Recent trends in Bi-based nanomaterials: challenges, fabrication, enhancement techniques, and environmental applications

  1. 1 ORCID Logo ,
  2. 2 ORCID Logo ,
  3. 3 ,
  4. 4 and
  5. 2 ORCID Logo
1School of Advanced Chemical Sciences, Shoolini University, Solan, Himachal Pradesh 173212, India
2Chettinad Hospital and Research Institute, Chettinad Academy of Research and Education (CARE), Chengalpattu district, Kelambakkam, Tamil Nadu, 603103, India
3University Centre for Research and Development, Chandigarh University, 140413, India
  1. Corresponding author email
Guest Editor: V. V. Pham
Beilstein J. Nanotechnol. 2022, 13, 1316–1336. https://doi.org/10.3762/bjnano.13.109
Received 15 Jul 2022, Accepted 13 Oct 2022, Published 11 Nov 2022
Review
cc by logo
Album

Abstract

One of the most enticing approaches to environmental restoration and energy conversion is photocatalysis powered by solar light. Traditional photocatalysts have limited practical uses due to inadequate light absorption, charge separation, and unknown reaction mechanisms. Discovering new visible-light photocatalysts and investigating their modification is crucial in photocatalysis. Bi-based photocatalytic nanomaterials have gotten much interest as they exhibit distinctive geometric shapes, flexible electronic structures, and good photocatalytic performance under visible light. They can be employed as stand-alone photocatalysts for pollution control and energy production, but they do not have optimum efficacy. As a result, their photocatalytic effectiveness has been significantly improved in the recent decades. Numerous newly created concepts and methodologies have brought significant progress in defining the fundamental features of photocatalysts, upgrading the photocatalytic ability, and understanding essential reactions of the photocatalytic process. This paper provides insights into the characteristics of Bi-based photocatalysts, making them a promising future nanomaterial for environmental remediation. The current review discusses the fabrication techniques and enhancement in Bi-based semiconductor photocatalysts. Various environmental applications, such as H2 generation and elimination of water pollutants, are also discussed in terms of semiconductor photocatalysis. Future developments will be guided by the uses, issues, and possibilities of Bi-based photocatalysts.

Introduction

Nanomaterials photocatalysis is a “green” integrative technique that combines physics, chemistry, and materials science with chemical engineering to catalyze chemical processes and transform constantly recoverable solar energy into productive chemical energy [1]. Various semiconductor nanoparticles have been used as effective photocatalysts in essential photocatalytic applications such as wastewater treatment, water dissociation, and energy conversion/storage due to their reactivity, surface area, and advantageous features compared to their bulk counterparts [2-4]. In recent years, many efforts have increased the photocatalytic performance. However, the relative photocatalytic performance is still deficient, and it does not fulfil the criteria for the practical implementation of photocatalysis techniques. Among many approaches, attention has been paid to altering and modifying the properties of photocatalytic materials [5,6]. Environmental treatment and energy conversion using photocatalytic technology have shown to be cost-effective and environmentally beneficial alternatives [7]. The choice of the photocatalysts is one of the most important steps in attaining high performance in photocatalysis. Semiconductors with bandgaps greater than 3 eV are called wide-bandgap photocatalysts. These semiconductors include oxides (e.g., TiO2, Bi2O3, Bi2WO6, and SrTiO3), sulfates (e.g., MoS2 and Bi2S3), selenides (e.g., MoSe2 and CdSe), and phosphates (e.g., Ag3PO4) [8-15].

The bandgap of photocatalysts sensitive to visible light is smaller than 3 eV. Wide-bandgap photocatalysts can only be stimulated by ultraviolet light, which makes up less than 5% of the absorbed solar radiation. Hence, developing photocatalysts that react to visible light is essential for photocatalysis since 43% of the total energy from the sun belongs to the visible spectrum [1,16]. Bi-based nanomaterials are photocatalysts that respond to visible light and have adequate bandgaps and performance. Bi2MO6 (M = Mo, W), (BiO)2CO3, Bi2S3, BiOX (X = I, Br, Cl), BiPO4, BiVO4, Bi2O3, and other Bi-based nanomaterials have been designed and examined for photocatalysis. The vast majority of these compounds have a layered structure, which causes an internal electric field (IEF) between the layers. This electric field allows photogenerated charge carriers to be separated and moved effectively [17-21].

A range of visible-light-active Bi-based photocatalysts has lately raised curiosity among semiconductor photocatalysts. Bi3+ has a higher stability than Bi5+. The earlier compounds have been examined more thoroughly than the latter. The overlap of O 2p and Bi 6s orbitals in the valence band (VB) of Bi3+-containing compounds improves photogenerated charge mobility and, hence, improves photocatalytic activity [22,23]. Furthermore, except for BiOF, BiOCl, and (BiO)2CO3, most Bi-based compounds have bandgaps that may be stimulated by visible light. As a result, much interest in environmental cleanup and energy conversion research has been sparked by Bi compounds [24]. Because of their advanced photocatalytic process, more and more publications on synthesizing and applying a semiconductor photocatalyst have been published in recent years. A survey on bismuth-based nanocomposites with the search keywords "Bismuth-based nanoparticles for environmental remediation" from 2011 to 2021 yields roughly 15,995 articles. This data illustrates the interest of the scientific community in environmental cleanup using bismuth-based nanoparticles (Figure 1). In recent years, an abundance of Bi-based photocatalysts has been reported.

[2190-4286-13-109-1]

Figure 1: Reported publications from 2011 to 2021 were searched on 6th June 2022 using the keyword “Bismuth-based nanoparticles for environmental remediation”. Data was extracted from Web of Science, Clarivate Analytics.

The most commonly used Bi-based photocatalysts include metallic Bi, Bi-based binary oxides, Bi-based oxyhalides, Bi-based multicomponent oxides, and binary Bi sulfides. Bismuth oxyhalides are indirect bandgap semiconductors in which photogenerated electrons and holes rarely recombine. BiOX is an excellent photocatalyst, and it is widely applied due to its small bandgap and high electron density, which are easily adjustable by changing the type of halogen used. The activation of otherwise inert CO2 and H2O molecules is greatly aided by the ease with which photoinduced oxygen vacancies (OVs) are produced on the surface [25]. The excellent photocatalytic performance can be attributed to the layered crystal structures and small bandgap energies. Many persistent organic contaminants can be degraded at room temperature through the oxidizing power of VB holes in bismuth oxyhalides [26]. BiOCl, BOI, BiOBr, and composites made from them have been widely reported due to their excellent photocatalytic properties [27-29]. However, the photocatalytic effectiveness of those semiconductors is inadequate for practical environmental and energy conservation applications because of substantial electron–hole recombination and a low capacity for the absorption of visible light. Numerous attempts have been made, with an emphasis on doping, the creation of heterojunctions, crystal plane management, and defect development, to enhance the photocatalytic efficacy of pristine Bi-based photocatalysts [1,30]. The photocatalytic processes of Bi-based photocatalytic applications have received little attention in environmental remediation and energy conversion. The rapidity with which this vital subject is advancing necessitates a thorough examination of recent breakthroughs regarding Bi-based photocatalysts. Consequently, in this work, Bi-based semiconductor photocatalysts and their manufacturing methods are discussed to make use of these photocatalysts in eco-friendly applications on a large scale. Bi-based nanomaterials as semiconductor photocatalysts are one of the study’s primary goals, as is the use of Bi-based nanomaterials for wastewater treatment, hydrogen generation, and photocatalytic degradation. Fabrication methods, reliability analogies, and future challenges of photocatalysts derived from bismuth-based nanomaterials are also discussed. There are many review reports on synthesis and enhancement techniques of Bi-based photocatalysts and the application of these photocatalysts in hydrogen generation, CO2 reduction, and water purification [31-35]. However, the present report focuses on understanding the role of different Bi-based photocatalysts concerning their synthesis method and enhancement. The mechanism of photocatalysts for different applications has been described for the type or nature of the photocatalyst. Hopefully, this work will help researchers understand how Bi-based nanomaterial photocatalysts may be employed in different environmental remediation systems by understanding the properties of Bi-based nanomaterials photocatalysts.

Review

Photocatalysis

Photocatalysis and its challenges

Photocatalysis can transform solar energy into storable chemical energy. Because of its minimal energy intake and carbon footprint, it is eco-friendly and promising. Two examples are the conversion of CO2 to hydrocarbons and water splitting to H2 and O2 [36,37]. Also, it is essential in domains including pollution degradation, antibiotic treatment, and sterilization [38]. The term “advanced oxidation processes” has become more common recently. In this process, many oxidizing agents (∙OH) are created. Electron–hole pairs are formed in AOPs when the VB electrons of semiconductor photocatalysts are driven into the conduction band (CB) through visible light [39]. The holes in the valence band of the catalyst split water to hydroxyl radicals (OH). Electrons in the CB of a semiconductor photocatalyst can generate the superoxide anion (O2) when they interact with oxygen molecules. During the photocatalytic oxidative degradation, the most notable oxidizing species are OH, photogenerated holes, and O2. These species are responsible for the photodegradation of organic and inorganic contaminants in wastewater [40]. To date, it is widely understood that the main limitation of photocatalysts is their low photocatalytic efficiency. The reason is that photogenerated electrons and holes recombine quickly [41]. We use an analogy for a more straightforward comprehension of the recombination timeframe. The force of gravity ensures that any item thrown into the air will return to the Earth below within a few seconds. After being exposed to light, electrons in a single photocatalyst undergo a transition akin to an item thrown into the air. This transition takes place from the VB to the CB [42]. After that, the very powerful Columbic force among photogenerated electrons and holes pulls them together, enabling recombination within a few picoseconds to nanoseconds in bulk or on the photocatalyst surface. It is possible to have a better understanding of the timescale by contrasting the calculation of the gravitational force with the computation of the Coulomb force. Because the gravitational constant (6.67 × 10−11 N·m2·kg−2) is significantly smaller than the Coulomb constant (8.99 × 109 N·m2·C−2), the recombination of the photogenerated electron and hole pairs takes significantly less time than the fall of the object. Within a single photocatalyst, photogenerated electrons and holes cannot withstand the tremendous force, which results in rapid recombination [43]. In addition to a low rate of recombination, other essential qualities of a superior photocatalyst include broad sunlight absorption and enough redox capacity. A small bandgap is desirable regarding a broad light absorption band. However, when redox ability is considered, the catalyst should have a high CB position and a low VB position, resulting in a large bandgap. These two prerequisites are not compatible. As a direct consequence of this, heterojunctions are produced. To create heterojunction photocatalysts, two semiconductor photocatalysts are combined [44,45]. Consequently, researchers focus their attention mainly on heterojunction photocatalysts.

Promising Bi-based nanomaterials

The overwhelming number of Bi-based semiconductors utilized in photocatalysis also have a distinctive layered structure and a bandgap of less than 3.0 eV. The connections between the layers are just van der Waals forces, which are weak [46]. Metal oxides such as TiO2 only have the O 2p orbital in their VB. In contrast , Bi-based oxide materials have an electronic structure in which O 2p and Bi 6s orbitals are paired in the VB. The bandgap of the semiconductor may be reduced to 3.0 eV thanks to the significant charge carrier dispersion provided by hybrid orbitals involving the Bi 6s orbital, as seen in Figure 2. Photoinduced electron–hole separation and charge carrier transfer in Bi-based materials are facilitated by a unique layered structure that creates an IEF. A magnetic field is generated between layers of Bi-based materials [47]. Many researchers have revealed that Bi-based nanomaterials have an adequate photocatalytic capacity for pollution remediation, water splitting, and the elimination of volatile organic compounds. Bi-based photocatalysts have substantial oxidative capabilities, as illustrated in Figure 2, since their VB potential is much higher than the oxidation potential of H2O, that is, 0.82 V vs NHE. Unfortunately, due to inadequate CB potential energy, most reduction processes, such as CO2 reduction, N2 fixation, and H2 creation, cannot be catalyzed with Bi-based nanomaterials. However, a few Bi-based semiconductor photocatalysts, such as Bi2S3, have a more substantial negative CB potential, making reduction reactions possible [34,48,49]. However, in real applications, the usefulness of Bi2S3 is limited because of the quick recombination of electrons and holes. New research has shown that it is possible to generate H2 or reduce CO2 by carefully controlling the production procedure of ultrathin photocatalysts based on Bi. Compared to Bi2WO6 formed using a solid-state reaction, single-unit-cell layers of Bi2WO6, synthesized using a hydrothermal process, offered three times greater CO2 adsorption and enhanced light absorption. The enhanced properties were reflected in the photocatalytic activity, resulting in a rate of methanol synthesis of 75 mol·g−1·h−1, which was much greater than the rate produced by the unaltered Bi2WO6. As a result of the decrease in dimensionality, it was projected that there would be a significant confinement effect and enlarged bandgap energy. Additionally, an upshift of the CB and a downshift of the VB were also measured [48,50]. The preparation process can be used to modify the electrical structure of Bi-based semiconductors. Most Bi-based semiconductors, especially oxide semiconductors such as Bi2O3, BiVO4, and Bi2WO6, have n-type properties with electrons as the majority carrier. Recent research suggests that the synthesis route can shift the conductivity types of Bi-based materials [45,51].

