistem, E.A.; Bini, R.D.; Silva, D.M.; Rosso, J.M.; Dias, G.S.; Cótica, L.F.; Santos, I.A. Intensifying the photocatalytic degradation of methylene blue by the formation of BiFeO3/Fe3O4 nanointerfaces. Ceram. Int. 2020, 46, 18768–18777. [Google Scholar] [CrossRef]
  57. Liu, T.; Li, Y.; Cheng, Z.; Peng, Y.; Shen, M.; Yang, S.; Zhang, Y. Crystal-face-selective Bi4Ti3O12/BiOI heterojunctions constructed for enhanced visible-light-driven photocatalytic activity. Appl. Surf. Sci. 2021, 552, 149507. [Google Scholar] [CrossRef]
  58. Bera, S.; Ghosh, S.; Maiyalagan, T.; Basu, R.N. Band edge engineering of BiOX/CuFe2O4 heterostructures for efficient water splitting. ACS Appl. Energy Mater. 2022, 5, 3821–3833. [Google Scholar] [CrossRef]
  59. Christopher, P.; Xin, H.L.; Marimuthu, A.; Linic, S. Singular characteristics and unique chemical bond activation mechanisms of photocatalytic reactions on plasmonic nanostructures. Nat. Mater. 2012, 11, 1044–1050. [Google Scholar] [CrossRef]
  60. Dong, X.A.; Zhang, W.D.; Sun, Y.J.; Li, J.Y.; Cen, W.L.; Cui, Z.H.; Huang, H.W.; Dong, F. Visible-light-induced charge transfer pathway and photocatalysis mechanism on Bi semimetal@defective BiOBr hierarchical microspheres. J. Catal. 2018, 357, 41–50. [Google Scholar] [CrossRef]
  61. He, W.J.; Sun, Y.J.; Jiang, G.M.; Li, Y.H.; Zhang, X.M.; Zhang, Y.X.; Zhou, Y.; Dong, F. Defective Bi4MoO9/Bi metal core/shell heterostructure: Enhanced visible light photocatalysis and reaction mechanism. Appl. Catal. B 2018, 239, 619–627. [Google Scholar] [CrossRef]
  62. Li, J.R.; Zhang, W.D.; Ran, M.X.; Sun, Y.J.; Huang, H.W.; Dong, F. Synergistic integration of Bi metal and phosphate defects on hexagonal and monoclinic BiPO4: Enhanced photocatalysis and reaction mechanism. Appl. Catal. B 2019, 243, 313–321. [Google Scholar] [CrossRef]
  63. Zhang, L.; Yang, C.; Lv, K.; Lu, Y.; Li, Q.; Wu, X.; Li, Y.; Li, X.; Fan, J.; Li, M. SPR effect of bismuth enhanced visible photoreactivity of Bi2WO6 for NO abatement. Chin. J. Catal. 2019, 40, 755–764. [Google Scholar] [CrossRef]
  64. Han, Q. Advances in preparation methods of bismuth-based photocatalysts. Chem. Eng. J. 2021, 414, 127877. [Google Scholar] [CrossRef]
  65. Wu, L.; Li, Z.; Li, Y.; Hu, H.; Liu, Y.; Zhang, Q. Mechanochemical syntheses of bismuth oxybromides BixOyBrz as visible-light responsive photocatalyts for the degradation of bisphenol A. J. Solid State Chem. 2019, 270, 458–462. [Google Scholar] [CrossRef]
  66. Maleki, H. Characterization and photocatalytic activity of Y-doped BiFeO3 ceramics prepared by solid-state reaction method. Adv. Powder Technol. 2019, 30, 2832–2840. [Google Scholar] [CrossRef]
  67. Wu, X.; Zhang, K.; Zhang, G.; Yin, S. Facile preparation of BiOX (X = Cl, Br, I) nanoparticles and up-conversion phosphors/BiOBr composites for efficient degradation of NO gas: Oxygen vacancy effect and near infrared light responsive mechanism. Chem. Eng. J. 2017, 325, 59–70. [Google Scholar] [CrossRef]
  68. Siao, C.-W.; Chen, H.-L.; Chen, L.-W.; Chang, J.-L.; Yeh, T.-W.; Chen, C.-C. Controlled hydrothermal synthesis of bismuth oxychloride/bismuth oxybromide/bismuth oxyiodide composites exhibiting visible-light photocatalytic degradation of 2-hydroxybenzoic acid and crystal violet. J. Colloid Inter. Sci. 2018, 526, 322–336. [Google Scholar] [CrossRef]
  69. Cai, J.; Huang, J.; Lai, Y. 3D Au-decorated BiMoO6 nanosheet/TiO2 nanotube array heterostructure with enhanced UV and visible-light photocatalytic activity. J. Mater. Chem. A 2017, 5, 16412–16421. [Google Scholar] [CrossRef]
  70. Lin, H.; Ye, H.; Li, X.; Cao, J.; Chen, S. Facile anion-exchange synthesis of BiOI/BiOBr composite with enhanced photoelectrochemical and photocatalytic properties. Ceram. Int. 2014, 40, 9743–9750. [Google Scholar] [CrossRef]
  71. Xiao, X.; Hao, R.; Zuo, X.; Nan, J.; Li, L.; Zhang, W. Microwave-assisted synthesis of hierarchical Bi7O9I3 microsheets for efficient photocatalytic degradation of bisphenol-A under visible light irradiation. Chem. Eng. J. 2012, 209, 293–300. [Google Scholar] [CrossRef]
  72. Liu, Z.; Xu, X.; Fang, J.; Zhu, X.; Chu, J.; Li, B. Microemulsion synthesis, characterization of bismuth oxyiodine/titanium dioxide hybrid nanoparticles with outstanding photocatalytic performance under visible light irradiation. Appl. Surf. Sci. 2012, 258, 3771–3778. [Google Scholar] [CrossRef]
  73. Lu, Y.; Zhao, Y.; Zhaon, J.; Song, Y.; Huang, Z.; Gao, F.; Li, N.; Li, Y. Photoactive β-Bi2O3 architectures prepared by a simple solution crystallization method. Ceram. Int. 2014, 40, 15057–15063. [Google Scholar] [CrossRef]
  74. Qin, F.; Li, G.; Wang, R.; Wu, J.; Sun, H.; Chen, R. Template-free fabrication of Bi2O3 and (BiO)2CO3 nanotubes and their application in water treatment. Chem. Eur. J. 2012, 18, 16491–16497. [Google Scholar] [CrossRef]
  75. Hou, J.; Yang, C.; Wang, Z.; Zhou, W.; Jiao, S.; Zhu, H. In situ synthesis of α−β phase heterojunction on Bi2O3 nanowires with exceptional visible-light photocatalytic performance. Appl. Catal. B Environ. 2013, 142–143, 504–511. [Google Scholar] [CrossRef] [Green Version]
  76. Liang, Y.-C.; Li, T.-H. Controllable morphology of Bi2S3 nanostructures formed via hydrothermal vulcanization of Bi2O3 thin-film layer and their photoelectrocatalytic performances. Nanotechnol. Rev. 2022, 11, 284–297. [Google Scholar] [CrossRef]
  77. Sood, S.; Umar, A.; Mehta, S.K.; Kansal, S. α−Bi2O3 nanorods: An efficient sunlight active photocatalyst for degradation of Rhodamine B and 2,4,6-trichlorophenol. Ceram. Int. 2015, 41, 3355–3364. [Google Scholar] [CrossRef]
  78. Tien, L.-C.; Lai, Y.-C. Nucleation control and growth mechanism of pure α-Bi2O3 nanowires. Appl. Surf. Sci. 2014, 290, 131–136. [Google Scholar] [CrossRef]
  79. Xiao, X.; Hu, R.; Liu, C.; Xing, C.; Qian, C.; Zuo, X.; Nan, J.; Wang, L. Facile large-scale synthesis of β-Bi2O3 nanospheres as a highly efficient photocatalyst for the degradation of acetaminophen under visible light irradiation. Appl. Catal. B Environ. 2013, 140–141, 433–443. [Google Scholar] [CrossRef]
  80. Verma, M.K.; Kumar, A.; Das, T.; Kumar, V.; Singh, S.; Rai, V.S.; Prajapati, D.; Sonwani, R.K.; Sahoo, K.; Mandal, K.D. BiFeO3 perovskite as an efficient photocatalyst synthesized by soft chemical route. Mater. Technol. 2021, 36, 594–602. [Google Scholar] [CrossRef]
  81. Soltani, T.; Lee, B.-K. Novel and facile synthesis of Ba-doped BiFeO3 nanoparticles and enhancement of their magnetic and photocatalytic activities for complete degradation of benzene in aqueous solution. J. Hazard. Mater. 2016, 316, 122–133. [Google Scholar] [CrossRef]
  82. Ponraj, C.; Vinitha, G.; Daniel, J. A review on the visible light active BiFeO3 nanostructures as suitable photocatalyst in the degradation of different textile dyes. Environ. Nanotechnol. Monitor. Manag. 2017, 7, 110–120. [Google Scholar] [CrossRef]
  83. Skiker, R.; Zouraibi, M.; Saidi, M.; Ziat, K. Facile co-precipitation synthesis of novel Bi12TiO20/BiFeO3 heterostructure series with enhanced photocatalytic activity for removal of methyl orange from water. J. Phys. Chem. Solids 2018, 119, 265–275. [Google Scholar] [CrossRef]
  84. Duan, Q.; Kong, F.; Han, X.; Jiang, Y.; Liu, T.; Chang, Y.; Zhou, L.; Qin, G.; Zhang, X. Synthesis and characterization of morphology-controllable BiFeO3 particles with efficient photocatalytic activity. Mater. Res. Bull. 2019, 112, 104–108. [Google Scholar] [CrossRef]
  85. Li, Y.; Zhang, X.; Chen, L.; Sun, H.; Zhang, H.; Si, W.; Wang, W.; Wang, L.; Li, J. Enhanced magnetic and photocatalytic properties of BiFeO3 nanotubes with ultrathin wall thickness. Vacuum 2021, 184, 109867. [Google Scholar] [CrossRef]
  86. Wu, T.; Liu, L.; Pi, M.; Zhang, D.; Chen, S. Enhanced magnetic and photocatalytic properties of Bi2Fe4O9 semiconductor with large exposed (001) surface. Appl. Surf. Sci. 2016, 377, 253–261. [Google Scholar] [CrossRef]
  87. Wang, G.; Liu, S.; He, T.; Liu, X.; Deng, Q.; Mao, Y.; Wang, S. Enhanced visible-light-driven photocatalytic activities of Bi2Fe4O9/g-C3N4 composite photocatalysts. Mater. Res. Bull. 2018, 104, 104–111. [Google Scholar] [CrossRef]
  88. Ma, M.; Chen, Y.; Liu, Y.; Jiang, J.; Jiao, Z.; Ma, Y. Highly efficient photocatalytic organic dyes degradation based on 1D magnetic Bi2Fe4O9/C@AgBr composite. Appl. Organomet. Chem. 2022, 36, e6619. [Google Scholar] [CrossRef]
  89. Cheng, H.; Huang, B.; Lu, J.; Wang, Z.; Xu, B.; Qin, X.; Zhang, X.; Dai, Y. Synergistic effect of crystal and electronic structures on the visible-light-driven photocatalytic performances of Bi2O3 polymorphs. Phys. Chem. Chem. Phys. 2010, 12, 15468–15475. [Google Scholar] [CrossRef]
  90. Bano, K.; Mittal, S.K.; Singh, P.P.; Kaushal, S. Sunlight driven photocatalytic degradation of organic pollutants using a MnV2O6/BiVO4 heterojunction: Mechanistic perception and degradation pathways. Nanoscale Adv. 2021, 3, 6446–6458. [Google Scholar] [CrossRef]
  91. Guan, X.; Tian, L.; Zhang, Y.; Shi, J.; Shen, S. Photocatalytic water splitting on BiVO4: Balanced charge-carrier consumption and selective redox reaction. Nano Res. 2022. [Google Scholar] [CrossRef]
  92. Tian, S.; Bello, I.A.; Peng, J.; Ren, H.; Ma, Q.; Zhu, M.; Li, K.; Wang, H.; Wu, J.; Zhu, G. Fabricating Mn3O4/β-Bi2O3 heterojunction microspheres with enhanced photocatalytic activity for organic pollutants degradation and NO removal. J. Alloys Compd. 2021, 854, 157223. [Google Scholar] [CrossRef]
  93. He, G.; Xing, C.; Xiao, X.; Hu, R.; Zuo, X.; Nan, J. Facile synthesis of flower-like Bi12O17Cl2/β-Bi2O3 composites with enhanced visible light photocatalytic performance for the degradation of 4-tert-butylphenol. Appl. Catal. B Environ. 2015, 170–171, 1–9. [Google Scholar] [CrossRef]
  94. Sun, Y.; Wang, W.; Zhang, L.; Zhang, Z. Design and controllable synthesis of α-/γ-Bi2O3 homojunctions have synergistic effects on photocatalytic activity. J. Chem. Eng. 2012, 211–212, 161–167. [Google Scholar] [CrossRef]
  95. Gadhi, T.A.; Hernández-Gordillo, A.; Bizarro, M.; Jagdale, P.; Tagliaferro, A.; Rodil, S.E. Efficient α/β-Bi2O3 composite for the sequential photodegradation of two-dyes mixture. Ceram. Int. 2016, 42, 13065–13073. [Google Scholar] [CrossRef]
  96. Liu, X.; Liu, Y.; Liu, T.; Jia, Y.; Deng, H.; Wang, W.; Zhang, F. Alkali-mediated dissolution-recrystallization strategy for in situ construction of a BiVO4/Bi25VO40 heterojunction with promoted interfacial charge transfer: Formation mechanism and photocatalytic tetracycline degradation studies. Chem. Eng. J. 2022, 431, 134181. [Google Scholar] [CrossRef]
  97. Duan, A.; Hou, X.; Yang, M.; Yu, H.; Yang, Y.; Ma, Q.; Dong, X. EDTA-2Na assisted facile synthesis of monoclinic bismuth vanadate (m-BiVO4) and m-BiVO4/rGO as a highly efficient visible-light-driven photocatalyst. Mater. Lett. 2022, 311, 131498. [Google Scholar] [CrossRef]
  98. El-Hakam, S.A.; Al-Shorifi, F.T.; Salama, R.S.; Gamal, S.; El-Yazeed, W.S.A.; Ibrahim, A.A.; Ahmed, A.I. Application of nanostructured mesoporous silica/ bismuth vanadate composite catalysts for the degradation of methylene blue and brilliant green. J. Mater. Res. Technol. 2022, 18, 1963–1976. [Google Scholar] [CrossRef]
  99. Wu, W.; Sun, Y.; Zhou, H. In-situ construction of β-Bi2O3/Ag2O photocatalyst from deactivated AgBiO3 for tetracycline degradation under visible light. Chem. Eng. J. 2022, 432, 134316. [Google Scholar] [CrossRef]
  100. Wu, W.; Xu, C.; Shi, X.; Zhao, J.; An, X.; Ma, H.; Tian, Y.; Zhou, H. Effective degradation of organic pollutants and reaction mechanism with flower-like AgBiO3/g-C3N4 composite. Colloids Surf. A 2021, 599, 124901. [Google Scholar] [CrossRef]