[2190-4286-13-109-2]

Figure 2: Bi-based photocatalysts exhibit substantial oxidative capabilities for various redox processes.

Fabrication routes

Synthesis approaches affect the topological size and size distribution of semiconductor photocatalysts, substantially influencing adsorption characteristics and photocatalytic efficacy. As well as affecting the environment, synthesis size, and cost, the fabrication method also affects manufacturing safety considerations [52]. Bi-based photocatalysts may currently be synthesized using various methods, the most common of which are hydrothermal synthesis, solid-state reaction, and design and optimization.

Today, NaBiO3·2H2O, Bi2O3, Bi(NO3)3·5H2O, BiCl3, and Bi are the most common Bi-based semiconductor photocatalysts. The hydrothermal process is one of the most frequent and widely utilized synthesis methods. Morphology, facets, size, surface flaws, and dimensionality of Bi-based photocatalysts are all susceptible to change when the synthesis conditions are altered. Generally, the hydrothermal technique yield a higher quality, and the nanoparticles that result from this process are more suited for specific applications than photocatalysts manufactured using dry methods [53]. However, the hydrothermal approach has a noticeable limitation in output yield. The usage of specific autoclaves has resulted in a long manufacturing time and the batch nature of production. For the hydrothermal technique, there is a danger of nanoparticle leakage, primarily in water, and the risk of toxic solvent emissions. Su et al. [54] reported the first hydrothermal fabrication of Bi5O7Br. Bi5O7Br rods of 50 μm length and 2 μm width were obtained. They found that Bi5O7Br effectively converts molecular oxygen to superoxide radicals and hydroxyl radicals in visible light. Under UV–vis irradiation, Bi5O7Br showed a higher photocatalytic activity in the degradation of rhodamine B (RhB) dye than BiOBr. The addition of Bi5O7Br photocatalysis to the Bi–O–X photocatalytic system improved the system. In this work, they found that the RhB elimination percentage over Bi5O7Br is 85% after 120 min of UV–visible-light irradiation, and the reaction rate constant was measured as 1.496 h−1·m−2. In contrast, the reaction rate constant for BiOBr was found to be only 0.154 h−1·m−2. Additionally, the hydrothermally fabricated Bi5O7Br has also excellent stability when exposed to light in aqueous environments. The photocatalytic performance was not much decreased after eight cycles. Recently, Lin et al. fabricated a monoclinic BiVO4 photocatalyst via a surfactant-free hydrothermal technique [55]. According to the results, the pH value has a significant impact on shape, surface area, particle size, and V–O bond length. The grain size was reduced when the pH value was raised, and the crystal structure became more closely stacked. Under visible-light irradiation, the RhB photodegradation efficiency of the coralloid particles produced at pH 7 was about four times greater than that of the sample synthesized at pH 0.5. The increased photocatalytic activity was caused by several factors, such as a synergistic effect of highly exposed (010) grain facets, an increase in the overlap between Bi 6s and O 2p orbitals, a reduced charge transfer route, and strange floating features. In another approach, Huang et al. reported a one-step hydrothermal synthesis of a BiIO4/Bi2MoO6 hybrid photocatalyst for photodegradation of RhB dye [56]. Under visible-light irradiation (λ > 420 nm), photoelectrochemical studies demonstrated that the RhB photodegradation effectiveness and photocurrent density of the BiIO4/Bi2MoO6 hybrid composite is much higher than that of the pure components. Because of the successful construction of the BiIO4/Bi2MoO6 hybrid, the photocatalytic activity was found to be significantly increased. This improvement was ascribed to the efficient interfacial charge transfer obtained as a consequence of the process. According to the active species trapping study, the photocatalysis process is significantly aided by the presence of photogenerated holes.

It has been shown that the hydrothermal method of fabricating Bi-based semiconductor photocatalysts offers some significant benefits. Nanostructured materials have several advantages, including a more adaptable area for more detailed reconstruction, confinement effects, superior mechanical stability, and large surface area, which make them excellent for photocatalysis. The progress of synthesis processes allows one to change the physical properties as needed.

In order to create economically viable Bi-based photocatalysts, substantial amounts of water are required. Solid-state reaction methods that do not require water are suitable for large-scale synthesis. However, in solid-state reaction methods, there is the risk of a release of nanoparticles into the air; as a result, they are not entirely eco-friendly [57]. He et al. reported the fabrication of a Bi4NbO8Br photocatalyst via solid-state synthesis [58]. They found that the temperature used during calcination plays a vital role. A higher calcination temperature results in a more crystalline morphology with more active sites for photocatalytic activity. The fabricated specimens were utilized for the photodegradation of RhB under visible light. The sample calcinated at 750 °C revealed the highest photocatalytic performance. Hamza et al. fabricated Bi2(CrO4)3 nanoparticles via a facile precipitation technique [59]. The photocatalytic activity of the Bi2(CrO4)3 nanoparticles was studied under UV, AM 1.5, and visible-light irradiation, and acceptable rates of 522.44, 174.15, and 88.24 μmol·g−1·h−1, respectively, were reached under these conditions. These rates outperform those of other similar Bi-based semiconductor photocatalysts reported in the literature.

To synthesize Bi-based photocatalysts with high anisotropy, hollow structures, or crystalline multidimensional forms, the template technique is one of several synthesis approaches. Direct manufacturing processes make this almost impossible. Therefore, using templates is a great workaround [60]. Template methods may be characterized either as a “hard templates”, “soft templates”, or “self-templates”. The high cost of template methods results from the lengthy process of creating and removing templates. Also, environmental aspects need to be considered as removing templates such as SiO2 requires using very corrosive acids or bases [61]. Numerous studies have focused on choosing templates for making nanostructures of functional materials. However, templates were chosen rather based on utility than on cost. The self-template technology does not require additional templates, which leads to less expensive production and increased efficiency. This makes the approach more realistic for practical applications [62]. A very good micro-/nanoscale hierarchical Bi7O9I3/NTC photocatalyst was created in a one-step, easy, and environmentally friendly way by Hou et al., who used an in situ ion exchange–recrystallization approach [63]. The used buffer provided a relatively stable environment for producing regular structures. The aqueous NH3 solution provided OH ions for the successful exchange of I, and the result was the synthesis of Bi7O9I3. The Bi7O9I3/NTC has evenly distributed Bi7O9I3 nanostructures in the shape of lanterns formed of extremely thin nanosheets with a thickness of less than 10 nm on both the surface and the inside. When exposed to visible light, Bi7O9I3/NTC displayed higher photocatalytic activity owing to the synergistic effect of the micro-/nanoscale hierarchical structure, low iodine content, and well-contacted interface. 93.5% methyl orange (MO) and 96.6% RhB were eliminated from solution during in two hours, suggesting a greater photocatalytic effectiveness than that of pure BiOI. To deposit metallic Bi on Bi2WO6 nanosheets, an in situ reduction approach using NaBH4 as the reducing agent was used [64]. Compared to pure Bi2WO6, Bi-coated Bi2WO6 absorbs more visible light, is more sensitive to photocurrent, and has a lower electrochemical impedance rate. This is because of surface plasmon resonances (SPRs) and the electron transport capabilities of Bi. The photocatalytic activity for the breakdown of phenol was significantly improved, compared to pristine Bi2WO6 under visible light. Xiao and colleagues have shown a straightforward synthesis approach for fabricating Bi2WO6 nanosheet rods [65]. They discovered that the hydrolysis of the precursor Bi(NO3)3 may quickly result in the formation of Bi6O5(OH)3(NO3)5·3H2O nanorods, which then acted as templates for the generation of Bi2WO6. It has been observed that the newly generated Bi2WO6 has a greater BET surface area and superior charge transfer kinetics. These properties point to an increase in photocatalytic activity. Other Bi-based hollow hierarchical structures, such as BiVO4, have the potential to be synthesized and used as CO2 reduction photocatalysts. This possibility exists since these structures are hollow. It is anticipated that hierarchical Bi-based photocatalysts produced would have a broad range of applications in environmental science and energy research.

Different preparation methods have been employed for the synthesis of Bi-based photocatalysts. Each method has unique advantages and disadvantages, which are compared and contrasted in Table 1. New strategies are still required to create suitable nanomaterials by overcoming faults and enhancing the synthesis process. The hydrothermal approach is the one that is most often used to produce nanomaterials. This is because it is easy to implement and allows for a complete control over the shape and size of the nanoparticles.

Table 1: Various fabrication techniques, their advantages, and limitations.

No. Technique Advantages Limits Ref.
1 solid-state technique high crystallinity, easy operation phase transition, high temperature, big grain particle, tiny surface area [66]
2 chemical precipitation convenient synthesis, low cost, and low energy consumption aggregation of particles, limited surface area, development of impurities, uncontrolled morphology [67]
3 sol–gel technique controllable morphology, nanoparticles, tiny and fine particles organic residues, high costs, and treatment procedure [68]
4 hydrolysis different particles size, easy synthesis conditions, and simple equipment uncontrollable morphology, solvent-dependent, poor dispersion [69]
5 hydrothermal simple operation, high crystallinity, a variety of morphologies, and particle size control high pressure, different reaction parameter [70]
6 drop casting method convenient precipitation, easy operation time-consuming, unpredictable thickness and homogeneity, and poor adhesion [71]
7 dip/spin-coating simple operation, thickness control, homogenous film inadequate attachment, specialized equipment [72]
8 spray pyrolysis quick and easy operation, cost-effective, scalable technique, controlled thickness, complex composites production high temperature, high cost, unwanted precipitates production, and high resistance [66]
9 hydrothermal coating easy synthesis, good conductivity, long durability, and controllable morphology high pressure, poor yield, complex parameters, and poor dispersion [68]
10 chemical vapor deposition regular thickness, minimal porosity, diverse materials, and a high degree of crystallinity extreme temperatures and specialized equipment [66]
11 anodization technique large-scale synthesis, large surface area, and morphological control applied bias, complex synthesis parameters [73]
12 electrodeposition technique thickness control, homogenous film specific equipment, imposed bias, and treatment procedure [74]

Enhancement techniques

The characteristics of a semiconductor photocatalyst alter as its size is reduced to the nanoscale. Increasing the proportion of atoms or ions exposed on the photocatalyst surface will increase the number of photocatalytically active sites [75]. Under light irradiation, the average amount of time it takes for a photogenerated carrier to diffuse from the bulk to the surface may be calculated as follows: τ = r22D, where r is the grain radius and D is the carrier’s diffusion coefficient. Consequently, as the particle radius decreases, a higher number of photogenerated carriers can reach the surface, where they might participate in a photocatalytic process [76]. Bismuth is often used as a nanoscale plasmonic photocatalyst. Nanospheres, nanorods, and nanosheets can be synthesized using various techniques. Hydrothermal calcination, template synthesis, precipitation, reverse micro-emulsion, sonochemical procedures, and microwave methods are typical techniques for fabricating Bi-based nanostructures [77]. Recombination of charge carriers and insufficient photon absorption are the two most common problems related to semiconductor photocatalysts. Also, it is necessary to have a wide bandgap to yield an adequate redox ability. However, a moderate bandgap is required to perform photocatalysis using visible light. This gap is essential for improving the material’s light-harvesting capabilities [78,79]. Consequently, in an attempt to improve the photocatalytic efficiency for water purification and other environmental applications, a variety of techniques, such as defect formation, metal/non-metal doping, heterostructure formation, interface modification, and Bi-content enhancement, have been employed.

Defect formation

Vacancies and defects affect the electrical properties of Bi-based semiconductor photocatalysts and, hence, govern the photocatalytic efficacy. Rao et al. reported an N2-assisted heat treatment approach for the in situ synthesis of a series of oxygen-vacancy (OV)-rich Bi0/Bi-based photocatalysts [80]. A new understanding of how Bi0 nanoparticles and OVs are created in situ in Bi-based photocatalysts has been reported. Compared to other Bi-based photocatalysts, Bi0/OV–(BiO)2CO3 showed high photocatalytic performance and stability for the photooxidative elimination of NO. The ohmic interaction between OV–(BiO)2CO3 and Bi0 has been shown to promote the synthesis of O2 and OH species. It was found that O2 had a significant impact on the photocatalytic elimination of NO.

In another approach, Huang et al. reported that BiOI microspheres served as self-sacrificing templates for in situ phase transformation and formation of phase junctions [81]. Different bismuth oxyiodides were formed as a result of this. Hierarchical BiOI, Bi4O5I2, Bi4O5I2–Bi5O7I phase-junction, and Bi5O7I may be synthesized from bismuth oxyiodides at different temperatures (Figure 3a). The photoabsorption wavelength of these bismuth oxyiodides has been tuned between 400 and 700 nm. Also, these compounds have a distinctive microstructure and a controllable band structure. (Figure 3b). The breakdown of antibiotics and pollutants such as tetracycline hydrochloride, bisphenol A (BPA), and RhB was used to measure the photocatalytic activity of the bismuth oxyiodides. The activity decreased in the sequence Bi4O5I2–Bi5O7I > Bi4O5I2 > BiOI, which is linked to charge separation efficiency and band structure. Engineered Bi vacancies in monolayered Bi2WO6 nanosheets with a thickness of 1.0 nm have recently been shown [82]. The Bi defects were shown to promote the adsorption and activation of reactant molecules, which reduced the energy barrier even more. The photocatalytic performance corroborated this. The presence of divacancies may help increase charge carrier separation by capturing photogenerated electrons close to the divacancies. In comparison, pure Bi2WO6 nanosheets had a photocatalytic performance 32 times lower for the elimination of gaseous toluene when exposed to visible light.

[2190-4286-13-109-3]

Figure 3: (a) The hierarchical structures of BiOI, Bi4O5I2, Bi4O5I2–Bi5O7I composite, and Bi5O7I are shown in this diagram. A schematic description of the charge segregation and transfer process for (b) the single photocatalyst of BiOI, Bi4O5I2, and Bi5O7I, and (c) Bi4O5I2–Bi5O7I phase junction. Figure 3a–c were reprinted from [81]. This article was published in Applied Catalysis B: Environmental, vol. 203, by H. Huang; K. Xiao; T. Zhang; F. Dong; Y. Zhang, “Rational design on 3D hierarchical bismuth oxyiodides via in situ self-template phase transformation and phase-junction construction for optimizing photocatalysis against diverse contaminants”, pages 879–888, Copyright Elsevier (2016). This content is not subject to CC BY 4.0.

Recently, Zou et al. proposed a simple hydrothermal synthesis for preparing 2D BiOCl nanosheets [83]. This was accomplished by altering the pH value of the precursor solution and using of dulcitol (C6H14O6) as surfactant. The pH value substantially influenced the thickness of the nanosheets and the fraction of exposed (001) facets. The sample synthesized at pH 4 demonstrated outstanding visible-light photocatalytic performance regarding the degradation of RhB. This was due to the low thickness, exposed (001) facets, and an appropriate number of oxygen vacancies. This work proposed that revolutionary photoexcitation mechanisms were found on oxygen vacancies. Irradiation with visible light excites the electrons in the VB to transition into defect states. In addition, photogenerated defect states cannot readily recombine with photogenerated holes because oxygen vacancies operate as electron traps. Because of this, electrons trapped inside the oxygen vacancies have a longer lifetime than those in the CB. Therefore, electrons in defect states have the potential to react with oxygen that has been adsorbed by oxygen vacancies, which results in the production of superoxide O2 radicals. These may subsequently be employed to drive photocatalytic processes. The absence of a metal atom would considerably affect the amount of absorbed light, the rate of charge transfer, and the number of reactive surface sites.

Metal/non-metal doping

Photocatalysts based on bivalent cations may be enhanced by doping them with additional elements. Self-doping and deposition and doping of metals and non-metallic elements are the most common doping methods. Metal ions modify the crystal structure of the Bi-based semiconductor photocatalysts or induce defects. Also, the photocatalytic properties may be altered by doping or deposition of metallic components [84,85]. Using a straightforward hydrothermal procedure, Hu et al. produced iron-doped Bi2WO6 nanocomposites [84].

In comparison to pristine Bi2WO6, Fe-doped Bi2WO6 exhibits superior visible-light photoabsorption, a considerably increased number of oxygen vacancies, and a noticeably improved capacity for separating photogenerated electrons and holes. As a result of the Fe doping, an impurity energy level was produced close to the VB, and an imperfection (oxygen vacancy) energy level was produced close to the CB. Both of these energy levels are in a position to potentially accept photoinduced holes and electrons, which significantly improves the electron–hole pair splitting. When Fe3+ substitutes W6+, the structure of the crystal is not compromised in any way; nevertheless, numerous additional oxygen vacancies are produced. The increased electron–hole pair separation is the driving factor behind the improved photocatalytic activity of the Fe-doped Bi2WO6 compound. When exposed to visible light, Fe-doped Bi2WO6 exhibited photocatalytic degradation rates that were 11.9 and 8.0 times higher than those of pristine Bi2WO6. This material was also found to be superior to the majority of modified Bi2WO6 photocatalysts that had been reported in the past. In addition to this, Fe-doped Bi2WO6 has a high degree of stability. The results of this study provide new information on boosting the photoactivity of Fe-doped Bi2WO6.

The SPR effect can be obtained by the deposition of metallic elements on a semiconductor surface. SPRs can potentially boost quantum yield by broadening the spectral response range of semiconductors. Fe, Au, Co, Ag, Ni, Bi, Al, and other metallic elements are often deposited and doped. For example, a nanostructure composite based on plasmonic Ag metal nanoclusters and monoclinic BiVO4 nanoparticles was fabricated using high-energy ball milling [86]. Ag clusters (5–10 nm) were homogeneously distributed on the flocculated BiVO4 particles (50–100 nm). The structure of the Ag-doped BiVO4 nanocomposite would promote the efficiency of the photodegradation of acid blue dyes. Huang and co-workers observed that a Bi-Bi2WO6 composite successfully degraded RhB and 4-chlorophenol under visible light [87]. They reported that the formed heterojunction yielded a two times higher RhB photodegradation and a three times higher 4-chlorophenol photodegradation than bare Bi2WO6.

Using nonmetal doping, doping energy levels may be created between CB and VB of Bi-based photocatalysts. This can increase light absorption, and charge transfer may be improved to enhance electron–hole segregation and reduce recombination. It is not uncommon to see nonmetals atoms replaced with heteroatoms, such as N (C), B (S), X (F, Cl), Br (I). Dong et al. [88] reported the fabrication of boron-doped Bi3O4Cl ultrathin nanosheets via a solvothermal technique, which were found to have enhanced solar absorption and efficient electron–hole separation. The B atoms enhance the photocatalytic performance via (1) producing mid-gap states to widen the light response region significantly up to 557 nm and (2) functioning as electron capture centers to accelerate charge carrier separation. According to ESR measurements, B-doped Bi3O4Cl can create more O2 and OH radicals. Consequently, the B-doped sample has a 3-fold and 2.1-fold better degradation efficiency for, respectively, BPA and ciprofloxacin than pristine Bi3O4Cl. This study offers fresh perspectives on photocatalyst design and underlines the importance of electronic structure modification in catalytic activity adjustment.

Self-doping is a novel approach for introducing intermediates from the synthesis process into photocatalysts to alter the energy band structure and increase photocatalytic activity [89]. A simple two-step technique was used to develop a novel compound photocatalyst of Bi/BiOBr-Bi5+ [90]. X-ray diffraction, field-emission transmission electron microscopy, and X-ray photoelectron spectroscopy revealed the coexistence of self-doped Bi5+ and in situ deposited Bi(M). Compared to Bi/BiOBr or BiOBr–Bi5+, the photocatalytic activity of Bi/BiOBr-Bi5+ regarding RhB degradation under visible light was significantly increased. OVs helped in separating photoexcited carriers, and SPRs enhanced the ability to absorb visible light. The photocatalytic activity of Bi(M) was further enhanced by the exposure of (010) facets (Figure 4a). Bi5+ reduced the bandgap of BiOBr, which led to an increase in the density of carriers. Examining electron transfer channels and identifying the most active species led to a credible mechanism for RhB degradation (Figure 4b,c).

[2190-4286-13-109-4]

Figure 4: (a) Crystal structure of self-doped Bi/BiOBr–Bi5+, (b) band structure, and (c) RhB degradation of Bi/BiOBr–Bi5+ in visible light. Figure 4a–c were adapted from [90]. This article was published in Separation and Purification Technology, vol. 253, by Q. Wu; S. Chai; H. Yang; Z. Gao; R. Zhang; L. Wang; L. Kang, “Enhancing visible-light driven photocatalytic performance of BiOBr by self-doping and in situ deposition strategy: A synergistic effect between Bi5+ and metallic Bi”, article no. 117388, Copyright Elsevier (2020). This content is not subject to CC BY 4.0.

It has been shown that the carrier combination centers may easily get doped at a deep level, considerably lowering photocatalytic activity. As a result, optimizing the electrical configuration by using suitable dopants and concentrations will increase photocatalytic activity. Photogenerated electron–hole separation efficiency and light absorption ability are critical functions that can be improved by appropriate doping.

Heterojunction formation

In general, for a photocatalyst to function appropriately, it is necessary to use a semiconductor with the following characteristics: a bandgap suitable for light harvesting, effective charge carrier separation capabilities, and suitable VB and CB edge potentials [91,92]. It is challenging to meet these requirements with only one single Bi-based photocatalyst. Constructing semi-conductor heterojunctions may be an effective technique for addressing the difficulties of individual Bi-based photocatalysts. This may be due to the changeable band structure and effective photoinduced electron–hole separation, which bestows them with higher capabilities [93]. It is possible to improve absorption of visible light, charge carrier segregation, and charge transport effectiveness by combining heterojunction and nanomaterials in photocatalysts constructed with care. As a result, this technique offers much potential for photocatalysis applications [94]. The band arrangement shown in Figure 5a may be used to classify the heterojunctions between semiconductors as either a straddling gap (type I), an uneven gap (type II), or a broken gap (type III). Type-II heterojunctions have garnered the most interest because of the improved photogenerated electron–hole separation they offer. They include n–n heterojunctions, p–n heterojunctions, p–p heterojunctions, and Z-scheme-based heterojunctions (Figure 5b). By using heterojunctions, it is possible to exert control over the electronic components of the photocatalyst to increase light absorption and photoinduced separation and migration [95,96]. Photoinduced holes in n-type semiconductors are transported to p-type semiconductors by an electric field at the interface, whereas photoinduced electrons from p-type semiconductors are transported to n-type semiconductors (Figure 5b).

[2190-4286-13-109-5]

Figure 5: (a) Band diagrams representing three different semiconductor heterojunctions, (b) band diagrams for p–n, n–n (p–p), and Z-scheme-based heterostructures.

Using simple and cost-effective experimental conditions, Sang et al. reported the fabrication of nanoflower-like Bi2O3/Bi2S3 heterojunctions via a one-step hydrothermal technique [97]. In the photocatalytic elimination of RhB and Cr(VI) under visible-light irradiation, the photocatalytic activity of this Bi2O3/Bi2S3 heterojunctions outperforms that of pristine Bi2O3 and Bi2S3. They found that photoinduced holes were the main oxidative species for eliminating RhB, whereas photogenerated electrons were responsible for the photoreduction of Cr(VI). Typically, the Fermi energy level of n-type Bi2O3 is quite close to the VB, while the Fermi energy level of p-type Bi2S3 is somewhat close to the CB. In a heterojunction from n-type Bi2O3 and p-type Bi2S3, electrons flow from n-type Bi2O3 to p-type Bi2S3, and holes flow from p-type Bi2S3 with a low Fermi level to n-type Bi2O3 with a high Fermi level. As a consequence of this, negative charges build up in Bi2S3 near the interface, which leads to an electric field. Also, the Fermi level of Bi2O3 goes down, while the level of Bi2S3 goes up. Along the Fermi level, the energy bands of both Bi2O3 and Bi2S3 are moving simultaneously in a downward and an upward direction. An equilibrium state, in which the Fermi levels of Bi2O3 and Bi2S3 equilibrate at the p–n junction, was also proposed. Bi2S3, which has a smaller bandgap than Bi2O3, is excited when it is exposed to visible light, while Bi2O3 is not. The photogenerated holes stay in the p-type Bi2S3 VB, while the excitons in the p-type Bi2S3 CB migrate to the n-type Bi2O3 CB. In the Bi2O3/Bi2S3 photocatalytic system, electrons and holes participate directly in the redox process. In another reported work, a BiOI/Bi2O2CO3/graphene ternary composite was fabricated via a facile and economic hydrothermal technique [98]. In order to assess the newly constructed semiconductor heterojunction, tetracycline and RhB were degraded using visible light. Better photocatalytic activity can be achieved with BiOI/Bi2O2CO3 photocatalysts instead of just using BiOI alone. This is because in BiOI/Bi2O2CO3 photocatalysts, an electric field emerges at the p–n heterojunction, which in turn helps to foster the movement of photogenerated carriers. Furthermore, the high photocatalytic activity of the BiOI/Bi2O2CO3/RGO composite can be attributed to the fact that the positively charged BiOI/Bi2O2CO3 was electrostatically paired with the negatively charged graphite oxide (GO) to form interlayer contacts. This caused the photocatalytic reaction sites to boost, the light response to broaden, and the separation of photoinduced charge to improve. Lv et al. [99] fabricated a p–n heterojunction-based novel CuS/Bi2WO6 semiconductor photocatalyst with 2D interfacial connections of CuS over the surface of Bi2WO6. The hydrothermal method was used, and it was discovered that the produced CuS/Bi2WO6 semiconductor photocatalyst had increased photocatalytic performance for the breakdown of glyphosate when exposed to visible light. Enhanced photocatalytic activity, excellent recyclability, high stability of CuS/Bi2WO6 photocatalysts may be primarily due to the presence of an electrical potential at the interface, which is responsible for both the enhanced visible light absorption and the efficient segregation of photoinduced charges. This study introduced a unique 2D interfacial coupling for the production and design of efficient photocatalysts. It also demonstrated that CuS may be used as an effective semiconductor in heterostructures, which is a concept that may be extended to other functional nanomaterials based on bismuth.