  101. Bautista-Ruiz, J.; Chaparro, A.; Bautista, W. Characterization of bismuth-silicate soles. J. Phys. Conf. Ser. 2019, 1386, 012020. [Google Scholar] [CrossRef]
  102. Pirovano, C.; Islam, M.S.; Vannier, R.-N.; Nowogrocki, G.; Mairesse, G. Modelling the crystal structures of Aurivillius phases. Solid State Ion. 2001, 140, 115–123. [Google Scholar] [CrossRef]
  103. Kim, Y.; Kim, J.; Fujiwara, A.; Taniguchi, H.; Kim, S.; Tanaka, H.; Sugimoto, K.; Kato, K.; Itoh, M.; Hosonog, H.; et al. Hierarchical dielectric order in hierarchical ferroelectric Bi2SiO5. IUCrJ 2014, 1, 160–164. [Google Scholar] [CrossRef] [PubMed]
  104. Taniguchi, H.; Kuwabara, A.; Kim, J.; Kim, Y.; Moriwake, H.; Kim, S.; Hoshiyama, T.; Koyama, T.; Mori, S.; Takata, M.; et al. Ferroelectricity driven by twisting of silicate tetrahedral chains. Angew. Chem. Int. Ed. 2013, 52, 8088–8092. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Wan, Z.; Zhang, G. Synthesis and facet-dependent enhanced photocatalytic activity of Bi2SiO5/AgI nanoplate photocatalysts. J. Mater. Chem. A 2015, 3, 16737–16745. [Google Scholar] [CrossRef]
  106. Liu, D.; Cai, W.; Wang, Y.; Zhu, Y. Constructing a novel Bi2SiO5/BiPO4 heterostructure with extended light response range and enhanced photocatalytic performance. Appl. Catal. B Environ. 2018, 236, 205–211. [Google Scholar] [CrossRef]
  107. Zou, C.; Liang, M.; Yang, Z.; Zhou, X.; Yang, Y.; Yang, S. Flower-like Bi2SiO5/Bi4MoO9 heterostructures for enhanced photocatalytic degradation of ciprofloxacin. Nanotechnology 2020, 31, 345604. [Google Scholar] [CrossRef]
  108. Veber, A.; Kunej, S.; Korosec, R.C.; Suvorov, D. The effects of solvents on the formation of sol–gel-derived Bi12SiO20 thin films. J. Eur. Ceram. Soc. 2010, 30, 2475–2480. [Google Scholar] [CrossRef]
  109. Di, J.; Xia, J.; Huang, Y.; Ji, M.; Fan, W.; Chen, Z.; Li, H. Constructing carbon quantum dots/Bi2SiO5 ultrathin nanosheets with enhanced photocatalytic activity and mechanism investigation. Chem. Eng. J. 2016, 302, 334–343. [Google Scholar] [CrossRef]
  110. Yang, C.-T.; Lee, W.W.; Lin, H.-P.; Dai, Y.-M.; Chia, H.-T.; Chen, C.-C. A novel heterojunction photocatalyst, Bi2SiO5/g-C3N4: Synthesis, characterization, photocatalytic activity, and mechanism. RSC Adv. 2016, 6, 40664–40675. [Google Scholar] [CrossRef]
  111. Wu, Y.; Lu, J.; Li, M.; Yuan, J.; Wu, P.; Chang, X.; Liu, C.; Wang, X. Bismuth silicate photocatalysts with enhanced light harvesting efficiency by photonic crystal. J. Alloys Compd. 2019, 810, 151839. [Google Scholar] [CrossRef]
  112. Jia, K.-L.; Qu, J.; Hao, S.-M.; An, F.; Jing, Y.-Q.; Yu, Z.-Z. One-pot synthesis of bismuth silicate heterostructures with tunable morphology and excellent visible light photodegradation performances. J. Colloid Interface Sci. 2017, 506, 255–262. [Google Scholar] [CrossRef] [PubMed]
  113. Liu, D.; Yao, W.; Wang, J.; Liu, Y.; Zhang, M.; Zhu, Y. Enhanced visible light photocatalytic performance of a novel heterostructured Bi4O5Br2/Bi24O31Br10/Bi2SiO5 photocatalyst. Appl. Catal. B Environ. 2015, 172–173, 100–107. [Google Scholar] [CrossRef]
  114. Wu, Z.; Shen, J.; Ma, N.; Li, Z.; Wu, M.; Xu, D.; Zhang, S.; Feng, W.; Zhu, Y. Bi4O5Br2 nanosheets with vertical aligned facets for efficient visible-light-driven photodegradation of BPA. Appl. Catal. B Environ. 2021, 286, 119937. [Google Scholar] [CrossRef]
  115. Zhang, H.; Feng, Y.; Jia, S.; Jiang, D.; Zhan, Q. Enhancing the photocatalytic performance of Bi12SiO20 by in situ grown Bi2O2CO3 and Bi through two-step light irradiation method. Appl. Surf. Sci. 2020, 520, 146355. [Google Scholar] [CrossRef]
  116. Wang, Z.; Chen, M.; Huang, D.; Zeng, G.; Xu, P.; Zhou, C.; Lai, C.; Wang, H.; Cheng, M.; Wang, W. Multiply structural optimized strategies for bismuth oxyhalide photocatalysis and their environmental application. Chem. Eng. J. 2019, 374, 1025–1045. [Google Scholar] [CrossRef]
  117. Wu, P.; Feng, L.; Liang, Y.; Zhang, X.; Li, X.; Tian, S.; Hu, H.; Yin, G.; Khan, S. Large-scale synthesis of 2D bismuth-enriched bismuth oxyiodides at low temperatures for high-performance supercapacitor and photocatalytic applications. J. Mater. Sci. Mater. Electron. 2020, 31, 5385–5401. [Google Scholar] [CrossRef]
  118. Xiao, X.; Liu, C.; Hu, R.; Zuo, X.; Nan, J.; Li, L.; Wang, L. Oxygen-rich bismuth oxyhalides: Generalized one-pot synthesis, band structures and visible-light photocatalytic properties. J. Mater. Chem. 2012, 22, 22840–22843. [Google Scholar] [CrossRef]
  119. Xiao, X.; Xing, C.; He, G.; Zuo, X.; Nan, J.; Wang, L. Solvothermal synthesis of novel hierarchical Bi4O5I2 nanoflakes with highly visible light photocatalytic performance for the degradation of 4-tert-butylphenol. Appl. Catal. B Environ. 2014, 148–149, 154–163. [Google Scholar] [CrossRef]
  120. Peng, H.; Wang, L.; Xu, J.; Jiang, S.; Xu, X.; Zhang, Q. Novel Bi4O5I2/Bi4O5I2 hybrids: Synthesis, characterization, performance test. Mater. Lett. 2019, 256, 126694. [Google Scholar] [CrossRef]
  121. Ding, C.; Ye, L.; Zhao, Q.; Zhong, Z.; Liu, K.; Xie, H.; Bao, K.; Zhang, X.; Huang, Z. Synthesis of BixOyIz from molecular precursor and selective photoreduction of CO2 into CO. J. CO2 Utiliz. 2016, 14, 135–142. [Google Scholar] [CrossRef]
  122. Yin, R.; Li, Y.; Zhong, K.; Yao, H.; Zhang, Y.; Lai, K. Multifunctional property exploration: Bi4O5I2 with visible light photocatalytic performance and a large non-linear optical effect. RSC Adv. 2019, 9, 4539–4544. [Google Scholar] [CrossRef] [PubMed]
  123. Chen, F.; Liu, H.; Bagwasi, S.; Shen, X.; Zhang, J. Photocatalytic study of BiOCl for degradation of organic pollutants under UV irradiation. J. Photochem. Photobio. A 2010, 215, 76–80. [Google Scholar] [CrossRef]
  124. Zhang, L.; Yuhan, L.; Li, Q.; Fan, J.; Carabineiro, S.A.C.; Lv, K. Recent advances on bismuth-based photocatalysts: Strategies and mechanisms. Chem. Eng. J 2021, 419, 129484. [Google Scholar] [CrossRef]
  125. Li, H.; Zhang, L. Photocatalytic performance of different exposed crystal facets of BiOCl. Cur. Opin. Green Sust. Chem. 2017, 6, 48–56. [Google Scholar] [CrossRef]
  126. Liu, W.-W.; Peng, R.-F. Recent advances of bismuth oxychloride photocatalytic material: Property, preparation and performance enhancement. J. Electron. Sci. Technol. 2020, 18, 100020. [Google Scholar] [CrossRef]
  127. Li, F.; Wang, Q.; Wang, X.; Li, B.; Hao, Y.; Liu, R.; Zhao, D. In-situ one-step synthesis of novel BiOCl/Bi24O31Cl10 heterojunctions via self-combustion of ionic liquid with enhanced visible-light photocatalytic activities. Appl. Catal. B Environ. 2014, 150–151, 574–584. [Google Scholar] [CrossRef]
  128. Bai, Y.; Chen, T.; Wang, P.; Wang, L.; Ye, L. Bismuth-rich Bi4O5X2 (X=Br, I) nanosheets with dominant {1,0,1} facets exposure for photocatalytic H2 evolution. Chem. Eng. J. 2016, 304, 454–460. [Google Scholar] [CrossRef]
  129. Jiang, J.; Zhao, K.; Xiao, X.; Zhang, L. Synthesis and facet-dependent photoreactivity of BiOCl single-crystalline nanosheets. J. Am. Chem. Soc. 2012, 134, 4473–4476. [Google Scholar] [CrossRef]
  130. Zhao, H.; Liu, X.; Dong, Y.; Xia, Y.; Wang, H. A special synthesis of BiOCl photocatalyst for efficient pollutants removal: New insight into the band structure regulation and molecular oxygen activation. Appl. Catal. B Environ. 2019, 256, 117872. [Google Scholar] [CrossRef]
  131. Dong, X.; Cui, W.; Wang, H.; Li, J.; Sun, Y.; Wang, H.; Zhang, Y.; Huang, H.; Dong, F. Promoting ring-opening efficiency for suppressing toxic intermediates during photocatalytic toluene degradation via surface oxygen vacancies. Sci. Bull. 2019, 64, 669–678. [Google Scholar] [CrossRef] [Green Version]
  132. Wang, C.Y.; Zhang, Y.J.; Wang, W.K.; Pei, D.N.; Huang, G.X.; Chen, J.J.; Zhang, X.; Yu, H.Q. Enhanced photocatalytic degradation of bisphenol A by Co-doped BiOCl nanosheets under visible light irradiation. Appl. Catal. B Environ. 2018, 221, 320–328. [Google Scholar] [CrossRef]
  133. Jia, Z.; Li, T.; Zheng, Z.; Zhang, J.; Liu, J.; Li, R.; Wang, Y.; Zhang, X.; Wang, Y.; Fan, C. The BiOCl/diatomite composites for rapid photocatalytic degradation of ciprofloxacin: Efficiency, toxicity evaluation, mechanisms and pathways. Chem. Eng. J. 2020, 380, 122422. [Google Scholar] [CrossRef]
  134. Alansi, A.M.; Qahtan, T.F.; Al-Abass, N.; Al-Qunaibit, M.; Saleh, T.A. In-situ sunlight-driven tuning of photo-induced electron-hole generation and separation rates in bismuth oxychlorobromide for highly efficient water decontamination under visible light irradiation. J. Colloid Interface Sci. 2022, 614, 58–65. [Google Scholar] [CrossRef] [PubMed]
  135. Di, J.; Chen, C.; Zhu, C.; Song, P.; Xiong, J.; Ji, M.; Zhou, J.; Fu, Q.; Xu, M.; Hao, W.; et al. Bismuth vacancy-tuned bismuth oxybromide ultrathin nanosheets toward photocatalytic CO2 reduction. ACS Appl. Mater. Interfaces 2019, 11, 30786–30792. [Google Scholar] [CrossRef]
  136. Sun, Y.F.; Wu, J.; Li, X.D.; Shi, W.; Ling, P.Q.; Jiao, X.C.; Gao, S.; Liang, L.; Xu, J.Q.; Yan, W.S.; et al. Efficient visible-light-driven CO2 Reduction realized by defect-mediated BiOBr atomic layers. Angew. Chem. Int. Ed. 2018, 57, 8719–8723. [Google Scholar]
  137. Wang, Q.; Liu, Z.; Liu, D.; Wang, W.; Zhao, Z.; Cui, F.; Li, G. Oxygen vacancy-rich ultrathin sulfur-doped bismuth oxybromide nanosheet as a highly efficient visible-light responsive photocatalyst for environmental remediation. Chem. Eng. J. 2019, 360, 838–847. [Google Scholar] [CrossRef]
  138. Yang, J.; Liang, Y.J.; Li, K.; Zhu, Y.L.; Liu, S.Q.; Xu, R.; Zhou, W. Design of 3D flowerlike BiOClxBr 1–x nanostructure with high surface area for visible light photocatalytic activities. J. Alloys Compd. 2017, 725, 1144–1157. [Google Scholar] [CrossRef]
  139. Liu, G.; Wang, T.; Ouyang, S.; Liu, L.; Jiang, H.; Yu, Q.; Kako, T.; Ye, J. Band-structure-controlled BiO(ClBr)(1–x)/2Ix solid solutions for visible-light photocatalysis. J. Mater. Chem. A 2015, 3, 8123–8132. [Google Scholar] [CrossRef]
  140. Dong, X.-D.; Zhang, Y.M.; Zhao, Z.Y. Role of the polar electric field in bismuth oxyhalides for photocatalytic water splitting. Inorg. Chem. 2021, 60, 8461–8474. [Google Scholar] [CrossRef]
  141. Wang, W.; Huang, B.; Ma, X.; Wang, Z.; Qin, X.; Zhang, X.; Dai, Y.; Whangbo, M.H. Efficient separation of photogenerated electron-hole pairs by the combination of a heterolayered structure and internal polar field in pyroelectric BiOIO3 nanoplates. Chem.-Eur. J. 2013, 19, 14777–14780. [Google Scholar] [CrossRef]
  142. Dong, X.-D.; Yao, G.-Y.; Liu, Q.-L.; Zhao, Q.-M.; Zhao, Z.-Y. Spontaneous polarization effect and photocatalytic activity of layered compound of BiOIO3. Inorg. Chem. 2019, 58, 15344–15353. [Google Scholar] [CrossRef] [PubMed]
  143. Chen, F.; Ma, Z.; Ye, L.; Ma, T.; Zhang, T.; Zhang, Y.; Huang, H. Macroscopic spontaneous polarization and surface oxygen vacancies collaboratively boosting CO2 photoreduction on BiOIO3 single crystals. Adv. Mater. 2020, 32, 1908350. [Google Scholar] [CrossRef] [PubMed]
  144. Huang, H.; Tu, S.; Zeng, C.; Zhang, T.; Reshak, A.H.; Zhang, Y. Macroscopic Polarization Enhancement Promoting Photo- and Piezoelectric-Induced Charge Separation and Molecular Oxygen Activation. Angew. Chem. Int. Ed. 2017, 56, 11860–11864. [Google Scholar] [CrossRef] [PubMed]
  145. Zhang, M.; Shi, Q.; Cheng, X.; Yang, J.; Liu, Z.; Chen, T.; Qu, Y.; Wang, J.; Xie, M.; Han, W. Accelerated generation of hydroxyl radical through surface polarization on BiVO4 microtubes for efficient chlortetracycline degradation. Chem. Eng. J. 2020, 400, 125871. [Google Scholar] [CrossRef]
  146. Cai, G.; Xu, L.; Wei, B.; Che, J.; Gao, H.; Sun, W. Facile synthesis of β-Bi2O3/Bi2O2CO3 nanocomposite with high visible-light photocatalytic activity. Mater. Lett. 2014, 120, 1–4. [Google Scholar] [CrossRef]