The most important heterojunctions are known as Z-scheme heterojunctions and have a band energy structure comparable to type-II heterojunctions. However, there are many alternative pathways open for charge carrier movement (Figure 6), including mediator-free or direct (Figure 6a), solid mediator (Figure 6b), and redox pair mediator (Figure 6c) [43,100]. A “mediator” is often employed to offer an intermediary conduit for electrons to flow from the CB of semiconductor II (SC II) to the VB of semiconductor I (SC I), making the charge transfer easier [38]. A solid substance or a redox couple in solution may serve as a mediator in the Z-scheme. This dual absorber system has the potential to get excited, which would result in the generation of photogenerated holes and electrons in CB and VB, respectively. Photogenerated electrons in the CB of SC II may migrate to the VB of SC I, where they may merge with photogenerated holes. This design is favorable for the electrons in the CB of SC I and the holes in the VB of SC II to keep their optimum and original redox potentials (Figure 6a). This method of electron transmission resembles the letter Z. As a result, the structure of this heterojunction is referred to as the direct Z-scheme [43,101]. Fu et al. [102] demonstrated the synthesis of a new direct Z-scheme photocatalyst made of ultrathin Bi2O3 and Bi2MoO6 microspheres. For the effective production of Bi2O3/Bi2MoO6 nanocomposites, researchers adopted a simple in situ alkali treatment of Bi2MoO6 followed by calcination. As a substrate for the production of Bi2O3 sheets, Bi2MoO6 microspheres were used. The 2D morphological properties of the Bi2O3 sheets resulted in enhanced charge carrier transfer. The relative mass ratio of Bi2MoO6 and Bi2O3 may be fine-tuned by adjusting the alkali dose (i.e., NaOH or KOH). Using phenol degradation and hydrogen generation as a test, the 1.5% Bi2O3/Bi2MoO6 sample was shown to be the most photocatalytically active. An easy hierarchical Z-scheme system with a ZnIn2S4/BiVO4 heterojunction has been proposed by Hu and co-workers [103]. This system can precisely regulate redox centers at the ZnIn2S4/BiVO4 heterojunction by expediting the detachment and mobility of photoinduced charges. This, in turn, increases the ability of holes and electrons to undergo oxidation and reduction, respectively. As a consequence of this, the ZnIn2S4/BiVO4 heterojunction has unusual photocatalytic activity. It has an H2 evolution rate much higher than pure ZnIn2S4, with a value of 5.944 mol·g−1·h−1. This value is over five times higher. In addition, this heterojunction has excellent stability and the capacity to be recycled, making it a promising photocatalyst for the formation of H2. Ternary composite heterojunction photocatalysts have been also reported. The Fenelon group successfully fabricated SnO2 and Bi2S3–Bi25 composites via a facile hydrothermal technique followed by thermal breakdown [27]. The photocatalytic performance of the fabricated ternary composite photocatalyst was 2.75 times higher than that of pristine Bi2S3 for the photodegradation of RhB under visible light. Their research also revealed that 15% SnO2 precursor solution was the most effective concentration for achieving a photocatalytic degradation efficiency of 80% after 180 min of exposure to visible light. Photogenerated holes were found to be responsible for the oxidation and breakdown of the pollutant during the photocatalytic reaction.

[2190-4286-13-109-6]

Figure 6: Different charge transfer Z-scheme types: (a) Mediator-free or direct, (b) solid mediator, and (c) redox pair mediator. Abbreviation lists: semiconductor I (SC I), semiconductor II (SC II), valence band (VB), and conduction band (CB).

One of the most popular ways to create an effective heterojunction structure is by combining two semiconductors with matching band alignment. When it comes to understanding the mechanisms behind the enhanced photocatalytic performance of heterojunction photocatalysts, researchers have recently proposed an S-scheme heterojunction [45]. Separating photoinduced electrons and holes with an S-scheme heterojunction efficiently preserves the promising redox properties of semiconductors. It is common for electrons to flow from the CB of one semiconductor to the VB of the other in an S-scheme heterojunction photocatalyst due to an IEF that typically exists at the interface of the two semiconductors [104]. Lately, Xu et al. reported the fabrication of a MoS2/BiVO4 heterojunction via solvothermal and electrospinning techniques [105]. Without any additional agent, the fabricated heterojunction completely degraded a RhB dye solution within 20 min. BiVO4 nanorods photogenerated hydroxyl radicals rather than super oxides because of the more positive oxidation potential of BiVO4 (2.31 V). MoS2 sheets favored the photogeneration of superoxide radicals because of the more negative CB position (−1.39 V). This finding demonstrated the spatial distribution of oxidation sites (BiVO4) and reduction sites (MoS2) via an S-scheme charge transfer path and significantly aided in the inactivation of bacteria under illumination. In another work, a simple hydrothermal technique was applied by Liu et al. to modify BiVO4 heterojunctions with carbon quantum dots [106]. Exceptional photocatalytic performance for the degradation of RhB dye under visible light was exhibited by CNQDs-ms/tz-BiVO4 and NCQDs-ms/tz-BiVO4 composites. The reason for their increased photocatalytic efficiency was the formation of heterojunctions together with the loading of quantum dots, which enhanced the light-harvesting efficiency and promoted the separation and migration of photogenerated carriers. The fabricated composites followed the S-scheme charge transfer mechanism, efficiently contributing to enhanced photocatalytic performance. Another research group reported that photoreduction and hydrothermal techniques were used to successfully synthesize a new 2D/2D Bi2MoO6/g-C3N4 S-scheme composite including Au as a co-catalyzer [107]. Bi2MoO6/g-C3N4/Au had a photocatalytic activity in RhB degradation that was 9.7 times and 13.1 times higher than that of Bi2MoO6 and g-C3N4, respectively. In the Bi2MoO6/g-C3N4/Au system, the higher photocatalytic activity can be attributed to the abundance of active sites and the enhanced separation efficiency of photogenerated carriers. The potential role of Au nanoparticles in the S-scheme heterostructure is noteworthy. They serve as a co-catalyst for improving electron separation and transmission due to the photogenerated potential.

By forming heterojunctions, the visible-light absorption as well as the carrier separation efficiency of Bi-based nanophotocatalysts can be improved, which in turn increases the photocatalytic activity. The zonal organization of several crystalline surfaces made from the same material may significantly increase charge separation. It can be used to create classic heterojunctions in addition to surface heterojunctions, which can be created using the same method. Surface heterojunctions have recently emerged as a novel concept that has garnered much interest. As a direct consequence of this, it is possible that in the not-too-distant future, it will be possible to construct and analyze further Bi-based surface heterojunction photocatalysts. Table 2 below provides an in-depth explanation of several different strategies that may be used to improve Bi-based nanocomposites.

Environmental applications

The use of Bi-based photocatalysts in power production and environmental remediation is widespread. Sewage treatment, environmental monitoring, disinfection, and sterilization are all areas where the photocatalytic breakdown of contaminants is used. Primary energy uses included photocatalytic hydrogen production from carbon dioxide, conversion of carbon dioxide to specific molecular organic matter, and nitrogen fixation [1,108,109]. Photocatalysts with Bi-based photocatalysts are discussed in this section in more depth.

Photocatalytic hydrogen generation

At the moment, humanity’s existence greatly depends on the consumption finite fossil fuels. Thus, finding long-term renewable energy sources is critical. Hydrogen is produced primarily by the electrolysis of water, in parts using solar energy, and the reformation of fossil fuels [110]. The conversion of water into hydrogen by using solar energy is considered the best way to produce hydrogen [111]. Photocatalysts still face the following issues regarding efficient water splitting. First, the quantum efficiency deteriorates as one moves into the visible region. Second, photocatalysts such as sulfides and nitrogen oxides in the visible region have poor stability and are frequently inactivated by photocorrosion. Third, the removal of O2 from the surface of a semiconductor photocatalyst is complex, which is why a h+-capturing substance is frequently added to the system to promote the production of H2 [112]. Pt, Pd, Au, and other precious metals, as well as photocatalysts containing Cd, Pb, and other elements (such as CdS), are some of the costliest and environmentally hazardous high-performance photocatalysts on the market. CdS is a more ecologically friendly alternative to other photocatalysts [113]. To improve the efficiency of photocatalytic water breakdown from the perspective of catalytic reaction kinetics, one method that is both practical and effective is to design photocatalysts in such a way as to change the processes involved in photocatalytic reaction kinetics.

For example, Cao et al. reported the fabrication of a bismuth-based Bi/Bi2MoO6/TiO2 nanocomposite photocatalytic material using a facile one-step solvothermal technique [114]. To reduce metal Bi on the surface of Bi2MoO6, they employed glucose as a reducing agent. To test the photocatalytic activity, they exposed the as-prepared materials to simulated solar irradiation and measured the degradation efficiency of RhB (7.321%), MB (92.98%), and Cr(VI) (70.54%). The highest hydrogen generation rate was 173.41 mol·h−1·cm−2, and after four cycles with the same parameters, there was no noticeable decrease in the hydrogen production of the Bi/Bi2MoO6/TiO2 system. The development of Bi2MoO6/TiO2 heterojunctions, the SPR effect of Bi, and synergistic effect may be responsible for the enhanced photocatalytic activity and hydrogen generation rate (Figure 7a). Developing semiconductor photocatalysts that are both low-cost and highly efficient is essential for the practical conversion of solar energy into fuel. A Ag–C3N4-modified (BiO)2CO3 semiconductor photocatalyst was synthesized using an in situ thermal approach in [115]. The activity of the (BiO)2CO3/g-C3N4 heterojunction reached 965 μmol·h−1·cm−2 after 5 h. This is almost three times the activity of pristine g-C3N4 or any other modified g-C3N4 nanomaterial (337 μmol·h−1·cm−2). An increase in photocatalytic activity has been found due to the usage of a direct Z-scheme system (Figure 7b). Theoretical simulations showed that charge carriers were redistributed at the junction between (BiO)2CO3 and g-C3N4.

[2190-4286-13-109-7]

Figure 7: Graphical representation of the proposed charge transfer mechanism of (a) the Bi/Bi2MoO6/TiO2 nanocomposite and (b) the (BiO)2CO3/g-C3N4 heterojunction. Figure 7a was reprinted from [114]. This article was published in Separation and Purification Technology, vol. 250, by D. Cao; Q. Wang; Y. Wu; S. Zhu; Y. Jia; R. Wang, “Solvothermal synthesis and enhanced photocatalytic hydrogen production of Bi/Bi2MoO6 co-sensitized TiO2 nanotube arrays”, article no. 117132, Copyright Elsevier (2020). This content is not subject to CC BY 4.0. Figure 7b was reprinted with permission from [115], Copyright 2019 American Chemical Society. This content is not subject to CC BY 4.0.

Photocatalytic elimination of water pollutants

A wide variety of contaminants in water, such as heavy metal ions, pharmaceuticals, and pesticides, have been used as models substances for photocatalytic degradation. Reactive oxygen species (ROSs) such as O2 and OH are critical to the photocatalytic removal of environmental pollutants.

Malathi et al. reported the photocatalytic and photoelectrochemical performance of BiFeWO6/BiOI nanohybrids, which were fabricated by wet impregnation, under visible-light irradiation [116]. Within 90 min, the composite degraded 92% of RhB. Similarly, Priya et al. used a simple wet-chemical method to make BiFeWO6/WO3 nanocomposites [117]. A 250 W tungsten halogen lamp was used to carry out the photoexcitation of the catalysts for RhB degradation. After a period of 60 min, the composite yielded a 83% breakdown of RhB. The multiferroic single-phase material known as nanoscale zero-valent iron-doped bismuth ferrite (NZVI@BiFeO3) displays concurrent ferroelectric, ferromagnetic, and ferroelastic properties. BiFeO3 has the ability to be utilized in heterogeneous photocatalysis when exposed to light from the sun. The NZVI@BiFeO3/g-C3N4 nanocomposite was produced using a straightforward hydrothermal process [118]. The composite was employed for the photodegradation of RhB. The generated composite could degrade 97% of the RhB dye when exposed to solar light. Compared to pure BiFeO3, the NZVI@BiFeO3/g-C3N4 composite has a better optical responsiveness, which contributes to the improved photocatalysis of the material. According to the theorized dye degradation mechanism, the composite provides efficient reactive species trapping sites. Because of this, the NZVI@BiFeO3/g-C3N4 nanocomposite has the potential to be utilized to oxidize a wide range of organic and inorganic contaminants that are found in wastewater.