  147. Huang, H.; Xiao, K.; He, Y.; Zhang, T.; Dong, F.; Du, X.; Zhang, Y. In situ assembly of BiOI@Bi12O17Cl2 p-n junction: Charge induced unique front-lateral surfaces coupling heterostructure with high exposure of BiOI {001} active facets for robust and non selective photocatalysis. Appl. Catal. B Environ. 2016, 199, 75–86. [Google Scholar] [CrossRef]
  148. Yin, H.Y.; Zheng, Y.F.; Song, X.C. Synthesis and enhanced visible light photocatalytic CO2 reduction of BiPO4–BiOBrxI1−x p–n heterojunctions with adjustable energy band. RSC Adv. 2019, 9, 11005–11012. [Google Scholar] [CrossRef] [Green Version]
  149. Tang, J.; Xue, Y.; Ma, C.; Zhang, S.; Li, Q. Facile preparation of BiOI/T-ZnOw p–n heterojunction photocatalysts with enhanced removal efficiency for rhodamine B and oxytetracycline. New J. Chem. 2022, 46, 13010–13020. [Google Scholar] [CrossRef]
  150. Nie, J.; Zhu, G.; Zhang, W.; Gao, J.; Zhong, P.; Xie, X.; Huang, Y.; Hojamberdiev, M. Oxygen vacancy defects-boosted deep oxidation of NO by β-Bi2O3/CeO2-δ p-n heterojunction photocatalyst in situ synthesized from Bi/Ce(CO3) (OH) precursor. Chem. Eng. J. 2021, 424, 130327. [Google Scholar] [CrossRef]
  151. Su, X.; Fan, D.; Sun, H.; Yang, J.; Yu, Z.; Zhang, D.; Pu, X.; Li, H.; Cai, P. One-dimensional rod-shaped Ag2Mo2O7/BiOI n-n junctions for efficient photodegradation of tetracycline and rhodamine B under visible light. J. Alloys Compd. 2022, 912, 165184. [Google Scholar] [CrossRef]
  152. Li, Y.; Li, Z.; Xia, Y.; Gao, L. AgBr/BiOI/g-C3N4 Photocatalyst with enhanced photocatalytic activity under visible-light irradiation via the formation of double Z-type heterojunction with the synergistic effect of metal Ag. Ind. Eng. Chem. Res. 2022, 61, 12918–12930. [Google Scholar] [CrossRef]
  153. Sahu, R.S.; Shih, Y.; Chen, W. New insights of metal free 2D graphitic carbon nitride for photocatalytic degradation of bisphenol A. J. Hazard. Mater. 2021, 402, 123509. [Google Scholar] [CrossRef] [PubMed]
  154. Jing, L.; He, M.; Xie, M.; Song, Y.; Wei, W.; Xu, Y.; Xu, H.; Li, H. Realizing the synergistic effect of electronic modulation over graphitic carbon nitride for highly efficient photodegradation of bisphenol A and 2-mercaptobenzothiazole: Mechanism, degradation pathway and density functional theory calculation. J. Colloid Interface Sci. 2021, 583, 113–127. [Google Scholar] [CrossRef] [PubMed]
  155. Ruan, X.; Hu, Y. Effectively enhanced photodegradation of bisphenol A by in-situ g-C3N4-Zn/Bi2WO6 heterojunctions and mechanism study. Chemosphere 2020, 246, 125782. [Google Scholar] [CrossRef]
  156. Du, F.; Lai, Z.; Tang, H.; Wang, H.; Zhao, C. Construction of dual Z-scheme Bi2WO6/g-C3N4/black phosphorus quantum dots composites for effective bisphenol A degradation. J. Environ. Sci. 2023, 124, 617–629. [Google Scholar] [CrossRef]
  157. Deng, F.; Zhang, Q.; Yang, L.X.; Luo, X.B.; Wang, A.J.; Luo, S.L.; Dionysiou, D.D. Visible-light-responsive graphene-functionalized Bi-bridge Z-scheme black BiOCl/Bi2O3 heterojunction with oxygen vacancy and multiple charge transfer channels for efficient photocatalytic degradation of 2-nitrophenol and industrial wastewater treatment. Appl. Catal. B Environ. 2018, 238, 61–69. [Google Scholar] [CrossRef]
  158. Liu, J.; Shu, S.; Li, Y.; Liu, J.; Yao, J.; Liu, S.; Zhu, M.; Huang, L.; Huang, L. Ternary hybrid Ag/SnO2-X/Bi4O5I2 photocatalysts: Impressive efficiency for photocatalytic degradation of antibiotics and inactivation of bacteria. Appl. Surf. Sci. 2022, 606, 154610. [Google Scholar] [CrossRef]
  159. Zhang, Q.; Ravindra, X.H.; Zhang, L.; Zeng, K.; Xu, Y.; Xin, C. Microwave hydrothermal synthesis of a Bi2SiO5/Bi12SiO20 heterojunction with oxygen vacancies and multiple charge transfer for enhanced photocatalytic activity. Appl. Surf. Sci. 2022, 581, 152337. [Google Scholar] [CrossRef]
  160. Weber, M.; Rodriguez, R.D.; Zahn, D.R.T.; Mehring, M. γ-Bi2O3—Presence or absence? Comparison of sillimanite γ-Bi2O3 and isocrystalline Bi12SiO20. Inorg. Chem. 2018, 57, 8540–8549. [Google Scholar] [CrossRef]
  161. Guo, J.H.; Shi, L.; Zhao, J.Y.; Wang, Y.; Tang, K.B.; Zhang, W.Q.; Xie, C.Z.; Yuan, X.Y. Enhanced visible-light photocatalytic activity of Bi2MoO6 nanoplates with heterogeneous Bi2MoO6-x@Bi2MoO6 core-shell structure. Appl. Catal. B Environ. 2018, 224, 692–704. [Google Scholar] [CrossRef]
  162. Yuan, S.; Zhao, Y.; Chen, W.; Wu, C.; Wang, X.; Zhang, L.; Wang, Q. Self-assembled 3D hierarchical porous Bi2MoO6 microspheres toward high capacity and ultra-long-life anode material for Li-ion batteries. ACS Appl. Mater. Interfaces 2017, 9, 21781–21790. [Google Scholar] [CrossRef] [PubMed]
  163. Lu, H.J.; Hao, Q.; Chen, T.; Zhang, L.H.; Chen, D.M.; Ma, C.; Yao, W.Q.; Zhu, Y.F. A high-performance Bi2O3 /Bi2SiO5 p-n heterojunction photocatalyst induced by phase transition of Bi2O3. Appl. Catal. B Environ. 2018, 237, 59–67. [Google Scholar] [CrossRef]
  164. Dou, L.; Jin, X.; Chen, J.; Zhong, J.; Li, J.; Zeng, Y.; Duan, R. One-pot solvothermal fabrication of S-scheme OVs-Bi2O3/Bi2SiO5 microsphere heterojunctions with enhanced photocatalytic performance toward decontamination of organic pollutants. Appl. Surf. Sci. 2020, 527, 146775. [Google Scholar] [CrossRef]
  165. He, R.; Liu, H.; Liu, H.; Xu, D.; Zhang, L. S-scheme photocatalyst Bi2O3/TiO2 nanofiber with improved photocatalytic performance. J. Mater. Sci. Technol. 2020, 52, 145–151. [Google Scholar]
  166. Xu, Y.F.; Zhou, Y.; Deng, Y.H.; Xiang, Y.; Tan, Y.W.; Tang, H.Q.; Zou, H. Synthesis of Bi2WO6@NH2-MIL-125(Ti): A S-scheme photocatalyst with enhanced visible light catalytic activity. Green Energy Environ. 2020, 5, 203–213. [Google Scholar] [CrossRef]
  167. Li, X.B.; Xiong, J.; Gao, X.M.; Ma, J.; Chen, Z.; Kang, B.B.; Liu, J.Y.; Li, H.; Feng, Z.J.; Huang, J.T. Novel BP/BiOBr S-scheme nano-heterojunction for enhanced visible-light photocatalytic tetracycline removal and oxygen evolution activity. J. Hazard. Mater. 2020, 387, 121690. [Google Scholar] [CrossRef]
  168. Xie, Q.; He, W.; Liu, S.; Li, C.; Zhang, J.; Wong, P.K. Bifunctional S-scheme g-C3N4/Bi/BiVO4 hybrid photocatalysts toward artificial carbon cycling. Chin. J. Catal. 2020, 41, 140–153. [Google Scholar] [CrossRef]
  169. Zhao, Z.; Zhang, W.; Sun, Y.; Yu, J.; Zhang, Y.; Wang, H.; Dong, F.; Bi, Z.W. Cocatalyst/Bi2MoO6 microspheres nanohybrid with SPR-Promoted visible-light photocatalysis. J. Phys. Chem. C 2016, 120, 11889–11898. [Google Scholar] [CrossRef]
  170. Weng, S.; Chen, B.; Xie, L.; Zheng, Z.; Liu, P. Facile in situ synthesis of a Bi/BiOCl nanocomposite with high photocatalytic activity. J. Mater. Chem. A 2013, 1, 3068. [Google Scholar] [CrossRef]
  171. Wang, H.; Zhang, W.D.; Li, W.D.; Li, J.Y.; Cen, W.L.; Li, Q.Y.; Dong, F. Highly enhanced visible light photocatalysis and in situ FT-IR studies on Bi metal@defective BiOCl hierarchical microspheres. Appl. Catal. B Environ. 2018, 225, 218–227. [Google Scholar] [CrossRef]
  172. Jiang, G.M.; Li, X.W.; Lan, M.N.; Shen, T.; Lv, X.S.; Dong, F.; Zhang, S. Monodisperse bismuth nanoparticles decorated graphitic carbon nitride: Enhanced visible-light-response photocatalytic NO removal and reaction pathway. Appl. Catal. B Environ. 2017, 205, 532–540. [Google Scholar] [CrossRef]
  173. Wang, J.; Shi, Y.; Sun, H.; Shi, W.; Guo, F. Fabrication of Bi4Ti3O12/ZnIn2S4 S-scheme heterojunction for achieving efficient photocatalytic hydrogen production. J. Alloys Compd. 2023, 930, 167450. [Google Scholar] [CrossRef]
  174. Chen, Y.; Zhang, Y.; Luo, L.; Shi, Y.; Wang, S.; Li, L.; Long, Y.; Jiang, F. A novel templated synthesis of C/N-doped β-Bi2O3 nanosheets for synergistic rapid removal of 17α-ethynylestradiol by adsorption and photocatalytic degradation. Ceram. Int. 2018, 44, 2178–2185. [Google Scholar] [CrossRef]
  175. Shahid, M.; Bashir, S.; Afzal, A.; Ibn Shamsah, S.M.; Jamil, A. Synergistic impacts of composite formation and doping techniques to boost the photocatalytic aptitude of the BiFeO3 nanostructure. Ceram. Int. 2022, 48, 2566–2576. [Google Scholar] [CrossRef]
  176. Sharmin, F.; Basith, M.A. Simple low temperature technique to synthesize sillenite bismuth ferrite with promising photocatalytic performance. ACS Omega 2022, 7, 34901–34911. [Google Scholar] [CrossRef]
  177. Tien, L.-C.; Lin, Y.-L.; Chen, S.-Y. Synthesis and characterization of Bi12O17Cl2 nanowires obtained by chlorination of α-Bi2O3 nanowires. Mater. Lett. 2013, 113, 30–33. [Google Scholar] [CrossRef]
  178. Prabhakar Vattikuti, S.V.; Bach, L.X.; Devarayapalli, K.C.; Reddy, A.N.R.; Shim, J.; Julien, C.M. Morphology-dependent photocatalytic and photoelectrochemical performance of bismuth oxybromide crystals applied to malachite green dye degradation. Colloids Surf. A Physicochem. Eng. Asp. 2022, 655, 130267. [Google Scholar] [CrossRef]
  179. Liu, Z.; Wu, B.T.; Xiang, D.; Zhu, Y. Effect of solvents on morphology and photocatalytic activity of BiOBr synthesized by solvothermal method. Mater. Res. Bull. 2012, 47, 3753–3757. [Google Scholar] [CrossRef]
  180. Patnam, H.; Bharat, L.K.; Hussain, S.K.; Yu, J.S. Effect of solvents on the morphology and optical properties of rare-earth ions doped BiOBr 3D flower-like microparticles via solvothermal method. J. Alloys Compd. 2018, 763, 478–485. [Google Scholar] [CrossRef]
  181. Liu, Z.; Lv, F.; Xiao, Y.; Chen, B.; Qiu, J.; Guo, W.; Wen, Y.; Li, Z.; Liu, Z. Morphology controllable synthesis of BiOBr architectures and their visible light photocatalytic activities. Mater. Technol. 2019, 34, 683–688. [Google Scholar] [CrossRef]
  182. Shi, Z.; Zhang, Y.; Shen, X.F.; Duoerkun, G.; Zhu, B.; Zhang, L.S.; Li, M.Q.; Chen, Z.G. Fabrication of g-C3N4/BiOBr heterojunctions on carbon fibers as weaveable photocatalyst for degrading tetracycline hydrochloride under visible light. Chem. Eng. J. 2020, 386, 124010. [Google Scholar] [CrossRef]
  183. Peng, P.; Chen, Z.; Li, X.M.; Wu, Y.; Xia, Y.T.; Duan, A.B.; Wang, D.B.; Yang, Q. Biomass-derived carbon quantum dots modified Bi2MoO6/Bi2S3 heterojunction for efficient photocatalytic removal of organic pollutants and Cr (VI). Sep. Purif. Technol. 2022, 291, 120901. [Google Scholar] [CrossRef]
  184. Bi, H.F.; Liu, J.S.; Wu, Z.Y.; Zhu, K.J.; Suo, H.; Lv, X.L.; Fu, Y.L.; Jian, R.; Sun, Z.B. Construction of Bi2WO6/ZnIn2S4 with Z-scheme structure for efficient photocatalytic performance. Chem. Phys. Lett. 2021, 769, 138449. [Google Scholar] [CrossRef]
  185. Zhou, Z.; Li, Y.; Lv, K.; Wu, X.; Li, Q.; Luo, J. Fabrication of walnut-like BiVO4@Bi2S3 heterojunction for efficient visible photocatalytic reduction of Cr(VI). Mater. Sci. Semicond. Process. 2018, 75, 334–341. [Google Scholar] [CrossRef]
  186. Madhusudan, P.; Ran, J.; Zhang, J.; Yu, J.; Liu, G. Novel urea assisted hydrothermal synthesis of hierarchical BiVO4/Bi2O2CO3 nanocomposites with enhanced visible-light photocatalytic activity. Appl. Catal. B Environ. 2011, 110, 286–295. [Google Scholar] [CrossRef]
  187. Wang, G.; Cheng, D.; He, T.; Hu, Y.; Deng, Q.; Mao, Y.; Wang, S. Enhanced visible- light responsive photocatalytic activity of Bi25FeO40/Bi2Fe4O9 composites and mechanism investigation. J. Mater. Sci. Mater. Electron. 2019, 30, 10923–10933. [Google Scholar] [CrossRef]
  188. Ren, X.; Wu, K.; Qin, Z.; Zhao, X.; Yang, H. The Construction of type II heterojunction of Bi2WO6/BiOBr photocatalyst with improved photocatalytic performance. J. Alloys Compd. 2019, 788, 102–109. [Google Scholar] [CrossRef]
  189. Guo, S.; Luo, H.; Li, Y.; Chen, J.; Mou, B.; Shi, X.; Sun, G. Structure-controlled three-dimensional BiOI/MoS2 microspheres for boosting visible-light photocatalytic degradation of tetracycline. J. Alloys Compd. 2021, 852, 157026. [Google Scholar] [CrossRef]