Recently, Chava et al. proposed the solvothermal method for preparing 1D/0D CdS/Bi heterostructures [119]. To evaluate their potential photocatalytic performance toward photodegradation of the organic pollutants, tetracycline was used as model substance. The 1D/0D CdS/Bi heterostructures exhibited excellent catalytic activity, compared to pure CdS. It was expected that 1D/0D CdS/Bi could act as an effective photocatalyst for the treatment of other organic pollutants in wastewater, such as organic dyes and pharmaceuticals. The enhanced photocatalytic degradation efficiency may be due to doped Bi3+ species, the SPR effect in the metallic Bi quantum dots, and increased photoinduced charge carrier separation. In another study, a visible-light-sensitive heterostructure made of InVO4/Bi2WO6 nanoflakes was synthesized using an in situ hydrothermal process [120]. The photocatalytic degradation of tetracycline was successfully investigated. Notably, the InVO4/Bi2WO6 semiconductor photocatalyst manufactured with 5.0 mg of Bi2WO6 exhibits the highest tetracycline degradation rate, that is, 97.42% in 72 min. Following the findings of quenching studies, hydroxyl radicals and holes prevail throughout the photocatalytic process. In addition, the improved nanocomposite does not lose its stability after being exposed to light for four cycles, highlighting the excellent reusability and photostability of the photocatalyst.

Photoreduction of carbon dioxide

The CO2 content in the atmosphere has risen from 280 to 408 ppm since the late 1800s. Officials and scholars are now concentrating their efforts on devising strategies that would significantly reduce atmospheric CO2 concentrations. The photochemical reduction of carbon dioxide to convert it to hydrocarbon fuels is one of the most promising options that have been found. Energy may be obtained from sustainable solar energy, which can be used directly or indirectly in this process, resulting in carbon recycling that lives up to its name since it can be conducted outside at ambient temperature and atmospheric pressure [121,122]. Therefore, using photocatalysis to remove excess CO2 from the atmosphere is of the utmost significance.

However, the photocatalytic reduction of CO2 is difficult for several reasons. First, breaking the bonds in CO2 requires significant energy. Second, reducing CO2 to methanol or methane requires processes involving the transfer of, respectively, six or eight electrons. These processes are far more intricate than the process transferring four electrons during the breakdown of water. Third, transforming carbon dioxide into methanol and other fuels requires complex, multi-step intermediate processes, such as proton transfer and hydroxylation. Fourth, in a system that exists in a liquid phase, the reduction of CO2 is typically followed by a competitive reaction (proton reduction) [123]. Furthermore, CO2 reduction produces a range of by-products, making catalyst selectivity crucial.

Among the candidates for CO2 photoreduction, bismuth-based photocatalysts have received enormous attention [124]. They offer many advantages, such as excellent bandgap features and a unique electronic structure. The vacancy defects in Bi-based photocatalysts could facilitate the adsorption and activation of CO2, resulting in further enhancing the efficiency of CO2 reduction. Moradi et al. combined sol–gel and photodeposition processes to fabricate Pt@Bi-TiO2 photocatalysts for converting CO2 [125]. As expected, methane production was noticeably improved over the Pt@Bi-TiO2 photocatalyst. Its methane yield was about 6.2 times greater than that of pure TiO2. The enhanced activity might be attributed to the efficient charge separation facilitated via Pt nanoparticles, and the increase of CO2 adsorption on phases containing Bi. Lui et al. [126] reported that a simple in situ one-step combustion approach was used to prepare Bi2Al4O9/β-Bi2O3 heterojunction. Urea was employed as the fuel for the reaction, while bismuth nitrate pentahydrate was used as a source of Bi2Al4O9 and β-Bi2O3. A transfer pathway of the photogenerated charges was suggested. Enhancement in CO yield was ascribed to oxygen vacancies, which improved the adsorption activation of CO2, and the photogenerated charge carriers effectively separated in the heterostructure. The CO2 photoreduction performance of the heterojunction of 0.14 Bi2Al4O9/β-Bi2O3was determined to be the highest. In contrast to the 1.5 μmol·g−1 CO yield on β-Bi2O3, the sample of 0.14 Bi2Al4O9/β-Bi2O3 gave 13.2 μmol·g−1 CO. Bi integration into TiO2 and subsequent loading of Pt on its surface led to a significant rise in methane generation. Methane yields were about 6.2 times higher with modified TiO2, which included 3 wt % Bi and 1.5 wt % Pt, compared to pure TiO2. Pt nanoparticles facilitated charge separation and boosted CO2 adsorption on phases containing Bi, which increased CO2 conversion activity.

Photocatalytic nitrogen fixation

Regarding the synthesis of NH3, photocatalytic N2 fixation (2N2 + 6H2O + light → 4NH3 + 3O2) is a promising concept. To further promote N2 photofixation, many studies have focused to develop effective photocatalysts.

Rong et al. proposed a Bi2Te3/BiOCl heterostructure as an effective candidate for N2 photofixation [127]. An NH3 release rate of 315.9 μmol·L−1·h−1 under UV-light irradiation was achieved. The improvement might be attributed to the effective prevention of hole–electron recombination. Bi2Te3 is a highly active semiconductor with a bandgap of 0.15 eV; however, its reduction potential of 0.57 eV is less than that of N2/NH3 (0.092 eV). As its CB potential (1.1 eV) is more negative than that of N2/NH3 (0.092 eV), BiOCl may photogenerate electrons and holes under UV light. So, the electrons photogenerated from BiOCl were crucial to the N2 photofixation. The photocatalytic activity of BiOCl is severely hampered by the recombination of photogenerated holes and charges. The photogenerated charges from the excitation of Bi2Te3 effectively suppress the recombination of photogenerated holes and electrons of BiOCl, which extends the lifetime of electrons in the CB of BiOCl. Without the presence of light or photocatalysts, no NH3 was measured. Fei et al. utilized graphene quantum dots and Bi2WO6 to construct a heterostructure for photocatalytic nitrogen fixation [128]. The photocatalytic performance was, respectively, 33.8 and 8.88 times better than that of pristine graphene quantum dots and Bi2WO6. Zhou et al. used an in situ bismuth reduction technique on Bi2WO6 [129]. Metallic Bi was used as a lattice junction to build Bi2WO6, which was highly oriented on the lattice structure. The directed interface and transfer channels made separating photogenerated carriers possible, leading to successful nitrogen fixation. In another work, Zhang et al. coated Bi2WO6 with cyclized polyacrylonitrile (c-PAN), which yielded active sites for N2 absorption and activation because the pyridinic N atom in c-PAN shifted into an unsaturated state to interact with small molecules and to transfer electrons to the small molecules [130].

Photocatalytic microbial disinfection

Heat and UV light are two methods to kill bacteria and viruses. Photocatalytic technology has also been used to disinfect and sterilize air, with positive results compared to traditional methods. Compared to TiO2 and ZnO, bismuth-based photocatalysts have a small bandgap and better visible light absorption [131]. Wang et al. reported a hydrothermally synthesized monoclinic dibismuth tetraoxide (m-Bi2O4) with a small bandgap of 2.0 eV and mixed valence states (Bi3+ and Bi5+) [132]. After 120 min of visible light irradiation, the m-Bi2O4 nanorods inactivated E. coli with a substantially better photocatalytic efficiency and photostability than CdS and Bi2O3. The primary ROS responsible for photocatalytic disinfection was found to be OH. This research also suggested that bismuth-based nanomaterials might be effective, stable, and long-lasting semiconductor photocatalysts for water disinfection under visible light. In another study, Liang et al. reported the fabrication of AgI/AgBr/BiOBr0.75I0.25 nanocomposites via solvothermal technique [133]. The fabricated semiconductor photocatalyst was used to deactivate Escherichia coli under visible light. A concentration of 80 mg/L AgI/AgBr/BiOBr0.75I0.25 was able to totally inactivate 3 × 107 CFU·mL−1 E. coli cells in 30 min. Furthermore, the bactericidal processes were thoroughly explored. The bactericidal activity of Ag+ ions generated from the nanocomposite was negligible, whereas active species such as h+, e, and O2 played critical roles in the disinfection. Direct interaction between bacterial cells and nanoparticles was discovered to be necessary for both the production of O2 and disinfection processes. E. coli cells were inactivated by disrupting the cell membrane and releasing cytoplasm. Furthermore, even after four repeated cycles, AgI/AgBr/BiOBr0.75I0.25 showed excellent antibacterial activity against E. coli. A novel Z-scheme AgI/BiVO4 heterojunction was synthesized via the chemical deposition–precipitation technique [134]. E. coli disinfection and the decomposition of oxytetracycline hydrochloride (OTC-HCl) were used to measure the photocatalytic activity under visible-light irradiation. The Z-scheme heterojunction took 50 min to kill bacteria and degraded OTC-HCl by 80% via photocatalysis, exhibiting high photocatalytic performance and photostability. h+ and O2 were shown to be the major reactive species during photoinactivation, with K+ permeability playing an important role in cell membrane collapse and bacterial deactivation. It was shown that the AgI/BiVO4 photocatalyst is an effective nanomaterial for wastewater treatment, specifically with extremely high concentrations of pathogenic microorganisms and antibiotics. Table 2 lists a Bi-based nanomaterials that have been studied for environmental applications.

Table 2: Different roles of Bi-based semiconductor photocatalysts in various environmental remediation applications.

No Photocatalyst Enhancement strategy Environmental Application Role of photocatalyst Ref.
1 BiOBr facet-dependent bacterial disinfection enhanced photocatalytic generation of radical [135]
2 Bi2O4/BiOBr nanosheets facet-dependent MO dye degradation, microbial disinfection enhanced light absorption, efficient photoinduced e/h+ pair separation, boosted surface-adsorbed ability [136]
3 BiO2−x Z-scheme heterojunction O2 evolution increased IEF [137]
4 Bi2MoO6/g-C3N4 heterojunction formation H2 evolution, microbial disinfection enhanced reduction capabilities [138]
5 BiOI/BiOBr heterojunction structure bacteriostatic activity advance separation of the photoinduced holes [139]
6 AgI/Bi2MoO6 Z-scheme water disinfection strong redox potential and enhanced separation of photogenerated charge carriers [140]
7 MoS2/Bi2WO6 p–n junction formation water detoxification generated electric field, efficient separation of photogenerated charges [141]
8 BiOCl surface defects CO2 photoreduction increased surface adsorption [142]
9 Bi4O5Br2 Bi-rich strategy CO2 photoreduction efficiently improved surface properties [143]
10 Bi2WO6/CuBi2O4 Z-scheme formation tetracycline dye degradation increased surface oxygen vacancy [144]
11 BiPO4/BiOCl0.9I0.1 Z-scheme phenol and RhB dye degradation boosted redox ability increased charge transfer [145]
12 Bi2S3/SnIn4S8 heterojunction formation microbial disinfection, RhB dye degradation advanced redox ability [146]
13 g-C3N4/Bi2WO6/AgI Z-scheme environmental decontamination enhanced the separation and transfer of photoinduced charges [147]
14 Bi2WO6/CdS heterojunction formation H2 evolution, RhB dye degradation enhanced electrochemical properties [148]
15 β-Bi2O3@g-C3N4 Z-scheme tetracycline dye degradation enhanced separation ability and prolonged lifespan of photogenerated charges [149]

Conclusion and Future Perspectives

There is great potential for Bi-based photocatalysts in the remediation of polluted environments and converting visible light into usable energy. In this study, we demonstrated that nanoscale Bi-based materials could be used for various photocatalytic ecological applications due to their distinctive electrical capabilities, crystal structure, and chemical attributes. These nanomaterials were obtained using various synthetic methods. Since there are constraints on employing just one Bi-based material as a photocatalyst, subsequent developments of Bi-based photocatalysts are explored in depth. The disadvantages are discussed in this article. We considered the most up-to-date research on Bi-based photocatalysts in our study. These developments stemmed from improvements in system design, microstructure management, and the synthesis of Bi-based composites. To further investigate the photocatalytic process, we analyzed the impact of several methods. The improved efficiency of photocatalysis based on Bi-valent cations was also covered.

Bi-based photocatalysts hold promise for environmental applications, but there is still much work to be done to improve the photocatalytic efficacy. Essential criteria in this field include the following: (1) Consideration must be given to the environmental and workplace safety and the economic impacts of the design process to overcome the “energy trilemma” when developing and building Bi-based photocatalysts. (2) The most difficult issue to address is obtaining and preserving the optimal nanostructure and morphology. (3) Thorough knowledge on reaction circumstances depending on structural traits is crucial. (4) Only there are only few studies on Bi-based photocatalysts in energy photocatalysis, such as H2 production, CO2 reduction, and selective organic transformation, because of the restricted number of photogenerated electrons from the less negative CB edge. (5) One of the most effective approaches to research reduction applications is to use Z-scheme heterojunctions with larger negative CB and photocatalytic Bi-based materials. (6) The stability of photocatalysts plays a vital role in large-scale application. Substantial research work has been already reported on it, but more ways to stabilize and recover a used Bi-based photocatalyst need to be found. Hence, more studies on the durability and renewability of Bi-based photocatalysts need to be carried out.

References

  1. He, R.; Xu, D.; Cheng, B.; Yu, J.; Ho, W. Nanoscale Horiz. 2018, 3, 464–504. doi:10.1039/c8nh00062j
    Return to citation in text: [1] [2] [3] [4]
  2. Liu, G.; Zhen, C.; Kang, Y.; Wang, L.; Cheng, H.-M. Chem. Soc. Rev. 2018, 47, 6410–6444. doi:10.1039/c8cs00396c
    Return to citation in text: [1]
  3. Di, J.; Xia, J.; Li, H.; Liu, Z. Nano Energy 2017, 35, 79–91. doi:10.1016/j.nanoen.2017.03.030
    Return to citation in text: [1]
  4. Xiong, J.; Song, P.; Di, J.; Li, H.; Liu, Z. J. Mater. Chem. A 2019, 7, 25203–25226. doi:10.1039/c9ta10144f
    Return to citation in text: [1]
  5. Dou, W.; Peng, X.; Kong, L.; Hu, X. Sci. Total Environ. 2022, 824, 153909. doi:10.1016/j.scitotenv.2022.153909
    Return to citation in text: [1]
  6. Wang, Z.; Mi, B. Environ. Sci. Technol. 2017, 51, 8229–8244. doi:10.1021/acs.est.7b01466
    Return to citation in text: [1]
  7. Navalón, S.; Dhakshinamoorthy, A.; Álvaro, M.; Garcia, H. ChemSusChem 2013, 6, 562–577. doi:10.1002/cssc.201200670
    Return to citation in text: [1]
  8. Parzinger, E.; Miller, B.; Blaschke, B.; Garrido, J. A.; Ager, J. W.; Holleitner, A.; Wurstbauer, U. ACS Nano 2015, 9, 11302–11309. doi:10.1021/acsnano.5b04979
    Return to citation in text: [1]
  9. Jin, J.; He, T. Appl. Surf. Sci. 2017, 394, 364–370. doi:10.1016/j.apsusc.2016.10.118
    Return to citation in text: [1]
  10. Chahkandi, M.; Zargazi, M. J. Hazard. Mater. 2019, 380, 120879. doi:10.1016/j.jhazmat.2019.120879
    Return to citation in text: [1]
  11. Zhang, L.-W.; Wang, Y.-J.; Cheng, H.-Y.; Yao, W.-Q.; Zhu, Y.-F. Adv. Mater. (Weinheim, Ger.) 2009, 21, 1286–1290. doi:10.1002/adma.200801354
    Return to citation in text: [1]
  12. Ouyang, S.; Tong, H.; Umezawa, N.; Cao, J.; Li, P.; Bi, Y.; Zhang, Y.; Ye, J. J. Am. Chem. Soc. 2012, 134, 1974–1977. doi:10.1021/ja210610h
    Return to citation in text: [1]
  13. Hsieh, P.-L.; Naresh, G.; Huang, Y.-S.; Tsao, C.-W.; Hsu, Y.-J.; Chen, L.-J.; Huang, M. H. J. Phys. Chem. C 2019, 123, 13664–13671. doi:10.1021/acs.jpcc.9b02081
    Return to citation in text: [1]
  14. Yu, H.; Jiang, L.; Wang, H.; Huang, B.; Yuan, X.; Huang, J.; Zhang, J.; Zeng, G. Small 2019, 15, 1901008. doi:10.1002/smll.201901008
    Return to citation in text: [1]
  15. Zhang, Z.; Wang, W.; Shang, M.; Yin, W. J. Hazard. Mater. 2010, 177, 1013–1018. doi:10.1016/j.jhazmat.2010.01.020
    Return to citation in text: [1]
  16. Yan, X.; Yuan, K.; Lu, N.; Xu, H.; Zhang, S.; Takeuchi, N.; Kobayashi, H.; Li, R. Appl. Catal., B 2017, 218, 20–31. doi:10.1016/j.apcatb.2017.06.022
    Return to citation in text: [1]
  17. Li, H.; Shang, J.; Ai, Z.; Zhang, L. J. Am. Chem. Soc. 2015, 137, 6393–6399. doi:10.1021/jacs.5b03105
    Return to citation in text: [1]
  18. Dong, C.; Lu, S.; Yao, S.; Ge, R.; Wang, Z.; Wang, Z.; An, P.; Liu, Y.; Yang, B.; Zhang, H. ACS Catal. 2018, 8, 8649–8658. doi:10.1021/acscatal.8b01645
    Return to citation in text: [1]
  19. Di, J.; Chen, C.; Zhu, C.; Ji, M.; Xia, J.; Yan, C.; Hao, W.; Li, S.; Li, H.; Liu, Z. Appl. Catal., B 2018, 238, 119–125. doi:10.1016/j.apcatb.2018.06.066
    Return to citation in text: [1]
  20. Zhao, Z.-Y.; Dai, W.-W. Inorg. Chem. 2015, 54, 10732–10737. doi:10.1021/acs.inorgchem.5b01714
    Return to citation in text: [1]
  21. Tan, C.; Cao, X.; Wu, X.-J.; He, Q.; Yang, J.; Zhang, X.; Chen, J.; Zhao, W.; Han, S.; Nam, G.-H.; Sindoro, M.; Zhang, H. Chem. Rev. 2017, 117, 6225–6331. doi:10.1021/acs.chemrev.6b00558
    Return to citation in text: [1]
  22. Di, J.; Xia, J.; Li, H.; Guo, S.; Dai, S. Nano Energy 2017, 41, 172–192. doi:10.1016/j.nanoen.2017.09.008
    Return to citation in text: [1]
  23. Xu, K.; Wang, L.; Xu, X.; Dou, S. X.; Hao, W.; Du, Y. Energy Storage Mater. 2019, 19, 446–463. doi:10.1016/j.ensm.2019.03.021
    Return to citation in text: [1]
  24. Dutta, V.; Sharma, S.; Raizada, P.; Khan, A. A. P.; Asiri, A. M.; Nadda, A.; Singh, P.; Van Le, Q.; Huang, C.-W.; Nguyen, D. L. T.; Pansambal, S.; Nguyen, V.-H. Surf. Interfaces 2021, 25, 101273. doi:10.1016/j.surfin.2021.101273
    Return to citation in text: [1]
  25. Ren, X.; Gao, M.; Zhang, Y.; Zhang, Z.; Cao, X.; Wang, B.; Wang, X. Appl. Catal., B 2020, 274, 119063. doi:10.1016/j.apcatb.2020.119063
    Return to citation in text: [1]
  26. Arthur, R. B.; Ahern, J. C.; Patterson, H. H. Catalysts 2018, 8, 604. doi:10.3390/catal8120604
    Return to citation in text: [1]
  27. Fenelon, E.; Bui, D.-P.; Tran, H. H.; You, S.-J.; Wang, Y.-F.; Cao, T. M.; Van Pham, V. ACS Omega 2020, 5, 20438–20449. doi:10.1021/acsomega.0c02461
    Return to citation in text: [1] [2]
  28. Xue, L.; An, F.; Yang, Y.; Ma, Y. J. Mater. Res. 2020, 35, 153–161. doi:10.1557/jmr.2019.401
    Return to citation in text: [1]
  29. Ding, S.; Han, M.; Dai, Y.; Yang, S.; Mao, D.; He, H.; Sun, C. ChemCatChem 2019, 11, 3490–3504. doi:10.1002/cctc.201900529
    Return to citation in text: [1]
  30. Ma, H.; He, Y.; Chen, P.; Wang, H.; Sun, Y.; Li, J.; Dong, F.; Xie, G.; Sheng, J. Chem. Eng. J. 2021, 417, 129305. doi:10.1016/j.cej.2021.129305
    Return to citation in text: [1]
  31. Sivasubramanian, P.; Chang, J.-H.; Nagendran, S.; Dong, C.-D.; Shkir, M.; Kumar, M. Chemosphere 2022, 307, 135652. doi:10.1016/j.chemosphere.2022.135652
    Return to citation in text: [1]
  32. Anku, W. W.; Oppong, S. O. B.; Govender, P. P. Bismuth-Based Nanoparticles as Photocatalytic Materials. In Bismuth - Advanced Applications and Defects Characterization; Zhou, Y.; Dong, F.; Jin, S., Eds.; IntechOpen: London, UK, 2018. doi:10.5772/intechopen.75104
    Return to citation in text: [1]
  33. Yang, C.; Chai, J.; Wang, Z.; Xing, Y.; Peng, J.; Yan, Q. Chem. Res. Chin. Univ. 2020, 36, 410–419. doi:10.1007/s40242-020-0069-3
    Return to citation in text: [1]
  34. Bhat, S. S. M.; Jang, H. W. ChemSusChem 2017, 10, 3001–3018. doi:10.1002/cssc.201700633
    Return to citation in text: [1] [2]
  35. Shahbazi, M.-A.; Faghfouri, L.; Ferreira, M. P. A.; Figueiredo, P.; Maleki, H.; Sefat, F.; Hirvonen, J.; Santos, H. A. Chem. Soc. Rev. 2020, 49, 1253–1321. doi:10.1039/c9cs00283a
    Return to citation in text: [1]
  36. Kudo, A.; Miseki, Y. Chem. Soc. Rev. 2009, 38, 253–278. doi:10.1039/b800489g
    Return to citation in text: [1]
  37. Meng, A.; Zhang, L.; Cheng, B.; Yu, J. Adv. Mater. (Weinheim, Ger.) 2019, 31, 1807660. doi:10.1002/adma.201807660
    Return to citation in text: [1]
  38. Low, J.; Yu, J.; Jaroniec, M.; Wageh, S.; Al-Ghamdi, A. A. Adv. Mater. (Weinheim, Ger.) 2017, 29, 1601694. doi:10.1002/adma.201601694
    Return to citation in text: [1] [2]
  39. Hasija, V.; Kumar, A.; Sudhaik, A.; Raizada, P.; Singh, P.; Van Le, Q.; Le, T. T.; Nguyen, V.-H. Environ. Chem. Lett. 2021, 19, 2941–2966. doi:10.1007/s10311-021-01231-w
    Return to citation in text: [1]
  40. Nosaka, Y.; Nosaka, A. Y. Chem. Rev. 2017, 117, 11302–11336. doi:10.1021/acs.chemrev.7b00161
    Return to citation in text: [1]
  41. Deng, Y.; Zhao, R. Curr. Pollut. Rep. 2015, 1, 167–176. doi:10.1007/s40726-015-0015-z
    Return to citation in text: [1]
  42. Huang, C. P.; Dong, C.; Tang, Z. Waste Manage. (Oxford, U. K.) 1993, 13, 361–377. doi:10.1016/0956-053x(93)90070-d
    Return to citation in text: [1]
  43. Xu, Q.; Zhang, L.; Yu, J.; Wageh, S.; Al-Ghamdi, A. A.; Jaroniec, M. Mater. Today 2018, 21, 1042–1063. doi:10.1016/j.mattod.2018.04.008
    Return to citation in text: [1] [2] [3]
  44. Serpone, N.; Maruthamuthu, P.; Pichat, P.; Pelizzetti, E.; Hidaka, H. J. Photochem. Photobiol., A 1995, 85, 247–255. doi:10.1016/1010-6030(94)03906-b
    Return to citation in text: [1]