  190. Sun, Q.; Hong, Y.; Liu, Q.; Dong, L. Synergistic operation of photocatalytic degradation and Fenton process by magnetic Fe3O4 loaded TiO2. Appl. Surf. Sci. 2018, 430, 399–406. [Google Scholar] [CrossRef]
  191. Jiao, S.; Zhao, Y.; Li, C.; Wang, B.; Qu, Y. Recyclable adsorbent of BiFeO3/Carbon for purifying industrial dye wastewater via photocatalytic reproducible. Green Energy Environ. 2019, 4, 66–74. [Google Scholar] [CrossRef]
  192. Subramanian, Y.; Ramasamy, V.; Karthikeyan, R.J.; Srinivasan, G.R.; Arulmozhi, D.; Gubendiran, R.K.; Sriramalu, M. Investigations on the enhanced dye degradation activity of heterogeneous BiFeO3–GdFeO3 nanocomposite photocatalyst. Heliyon 2019, 5, e01831. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Si, Y.; Li, Y.; Zou, J.; Xiong, X.; Zeng, X.; Zhou, J. Photocatalytic performance of a novel MOF/BiFeO3 composite. Materials 2017, 10, 1161. [Google Scholar] [CrossRef] [PubMed]
  194. Zhao, W.; Wang, Y.; Yang, Y.; Tang, J.; Yang, Y. Carbon spheres supported visible light-driven CuO-BiVO4 heterojunction: Preparation, characterization, and photocatalytic properties. Appl. Catal. B Environ. 2012, 115–116, 90–99. [Google Scholar] [CrossRef]
  195. Liu, Y.; Yang, Z.-H.; Song, P.-P.; Xu, R.; Wang, H. Facile synthesis of Bi2MoO6/ZnSnO3 heterojunction with enhanced visible light photocatalytic degradation of methylene blue. Appl. Surf. Sci. 2018, 430, 561–570. [Google Scholar] [CrossRef]
  196. Wang, Q.; Jiao, D.; Lian, J.; Ma, Q.; Yu, J.; Huang, H.; Zhong, J.; Li, J. Preparation of efficient visible-light-driven BiOBr/Bi2O3 heterojunction composite with enhanced photocatalytic activities. J. Alloys Compd. 2015, 649, 474–482. [Google Scholar] [CrossRef]
  197. Chaiwichian, S.; Wetchakun, K.; Kangwansupamonkon, W.; Wetchakun, N. Novel visible-light-driven BiFeO3-Bi2WO6 nanocomposites toward degradation of dyes. J. Photochem. Photobiol. A Chem. 2017, 349, 183–192. [Google Scholar] [CrossRef]
  198. Shtarev, D.S.; Serpone, N. A new generation of visible-light-active photocatalysts—The alkaline earth metal bismuthates: Syntheses, compositions, structures, and properties. J. Photochem. Photobiol. C Photochem. Rev. 2022, 50, 100501. [Google Scholar] [CrossRef]
  199. Shtarev, S.; Kevorkyants, R.; Molokeev, M.S.; Shtareva, A.V. The effect of composition on optical and photocatalytic properties of visible light response materials Bi26-xMgxO40. Inorg. Chem. 2020, 59, 8173–8183. [Google Scholar] [CrossRef]
  200. Li, W.; Kong, D.; Cui, X.; Du, D.; Yan, T.; You, J. Hydrothermal synthesis of Ca3Bi8O15 rods and their visible light photocatalytic properties. Mater. Res. Bull. 2014, 51, 69–73. [Google Scholar] [CrossRef]
  201. Yang, Y.; Wang, X.; Qu, J. Preparation and photocatalytic degradation of malachite green by photocatalyst SrBi4O7 under visible light irradiation. Appl. Mech. Mater. 2014, 522–524, 411–415. [Google Scholar] [CrossRef]
Figure 1. Degradation mechanisms of antibiotics by bismuth-based photocatalysts: absorption of photons with energy higher than the material bandgap, excitation, and reaction. Reproduced with permission from [45]. Copyright 2021 Elsevier.
Figure 1. Degradation mechanisms of antibiotics by bismuth-based photocatalysts: absorption of photons with energy higher than the material bandgap, excitation, and reaction. Reproduced with permission from [45]. Copyright 2021 Elsevier.
Ijms 24 00663 g001
Figure 2. The band edge positions of bismuth-based photocatalysts. Almost all the tops of the valence bands of bismuth-based photocatalysts are higher than the redox potential of •OH/OH (+1.99 eV). Reproduced with permission from [45]. Copyright 2021 Elsevier.
Figure 2. The band edge positions of bismuth-based photocatalysts. Almost all the tops of the valence bands of bismuth-based photocatalysts are higher than the redox potential of •OH/OH (+1.99 eV). Reproduced with permission from [45]. Copyright 2021 Elsevier.
Ijms 24 00663 g002
Figure 3. Heterojunction types by band position: (a) type I (straddling gap), (b) type II (staggered gap), (c) direct Z-scheme or mediator-free, (d) solid mediator, and (e) redox pair mediator. Reproduced with permission from [54]. Copyright 2020 Elsevier.
Figure 3. Heterojunction types by band position: (a) type I (straddling gap), (b) type II (staggered gap), (c) direct Z-scheme or mediator-free, (d) solid mediator, and (e) redox pair mediator. Reproduced with permission from [54]. Copyright 2020 Elsevier.
Ijms 24 00663 g003
Figure 4. Sketch of growth mechanisms of α-Bi2O3 nanowires. Processes I–II are driven by the Au-catalyzed mechanism for the formation of the (010) facet; processes III–IV are driven by the Bi-catalyzed mechanism for the formation of beaklike nanowires. Reproduced with permission from [78]. Copyright 2014 Elsevier.
Figure 4. Sketch of growth mechanisms of α-Bi2O3 nanowires. Processes I–II are driven by the Au-catalyzed mechanism for the formation of the (010) facet; processes III–IV are driven by the Bi-catalyzed mechanism for the formation of beaklike nanowires. Reproduced with permission from [78]. Copyright 2014 Elsevier.
Ijms 24 00663 g004
Figure 5. Possible reaction mechanism of the silver–bismuthate system. Reproduced with permission from [99]. Copyright 2022 Elsevier.
Figure 5. Possible reaction mechanism of the silver–bismuthate system. Reproduced with permission from [99]. Copyright 2022 Elsevier.
Ijms 24 00663 g005
Figure 6. Schematic drawing of the crystal structure of Bi2SiO5 (a) and ECD using the MEM distribution of the Bi2O2 (b,c) and the SiO3 layer (d,e) in the ferroelectric (300 K) and paraelectric (773 K) phases. The isosurface of ECD is 0.85 e Å−3 and 1.50 e Å−3 for the Bi2O2 and the SiO3 layers, respectively. Reproduced with permission from [103]. Copyright 2014 under Creative Commons Attribution (CC-BY) Licence.