  45. Xu, Q.; Zhang, L.; Cheng, B.; Fan, J.; Yu, J. Chem 2020, 6, 1543–1559. doi:10.1016/j.chempr.2020.06.010
    Return to citation in text: [1] [2] [3]
  46. He, R.; Cao, S.; Zhou, P.; Yu, J. Chin. J. Catal. 2014, 35, 989–1007. doi:10.1016/s1872-2067(14)60075-9
    Return to citation in text: [1]
  47. Dutta, V.; Sharma, S.; Raizada, P.; Kumar, R.; Thakur, V. K.; Nguyen, V.-H.; Asiri, A. M.; Khan, A. A. P.; Singh, P. J. Environ. Chem. Eng. 2020, 8, 104505. doi:10.1016/j.jece.2020.104505
    Return to citation in text: [1]
  48. Kwolek, P.; Szaciłowski, K. Electrochim. Acta 2013, 104, 448–453. doi:10.1016/j.electacta.2012.10.001
    Return to citation in text: [1] [2]
  49. Sharma, K.; Dutta, V.; Sharma, S.; Raizada, P.; Hosseini-Bandegharaei, A.; Thakur, P.; Singh, P. J. Ind. Eng. Chem. (Amsterdam, Neth.) 2019, 78, 1–20. doi:10.1016/j.jiec.2019.06.022
    Return to citation in text: [1]
  50. Liang, L.; Lei, F.; Gao, S.; Sun, Y.; Jiao, X.; Wu, J.; Qamar, S.; Xie, Y. Angew. Chem., Int. Ed. 2015, 54, 13971–13974. doi:10.1002/anie.201506966
    Return to citation in text: [1]
  51. Wu, X.; Toe, C. Y.; Su, C.; Ng, Y. H.; Amal, R.; Scott, J. J. Mater. Chem. A 2020, 8, 15302–15318. doi:10.1039/d0ta01180k
    Return to citation in text: [1]
  52. Chou, S.-Y.; Chen, C.-C.; Dai, Y.-M.; Lin, J.-H.; Lee, W. W. RSC Adv. 2016, 6, 33478–33491. doi:10.1039/c5ra28024a
    Return to citation in text: [1]
  53. Liu, F.-Y.; Dai, Y.-M.; Chen, F.-H.; Chen, C.-C. J. Colloid Interface Sci. 2020, 562, 112–124. doi:10.1016/j.jcis.2019.12.006
    Return to citation in text: [1]
  54. Su, Y.; Ding, C.; Dang, Y.; Wang, H.; Ye, L.; Jin, X.; Xie, H.; Liu, C. Appl. Surf. Sci. 2015, 346, 311–316. doi:10.1016/j.apsusc.2015.04.021
    Return to citation in text: [1]
  55. Lin, Y.; Lu, C.; Wei, C. J. Alloys Compd. 2019, 781, 56–63. doi:10.1016/j.jallcom.2018.12.071
    Return to citation in text: [1]
  56. Huang, H.; Liu, L.; Zhang, Y.; Tian, N. J. Alloys Compd. 2015, 619, 807–811. doi:10.1016/j.jallcom.2014.08.262
    Return to citation in text: [1]
  57. Le, M. T.; Van Well, W. J. M.; Van Driessche, I.; Hoste, S. Appl. Catal., A 2004, 267, 227–234. doi:10.1016/j.apcata.2004.03.007
    Return to citation in text: [1]
  58. He, Y.; Zhang, Y.; Huang, H.; Tian, N.; Guo, Y.; Luo, Y. Colloids Surf., A 2014, 462, 131–136. doi:10.1016/j.colsurfa.2014.07.034
    Return to citation in text: [1]
  59. Hamza, M. A.; El-Shazly, A. N.; Tolba, S. A.; Allam, N. K. Chem. Eng. J. 2020, 384, 123351. doi:10.1016/j.cej.2019.123351
    Return to citation in text: [1]
  60. Tian, N.; Zhang, Y.; Huang, H.; He, Y.; Guo, Y. J. Phys. Chem. C 2014, 118, 15640–15648. doi:10.1021/jp500645p
    Return to citation in text: [1]
  61. Dong, F.; Li, Q.; Sun, Y.; Ho, W.-K. ACS Catal. 2014, 4, 4341–4350. doi:10.1021/cs501038q
    Return to citation in text: [1]
  62. Dong, F.; Xiong, T.; Sun, Y.; Huang, H.; Wu, Z. J. Mater. Chem. A 2015, 3, 18466–18474. doi:10.1039/c5ta05099e
    Return to citation in text: [1]
  63. Hou, J.; Wei, R.; Wu, X.; Tahir, M.; Wang, X.; Butt, F. K.; Cao, C. Dalton Trans. 2018, 47, 6692–6701. doi:10.1039/c8dt00570b
    Return to citation in text: [1]
  64. Yu, S.; Zhang, Y.; Li, M.; Du, X.; Huang, H. Appl. Surf. Sci. 2017, 391, 491–498. doi:10.1016/j.apsusc.2016.07.028
    Return to citation in text: [1]
  65. Xiao, L.; Lin, R.; Wang, J.; Cui, C.; Wang, J.; Li, Z. J. Colloid Interface Sci. 2018, 523, 151–158. doi:10.1016/j.jcis.2018.03.064
    Return to citation in text: [1]
  66. Jamkhande, P. G.; Ghule, N. W.; Bamer, A. H.; Kalaskar, M. G. J. Drug Delivery Sci. Technol. 2019, 53, 101174. doi:10.1016/j.jddst.2019.101174
    Return to citation in text: [1] [2] [3]
  67. Rajput, N. Int. J. Adv. Eng. Technol. 2015, 7, 1806–1811. doi:10.7323/ijaet/v7_iss6
    Return to citation in text: [1]
  68. Rane, A. V.; Kanny, K.; Abitha, V. K.; Thomas, S. Methods for synthesis of nanoparticles and fabrication of nanocomposites. Synthesis of inorganic nanomaterials; Woodhead Publishing: Duxford, UK, 2018; pp 121–139. doi:10.1016/b978-0-08-101975-7.00005-1
    Return to citation in text: [1] [2]
  69. Cipolatti, E. P.; Valério, A.; Henriques, R. O.; Moritz, D. E.; Ninow, J. L.; Freire, D. M. G.; Manoel, E. A.; Fernandez-Lafuente, R.; de Oliveira, D. RSC Adv. 2016, 6, 104675–104692. doi:10.1039/c6ra22047a
    Return to citation in text: [1]
  70. Zhu, X. H.; Hang, Q. M. Micron 2013, 44, 21–44. doi:10.1016/j.micron.2012.06.005
    Return to citation in text: [1]
  71. Abbas, A.; Amin, H. M. A. Microchem. J. 2022, 175, 107166. doi:10.1016/j.microc.2021.107166
    Return to citation in text: [1]
  72. Trino, L. D.; Bronze-Uhle, E. S.; George, A.; Mathew, M. T.; Lisboa-Filho, P. N. Colloids Surf., A 2018, 546, 168–178. doi:10.1016/j.colsurfa.2018.03.019
    Return to citation in text: [1]
  73. Nasr, M.; Eid, C.; Habchi, R.; Miele, P.; Bechelany, M. ChemSusChem 2018, 11, 3023–3047. doi:10.1002/cssc.201800874
    Return to citation in text: [1]
  74. Welch, C. M.; Compton, R. G. Anal. Bioanal. Chem. 2006, 384, 601–619. doi:10.1007/s00216-005-0230-3
    Return to citation in text: [1]
  75. Zhang, L.; Wang, W.; Zhou, L.; Xu, H. Small 2007, 3, 1618–1625. doi:10.1002/smll.200700043
    Return to citation in text: [1]
  76. Hagfeldt, A.; Graetzel, M. Chem. Rev. 1995, 95, 49–68. doi:10.1021/cr00033a003
    Return to citation in text: [1]
  77. Wang, Z.; Jiang, C.; Huang, R.; Peng, H.; Tang, X. J. Phys. Chem. C 2014, 118, 1155–1160. doi:10.1021/jp4065505
    Return to citation in text: [1]
  78. Ma, D.; Zhao, J.; Chu, R.; Yang, S.; Zhao, Y.; Hao, X.; Li, L.; Zhang, L.; Lu, Y.; Yu, C. Adv. Powder Technol. 2013, 24, 79–85. doi:10.1016/j.apt.2012.02.004
    Return to citation in text: [1]
  79. Ma, D.; Zhao, J.; Li, Y.; Su, X.; Hou, S.; Zhao, Y.; Hao, X.; Li, L. Colloids Surf., A 2010, 368, 105–111. doi:10.1016/j.colsurfa.2010.07.022
    Return to citation in text: [1]
  80. Rao, F.; Zhu, G.; Zhang, W.; Gao, J.; Zhang, F.; Huang, Y.; Hojamberdiev, M. Appl. Catal., B 2021, 281, 119481. doi:10.1016/j.apcatb.2020.119481
    Return to citation in text: [1]
  81. Huang, H.; Xiao, K.; Zhang, T.; Dong, F.; Zhang, Y. Appl. Catal., B 2017, 203, 879–888. doi:10.1016/j.apcatb.2016.10.082
    Return to citation in text: [1] [2]
  82. Xie, T.; Zhang, Y.; Yao, W.; Liu, Y.; Wang, H.; Wu, Z. Catal. Sci. Technol. 2019, 9, 1178–1188. doi:10.1039/c8cy02344a
    Return to citation in text: [1]
  83. Zou, Z.; Xu, H.; Li, D.; Sun, J.; Xia, D. Appl. Surf. Sci. 2019, 463, 1011–1018. doi:10.1016/j.apsusc.2018.09.025
    Return to citation in text: [1]
  84. Hu, T.; Li, H.; Du, N.; Hou, W. ChemCatChem 2018, 10, 3040–3048. doi:10.1002/cctc.201701965
    Return to citation in text: [1] [2]
  85. Zhou, S.; Zhang, Q.; Zhao, D.; Zong, W.; Fan, Z.; Sun, Y.; Xu, X. Appl. Organomet. Chem. 2018, 32, e4230. doi:10.1002/aoc.4230
    Return to citation in text: [1]
  86. Merupo, V. I.; Velumani, S.; Oza, G.; Tabellout, M.; Bizarro, M.; Coste, S.; Kassiba, A. H. ChemistrySelect 2016, 1, 1278–1286. doi:10.1002/slct.201600090
    Return to citation in text: [1]
  87. Huang, Y.; Kang, S.; Yang, Y.; Qin, H.; Ni, Z.; Yang, S.; Li, X. Appl. Catal., B 2016, 196, 89–99. doi:10.1016/j.apcatb.2016.05.022
    Return to citation in text: [1]
  88. Dong, Y.; Xu, D.; Wang, Q.; Zhang, G.; Zhang, Q.; Zhang, Z.; Lv, L.; Xia, Y.; Ren, Z.; Wang, P. Appl. Surf. Sci. 2021, 542, 148521. doi:10.1016/j.apsusc.2020.148521
    Return to citation in text: [1]
  89. Huang, H.; Li, X.; Wang, J.; Dong, F.; Chu, P. K.; Zhang, T.; Zhang, Y. ACS Catal. 2015, 5, 4094–4103. doi:10.1021/acscatal.5b00444
    Return to citation in text: [1]
  90. Wu, Q.; Chai, S.; Yang, H.; Gao, Z.; Zhang, R.; Wang, L.; Kang, L. Sep. Purif. Technol. 2020, 253, 117388. doi:10.1016/j.seppur.2020.117388
    Return to citation in text: [1] [2]
  91. Xiong, T.; Wen, M.; Dong, F.; Yu, J.; Han, L.; Lei, B.; Zhang, Y.; Tang, X.; Zang, Z. Appl. Catal., B 2016, 199, 87–95. doi:10.1016/j.apcatb.2016.06.032
    Return to citation in text: [1]
  92. Zhou, F. Q.; Fan, J. C.; Xu, Q. J.; Min, Y. L. Appl. Catal., B 2017, 201, 77–83. doi:10.1016/j.apcatb.2016.08.027
    Return to citation in text: [1]
  93. He, Z.; Shi, Y.; Gao, C.; Wen, L.; Chen, J.; Song, S. J. Phys. Chem. C 2014, 118, 389–398. doi:10.1021/jp409598s
    Return to citation in text: [1]
  94. Wang, S.; Yang, X.; Zhang, X.; Ding, X.; Yang, Z.; Dai, K.; Chen, H. Appl. Surf. Sci. 2017, 391, 194–201. doi:10.1016/j.apsusc.2016.07.070
    Return to citation in text: [1]
  95. Zou, X.; Dong, Y.; Zhang, X.; Cui, Y.; Ou, X.; Qi, X. Appl. Surf. Sci. 2017, 391, 525–534. doi:10.1016/j.apsusc.2016.06.003
    Return to citation in text: [1]
  96. Yi, J.; Jiao, C.; Mo, H.; Chen, Q.; She, Q.; Li, Z. Ceram. Int. 2018, 44, 22102–22107. doi:10.1016/j.ceramint.2018.08.320
    Return to citation in text: [1]
  97. Sang, Y.; Cao, X.; Dai, G.; Wang, L.; Peng, Y.; Geng, B. J. Hazard. Mater. 2020, 381, 120942. doi:10.1016/j.jhazmat.2019.120942
    Return to citation in text: [1]
  98. Zhao, Q.; Lu, L.; Wang, B.; Jiang, T. React. Kinet., Mech. Catal. 2021, 132, 581–597. doi:10.1007/s11144-020-01916-3
    Return to citation in text: [1]
  99. Lv, Y.-R.; He, R.-K.; Chen, Z.-Y.; Li, X.; Xu, Y.-H. J. Colloid Interface Sci. 2020, 560, 293–302. doi:10.1016/j.jcis.2019.10.064
    Return to citation in text: [1]
  100. Bard, A. J. J. Photochem. 1979, 10, 59–75. doi:10.1016/0047-2670(79)80037-4
    Return to citation in text: [1]
  101. Wang, Y.; Zhang, Z.; Zhang, L.; Luo, Z.; Shen, J.; Lin, H.; Long, J.; Wu, J. C. S.; Fu, X.; Wang, X.; Li, C. J. Am. Chem. Soc. 2018, 140, 14595–14598. doi:10.1021/jacs.8b09344
    Return to citation in text: [1]
  102. Fu, F.; Shen, H.; Xue, W.; Zhen, Y.; Soomro, R. A.; Yang, X.; Wang, D.; Xu, B.; Chi, R. J. Catal. 2019, 375, 399–409. doi:10.1016/j.jcat.2019.06.033
    Return to citation in text: [1]
  103. Hu, J.; Chen, C.; Zheng, Y.; Zhang, G.; Guo, C.; Li, C. M. Small 2020, 16, 2002988. doi:10.1002/smll.202002988
    Return to citation in text: [1]
  104. Bao, Y.; Song, S.; Yao, G.; Jiang, S. Sol. RRL 2021, 5, 2100118. doi:10.1002/solr.202100118
    Return to citation in text: [1]
  105. Xu, A.; Tu, W.; Shen, S.; Lin, Z.; Gao, N.; Zhong, W. Appl. Surf. Sci. 2020, 528, 146949. doi:10.1016/j.apsusc.2020.146949
    Return to citation in text: [1]
  106. Liu, Y.; Liu, C.; Shi, C.; Sun, W.; Lin, X.; Shi, W.; Hong, Y. J. Alloys Compd. 2021, 881, 160437. doi:10.1016/j.jallcom.2021.160437
    Return to citation in text: [1]
  107. Li, Q.; Zhao, W.; Zhai, Z.; Ren, K.; Wang, T.; Guan, H.; Shi, H. J. Mater. Sci. Technol. 2020, 56, 216–226. doi:10.1016/j.jmst.2020.03.038
    Return to citation in text: [1]
  108. Kong, X. Y.; Choo, Y. Y.; Chai, S.-P.; Soh, A. K.; Mohamed, A. R. Chem. Commun. 2016, 52, 14242–14245. doi:10.1039/c6cc07750a
    Return to citation in text: [1]
  109. Jin, X.; Ye, L.; Xie, H.; Chen, G. Coord. Chem. Rev. 2017, 349, 84–101. doi:10.1016/j.ccr.2017.08.010
    Return to citation in text: [1]
  110. Wang, Z.; Li, C.; Domen, K. Chem. Soc. Rev. 2019, 48, 2109–2125. doi:10.1039/c8cs00542g
    Return to citation in text: [1]
  111. Maeda, K.; Domen, K. J. Phys. Chem. Lett. 2010, 1, 2655–2661. doi:10.1021/jz1007966
    Return to citation in text: [1]
  112. Zou, Z.; Ye, J.; Sayama, K.; Arakawa, H. Nature 2001, 414, 625–627. doi:10.1038/414625a
    Return to citation in text: [1]
  113. Martinez Suarez, C.; Hernández, S.; Russo, N. Appl. Catal., A 2015, 504, 158–170. doi:10.1016/j.apcata.2014.11.044
    Return to citation in text: [1]
  114. Cao, D.; Wang, Q.; Wu, Y.; Zhu, S.; Jia, Y.; Wang, R. Sep. Purif. Technol. 2020, 250, 117132. doi:10.1016/j.seppur.2020.117132
    Return to citation in text: [1] [2]
  115. Yang, C.; Xue, Z.; Qin, J.; Sawangphruk, M.; Rajendran, S.; Zhang, X.; Liu, R. J. Phys. Chem. C 2019, 123, 4795–4804. doi:10.1021/acs.jpcc.8b10604
    Return to citation in text: [1] [2]
  116. Malathi, A.; Arunachalam, P.; Grace, A. N.; Madhavan, J.; Al-Mayouf, A. M. Appl. Surf. Sci. 2017, 412, 85–95. doi:10.1016/j.apsusc.2017.03.199
    Return to citation in text: [1]
  117. Priya, A.; Arunachalam, P.; Selvi, A.; Madhavan, J.; Al-Mayouf, A. M. Colloids Surf., A 2018, 559, 83–91. doi:10.1016/j.colsurfa.2018.09.031
    Return to citation in text: [1]
  118. Rahman, M. U.; Qazi, U. Y.; Hussain, T.; Nadeem, N.; Zahid, M.; Bhatti, H. N.; Shahid, I. Opt. Mater. (Amsterdam, Neth.) 2021, 120, 111408. doi:10.1016/j.optmat.2021.111408
    Return to citation in text: [1]
  119. Chava, R. K.; Son, N.; Kang, M. Chemosphere 2022, 300, 134570. doi:10.1016/j.chemosphere.2022.134570
    Return to citation in text: [1]
  120. Sreeram, N.; Aruna, V.; Koutavarapu, R.; Lee, D.-Y.; Shim, J. Chemosphere 2022, 299, 134477. doi:10.1016/j.chemosphere.2022.134477
    Return to citation in text: [1]
  121. Larimi, A.; Rahimi, M.; Khorasheh, F. Renewable Energy 2020, 145, 1862–1869. doi:10.1016/j.renene.2019.07.105
    Return to citation in text: [1]
  122. Lee, J. H.; Lee, H.; Kang, M. Mater. Lett. 2016, 178, 316–319. doi:10.1016/j.matlet.2016.04.193
    Return to citation in text: [1]
  123. Li, L.; Li, P.; Wang, Y.; Lin, L.; Shah, A. H.; He, T. Appl. Surf. Sci. 2018, 452, 498–506. doi:10.1016/j.apsusc.2018.04.256
    Return to citation in text: [1]
  124. Li, X.; Zhuang, Z.; Li, W.; Pan, H. Appl. Catal., A 2012, 429-430, 31–38. doi:10.1016/j.apcata.2012.04.001
    Return to citation in text: [1]
  125. Moradi, M.; Khorasheh, F.; Larimi, A. Sol. Energy 2020, 211, 100–110. doi:10.1016/j.solener.2020.09.054
    Return to citation in text: [1]
  126. Liu, Y.; Guo, J.-g.; Wang, Y.; Hao, Y.-j.; Liu, R.-h.; Li, F.-t. Green Energy Environ. 2021, 6, 244–252. doi:10.1016/j.gee.2020.04.014
    Return to citation in text: [1]
  127. Rong, X.; Mao, Y.; Xu, J.; Zhang, X.; Zhang, L.; Zhou, X.; Qiu, F.; Wu, Z. Catal. Commun. 2018, 116, 16–19. doi:10.1016/j.catcom.2018.07.018
    Return to citation in text: [1]
  128. Fei, T.; Yu, L.; Liu, Z.; Song, Y.; Xu, F.; Mo, Z.; Liu, C.; Deng, J.; Ji, H.; Cheng, M.; Lei, Y.; Xu, H.; Li, H. J. Colloid Interface Sci. 2019, 557, 498–505. doi:10.1016/j.jcis.2019.09.011
    Return to citation in text: [1]
  129. Zhou, S.; Zhang, C.; Liu, J.; Liao, J.; Kong, Y.; Xu, Y.; Chen, G. Catal. Sci. Technol. 2019, 9, 5562–5566. doi:10.1039/c9cy00972h
    Return to citation in text: [1]
  130. Zhang, Y.; Yu, L.; Tang, L.; Zhu, M.; Jin, Y.; Wang, Z.; Li, L. Engineering (Irvine, CA, U. S.) 2021, 7, 11–13. doi:10.1016/j.eng.2020.03.006
    Return to citation in text: [1]
  131. Ning, S.; Lin, H.; Tong, Y.; Zhang, X.; Lin, Q.; Zhang, Y.; Long, J.; Wang, X. Appl. Catal., B 2017, 204, 1–10. doi:10.1016/j.apcatb.2016.11.006
    Return to citation in text: [1]
  132. Wang, W.; Chen, X.; Liu, G.; Shen, Z.; Xia, D.; Wong, P. K.; Yu, J. C. Appl. Catal., B 2015, 176-177, 444–453. doi:10.1016/j.apcatb.2015.04.026
    Return to citation in text: [1]
  133. Liang, J.; Deng, J.; Li, M.; Tong, M. Colloids Surf., B 2016, 138, 102–109. doi:10.1016/j.colsurfb.2015.11.060
    Return to citation in text: [1]
  134. Guan, D.-L.; Niu, C.-G.; Wen, X.-J.; Guo, H.; Deng, C.-H.; Zeng, G.-M. J. Colloid Interface Sci. 2018, 512, 272–281. doi:10.1016/j.jcis.2017.10.068
    Return to citation in text: [1]
  135. Wu, D.; Wang, B.; Wang, W.; An, T.; Li, G.; Ng, T. W.; Yip, H. Y.; Xiong, C.; Lee, H. K.; Wong, P. K. J. Mater. Chem. A 2015, 3, 15148–15155. doi:10.1039/c5ta02757h
    Return to citation in text: [1]
  136. Wu, D.; Ye, L.; Yue, S.; Wang, B.; Wang, W.; Yip, H. Y.; Wong, P. K. J. Phys. Chem. C 2016, 120, 7715–7727. doi:10.1021/acs.jpcc.6b02365
    Return to citation in text: [1]
  137. Wang, M.; Tan, G.; Zhang, D.; Li, B.; Lv, L.; Wang, Y.; Ren, H.; Zhang, X.; Xia, A.; Liu, Y. Appl. Catal., B 2019, 254, 98–112. doi:10.1016/j.apcatb.2019.04.044
    Return to citation in text: [1]
  138. Li, J.; Yin, Y.; Liu, E.; Ma, Y.; Wan, J.; Fan, J.; Hu, X. J. Hazard. Mater. 2017, 321, 183–192. doi:10.1016/j.jhazmat.2016.09.008
    Return to citation in text: [1]
  139. Wang, Y.; Lin, L.; Li, F.; Chen, L.; Chen, D.; Yang, C.; Huang, M. Photochem. Photobiol. Sci. 2016, 15, 666–672. doi:10.1039/c5pp00406c
    Return to citation in text: [1]
  140. Liang, J.; Liu, F.; Deng, J.; Li, M.; Tong, M. Water Res. 2017, 123, 632–641. doi:10.1016/j.watres.2017.06.060
    Return to citation in text: [1]
  141. Meng, X.; Li, Z.; Zeng, H.; Chen, J.; Zhang, Z. Appl. Catal., B 2017, 210, 160–172. doi:10.1016/j.apcatb.2017.02.083
    Return to citation in text: [1]
  142. Gong, S.; Zhu, G.; Wang, R.; Rao, F.; Shi, X.; Gao, J.; Huang, Y.; He, C.; Hojamberdiev, M. Appl. Catal., B 2021, 297, 120413. doi:10.1016/j.apcatb.2021.120413
    Return to citation in text: [1]
  143. Ye, L.; Jin, X.; Liu, C.; Ding, C.; Xie, H.; Chu, K. H.; Wong, P. K. Appl. Catal., B 2016, 187, 281–290. doi:10.1016/j.apcatb.2016.01.044
    Return to citation in text: [1]
  144. Yuan, X.; Shen, D.; Zhang, Q.; Zou, H.; Liu, Z.; Peng, F. Chem. Eng. J. 2019, 369, 292–301. doi:10.1016/j.cej.2019.03.082
    Return to citation in text: [1]
  145. Yue, P.; Zhang, G.; Cao, X.; Wang, B.; Zhang, Y.; Wei, Y. Sep. Purif. Technol. 2019, 213, 34–44. doi:10.1016/j.seppur.2018.12.003
    Return to citation in text: [1]
  146. Shi, H.; Zhao, Y.; Fan, J.; Tang, Z. Appl. Surf. Sci. 2019, 465, 212–222. doi:10.1016/j.apsusc.2018.09.164
    Return to citation in text: [1]
  147. Xue, W.; Huang, D.; Li, J.; Zeng, G.; Deng, R.; Yang, Y.; Chen, S.; Li, Z.; Gong, X.; Li, B. Chem. Eng. J. 2019, 373, 1144–1157. doi:10.1016/j.cej.2019.05.069
    Return to citation in text: [1]
  148. Arif, M.; Min, Z.; Yuting, L.; Yin, H.; Liu, X. J. Ind. Eng. Chem. (Amsterdam, Neth.) 2019, 69, 345–357. doi:10.1016/j.jiec.2018.09.026
    Return to citation in text: [1]
  149. Hong, Y.; Li, C.; Yin, B.; Li, D.; Zhang, Z.; Mao, B.; Fan, W.; Gu, W.; Shi, W. Chem. Eng. J. 2018, 338, 137–146. doi:10.1016/j.cej.2017.12.108
    Return to citation in text: [1]
Other Beilstein-Institut Open Science Activities