Figure 6. Schematic drawing of the crystal structure of Bi2SiO5 (a) and ECD using the MEM distribution of the Bi2O2 (b,c) and the SiO3 layer (d,e) in the ferroelectric (300 K) and paraelectric (773 K) phases. The isosurface of ECD is 0.85 e Å−3 and 1.50 e Å−3 for the Bi2O2 and the SiO3 layers, respectively. Reproduced with permission from [103]. Copyright 2014 under Creative Commons Attribution (CC-BY) Licence.
Ijms 24 00663 g006
Figure 7. Energy band structures of Bi2SiO5 (a) and Bi4MoO9 (b). Density of states of Bi2SiO5 (c) and Bi4MoO9 (d). Reproduced with permission from [107]. Copyright 2020 under Creative Commons Attribution (CC-BY) Licence.
Figure 7. Energy band structures of Bi2SiO5 (a) and Bi4MoO9 (b). Density of states of Bi2SiO5 (c) and Bi4MoO9 (d). Reproduced with permission from [107]. Copyright 2020 under Creative Commons Attribution (CC-BY) Licence.
Ijms 24 00663 g007
Figure 8. Schematic illustration of the band-gap structure and possible flow of charge carriers through the ternary heterostructure under visible light irradiation. Reproduced with permission from [113]. Copyright 2015 Elsevier.
Figure 8. Schematic illustration of the band-gap structure and possible flow of charge carriers through the ternary heterostructure under visible light irradiation. Reproduced with permission from [113]. Copyright 2015 Elsevier.
Ijms 24 00663 g008
Figure 9. Schematic illustration of the SPR effect on the enhanced photoreactivity of Bi2WO6 toward NO oxidation after being loaded with Bi nanospheres. Reproduced with permission from [63]. Copyright 2019 Elsevier.
Figure 9. Schematic illustration of the SPR effect on the enhanced photoreactivity of Bi2WO6 toward NO oxidation after being loaded with Bi nanospheres. Reproduced with permission from [63]. Copyright 2019 Elsevier.
Ijms 24 00663 g009
Figure 10. Schematic diagrams for (a) the efficient charge separation process through the {001} active facets of BiOI and (b) the proposed charge-transfer mechanism via the BiOI@Bi12O17Cl2 p–n junction [147]. Copyright 2016 Elsevier.
Figure 10. Schematic diagrams for (a) the efficient charge separation process through the {001} active facets of BiOI and (b) the proposed charge-transfer mechanism via the BiOI@Bi12O17Cl2 p–n junction [147]. Copyright 2016 Elsevier.
Ijms 24 00663 g010
Figure 11. Schematic diagrams of the energy levels of T-ZnOw and BiOI (a) before contact and (b) after the formation of a p–n BiOI/T-ZnOw heterojunction and possible photocatalytic degradation mechanism under visible light irradiation. Reproduced with permission from [149]. Copyright 2022 The Royal Society of Chemistry.
Figure 11. Schematic diagrams of the energy levels of T-ZnOw and BiOI (a) before contact and (b) after the formation of a p–n BiOI/T-ZnOw heterojunction and possible photocatalytic degradation mechanism under visible light irradiation. Reproduced with permission from [149]. Copyright 2022 The Royal Society of Chemistry.
Ijms 24 00663 g011
Figure 12. X-ray diffraction patterns (a) and Raman spectra (b) of CeO2-δ, β-Bi2O3, and β-Bi2O3/CeO2-δ samples. (c) Scanning electron microscopy images and schematic representation of the formation process of 4% β-Bi2O3/CeO2-δ. (d,e) Transmission electron microscopy (TEM) and high-resolution TEM showing the various species in sample (f,g) images of 4% β-Bi2O3/CeO2-δ. Reproduced with permission from [150]. Copyright 2021 Elsevier.
Figure 12. X-ray diffraction patterns (a) and Raman spectra (b) of CeO2-δ, β-Bi2O3, and β-Bi2O3/CeO2-δ samples. (c) Scanning electron microscopy images and schematic representation of the formation process of 4% β-Bi2O3/CeO2-δ. (d,e) Transmission electron microscopy (TEM) and high-resolution TEM showing the various species in sample (f,g) images of 4% β-Bi2O3/CeO2-δ. Reproduced with permission from [150]. Copyright 2021 Elsevier.
Ijms 24 00663 g012
Figure 13. A plausible photocatalytic reaction mechanism diagram for an n–n Ag2Mo2O7/BiOI nanostructure. Reproduced with permission from [151]. Copyright 2022 Elsevier.
Figure 13. A plausible photocatalytic reaction mechanism diagram for an n–n Ag2Mo2O7/BiOI nanostructure. Reproduced with permission from [151]. Copyright 2022 Elsevier.
Ijms 24 00663 g013
Figure 14. Double Z-type electron transfer mechanism of heterojunction for the photodegradation of MO based on AgBr/BiOI/g-C3N4. Reproduced with permission from [152]. Copyright 2022 American Chemical Society.
Figure 14. Double Z-type electron transfer mechanism of heterojunction for the photodegradation of MO based on AgBr/BiOI/g-C3N4. Reproduced with permission from [152]. Copyright 2022 American Chemical Society.
Ijms 24 00663 g014
Figure 15. (a) Schematic diagram for the enhanced photogenerated electron transfer processes induced by OVs. (b) Schematic diagram for the migration and separation of electron-hole pairs and the photocatalytic process of the Bi2SiO5/Bi12SiO20 heterojunction photocatalyst: Crystal structure of Bi2SiO5 (left) (SiBiO4) tetrahedron and (BiO5) pyramids in Bi12SiO20 crystal structure (right). Reproduced with permission from [159]. Copyright 2022 Elsevier.
Figure 15. (a) Schematic diagram for the enhanced photogenerated electron transfer processes induced by OVs. (b) Schematic diagram for the migration and separation of electron-hole pairs and the photocatalytic process of the Bi2SiO5/Bi12SiO20 heterojunction photocatalyst: Crystal structure of Bi2SiO5 (left) (SiBiO4) tetrahedron and (BiO5) pyramids in Bi12SiO20 crystal structure (right). Reproduced with permission from [159]. Copyright 2022 Elsevier.
Ijms 24 00663 g015
Figure 16. The photo-excited electron-hole separation process over OVs-Bi2O3/Bi2SiO5 composite photocatalysts. Reproduced with permission from [164]. Copyright 2020 Elsevier.
Figure 16. The photo-excited electron-hole separation process over OVs-Bi2O3/Bi2SiO5 composite photocatalysts. Reproduced with permission from [164]. Copyright 2020 Elsevier.
Ijms 24 00663 g016
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Prabhakar Vattikuti, S.V.; Zeng, J.; Ramaraghavulu, R.; Shim, J.; Mauger, A.; Julien, C.M. High-Throughput Strategies for the Design, Discovery, and Analysis of Bismuth-Based Photocatalysts. Int. J. Mol. Sci. 2023, 24, 663. https://doi.org/10.3390/ijms24010663

AMA Style

Prabhakar Vattikuti SV, Zeng J, Ramaraghavulu R, Shim J, Mauger A, Julien CM. High-Throughput Strategies for the Design, Discovery, and Analysis of Bismuth-Based Photocatalysts. International Journal of Molecular Sciences. 2023; 24(1):663. https://doi.org/10.3390/ijms24010663

Chicago/Turabian Style

Prabhakar Vattikuti, Surya V., Jie Zeng, Rajavaram Ramaraghavulu, Jaesool Shim, Alain Mauger, and Christian M. Julien. 2023. "High-Throughput Strategies for the Design, Discovery, and Analysis of Bismuth-Based Photocatalysts" International Journal of Molecular Sciences 24, no. 1: 663. https://doi.org/10.3390/ijms24010663

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop