Next Article in Journal
Antibacterial Applications of Nanomaterials
Previous Article in Journal
Sorption and Desorption of Vapor of n-Pentane by Porphyrin Aluminum Metal–Organic Framework: Mechanism of Bonding, Kinetics and Stoichiometry by Complementary In-Situ Time-Dependent and Ex-Situ Methods
Previous Article in Special Issue
Eco-Friendly Reduction of Graphene Oxide by Aqueous Extracts for Photocatalysis Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

BiVO4 As a Sustainable and Emerging Photocatalyst: Synthesis Methodologies, Engineering Properties, and Its Volatile Organic Compounds Degradation Efficiency

by
Ganesh S. Kamble
1,*,
Thillai Sivakumar Natarajan
2,3,
Santosh S. Patil
4,
Molly Thomas
5,
Rajvardhan K. Chougale
1,
Prashant D. Sanadi
1,
Umesh S. Siddharth
6 and
Yong-Chein Ling
7,*
1
Department of Engineering Chemistry, Kolhapur Institute of Technology’s College of Engineering (Autonomous), Kolhapur Affiliated Shivaji University Kolhapur Maharashtra, Kolhapur 416004, Maharashtra, India
2
Environmental Science Laboratory, CSIR-Central Leather Research Institute (CSIR-CLRI), Chennai 600020, Tamil Nadu, India
3
Academy of Scientific and Innovative Research (AcSIR), Ghaziabad 600113, Uttar Pradesh, India
4
Department of Applied Mechanics, ECTO Group, FEMTO-ST Institute, 24, Rue de l’Epitaph, 25000 Besançon, France
5
School of Studies in Chemistry & Research Centre, Maharaja Chhatrasal Bundelkhand University, Chhatarpur 471001, Madhya Pradesh, India
6
Department of Basic Sciences and Humanities, Sharad Institute of Technology College of Engineering Yadrav (Ichalkaranji), Ichalkaranji 416115, Maharashtra, India
7
Department of Chemistry, National Tsing Hua University, Hsinchu 300044, Taiwan
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(9), 1528; https://doi.org/10.3390/nano13091528
Submission received: 10 February 2023 / Revised: 10 April 2023 / Accepted: 13 April 2023 / Published: 1 May 2023
(This article belongs to the Special Issue Nanomaterials for Photodegradation of Pollutants)

Abstract

:
Bismuth vanadate (BiVO4) is one of the best bismuth-based semiconducting materials because of its narrow band gap energy, good visible light absorption, unique physical and chemical characteristics, and non-toxic nature. In addition, BiVO4 with different morphologies has been synthesized and exhibited excellent visible light photocatalytic efficiency in the degradation of various organic pollutants, including volatile organic compounds (VOCs). Nevertheless, the commercial scale utilization of BiVO4 is significantly limited because of the poor separation (faster recombination rate) and transport ability of photogenerated electron–hole pairs. So, engineering/modifications of BiVO4 materials are performed to enhance their structural, electronic, and morphological properties. Thus, this review article aims to provide a critical overview of advanced oxidation processes (AOPs), various semiconducting nanomaterials, BiVO4 synthesis methodologies, engineering of BiVO4 properties through making binary and ternary nanocomposites, and coupling with metals/non-metals and metal nanoparticles and the development of Z-scheme type nanocomposites, etc., and their visible light photocatalytic efficiency in VOCs degradation. In addition, future challenges and the way forward for improving the commercial-scale application of BiVO4-based semiconducting nanomaterials are also discussed. Thus, we hope that this review is a valuable resource for designing BiVO4-based nanocomposites with superior visible-light-driven photocatalytic efficiency in VOCs degradation.

1. Introduction

In the 21st century, environmental protection and remediation are the greatest challenges for human beings due to massive population increases and the growth of industrialization [1]. Our natural water has been significantly damaged and continues to deteriorate due to human activity and the growth of chemical, agriculture, and pharmaceutical industries. Textile industries, in particular, discharge annually 15% (one thousand tons) of hazardous dyes as effluents [2]. Along with these, a vast amount of toxic and harmful chemicals, heavy metals [3], ionic liquids [4], surfactants [5] agrochemicals, and pharmaceutical industry wastes, including drugs and antibiotics wastes [6], are discharged into fresh water. Organic effluents are highly carcinogenic, leading to a negative impact on the sustainability of water resources [7]. It is anticipated that up to 50% of people will face clean water disasters by 2025 because approximately 70% of industrial wastewater is not well treated and waste effluent is released directly into freshwater sources, causing the severe pollution of natural water bodies [8]. Therefore, the removal of these hazardous pollutants from wastewater before their discharge into the environment is a remarkable challenge across the globe to protect the environment and natural water resources.
Volatile organic compounds (VOCs) are one of the most hazardous organic effluents discharged directly into water by the paint, pharmaceutical, chemical, mining, printing, and petrochemical industries. The most common VOCs used industrially are acetone, formaldehyde, benzene, toluene, xylene, 1,3-butadiene, ethylene glycol, perchloroethylene, methylene chloride, and chlorobenzene, which cause severe health hazards to human health, including toxic, carcinogenic, mutagenic, and teratogenic effects [9,10]. For the elimination of VOCs, over the years, various environmental pollutant remediation technologies, such as membrane separation, adsorption, biological oxidation, and chemical oxidation, have been used, but these can generate secondary pollutants, which require post-treatment and are inefficient for field-level application [11,12]. In order to overcome these problems, heterogeneous semiconductor photocatalysis is one of the promising advanced oxidation processes (AOPs) used for the removal of VOCs and pollutants from wastewater that has been studied as it has the advantages of an environmentally friendly process, low cost, mild reaction conditions, no generation of secondary pollution, and greater effectiveness in the degradation of even low concentrations of pollutants compared to conventional technologies [13,14,15,16,17]. Various well-known stable metal oxides (e.g., TiO2, WO3, and ZnO, etc.), metal titanates (SrTiO3), etc., have been developed and have demonstrated an efficient photocatalytic performance for the degradation of various organic compounds. However, the photocatalytic efficiency of these metal oxides is limited due to their large band gap energy (3.2 eV), response to ultraviolet (UV) light only, and utilization of less than 4% of the available solar spectrum, which restricts their potential commercial application [18,19,20,21]. Therefore, during the last decade, research has been focused on constructing various visible-light-responsive semiconductor photocatalysts, such as metal sulfides, oxides, oxynitrides, chalcogenides, halides, and oxyhalides, for the degradation of hazardous organic pollutants [22,23,24,25,26,27,28,29,30]. However, metal sulfides and some chalcogenides-based catalytic systems show poor stability under light irradiation, which restricts their commercial feasibility. Therefore, the development of metal-oxide-based, visible-light-responsive photocatalysts, especially bismuth-based metal oxide (BiVO4, Bi2MoO6, Bi2WO6, BiFe2O3, BiFe4O9, and BiOX (X = Cl, Br, I, etc.,)) systems, has received significant attention, and these are being used for the degradation of various organic pollutants, including VOCs [31,32,33,34,35,36].
Among bismuth metal oxides, bismuth vanadate (BiVO4) has attracted significant interest due to its outstanding visible light absorption (~2.3–2.4 eV band gap energy), suitable band potentials, facet-dependent catalytic activity, non-toxicity, and resistance to photo and chemical corrosion [37,38,39,40,41,42]. Kudo et al., in 1999, developed highly crystalline monoclinic and tetragonal BiVO4 by changing the ratio of vanadium to bismuth in starting materials, producing O2 evolution under visible light irradiation. Subsequently, significant research efforts have been made to develop BiVO4-based systems for the degradation of pollutants [43]. However, there is a high recombination rate of photo-generated charge carriers, and the conduction band position of BiVO4 is lower than the superoxide radical anion production potential, which decreases the photocatalytic activity. Consequently, different modifications to BiVO4, such as metal and non-metal loading, the control of morphologies, and the formation of heterojunctions with bismuth and non-bismuth-based oxides, have been performed, which has led to various reviews and research articles on the degradation of water pollutants. However, review articles on BiVO4-based materials for VOCs degradation are rarely reported. Therefore, the present review covers recent developments in BiVO4-based materials and their potential visible light photocatalytic applications for VOCs degradation. Specifically, the synthesis methodologies of BiVO4-based materials, the engineering of BiVO4 to achieve changes in its structural and electronic properties, and the correlation of these properties with improvement in visible light photocatalytic activity are also discussed. Finally, further prospects for BiVO4-based materials are presented, which may provide a better understanding and encourage the field-scale application of BiVO4-based materials for VOCs degradation.

2. Advanced Oxidation Process and Semiconductor Photocatalysis

Advanced oxidation processes (AOPs) are amongst the widely accepted eco-friendly processes for the treatment of different wastewaters. They involve the in-situ generation of hydroxyl (OH) and sulphate (SO4) radicals, which are strong oxidants for the oxidation of various types of toxic organic pollutants. Among these, the hydroxyl (OH) radical is the most efficient species in AOPs. Some AOPs, including photolysis (UV) and photochemical (UV/H2O2, UV/O3) reactions, the Fenton reaction (Fe2+/H2O2), photo-Fenton reactions (light/Fe2+/H2O2), cavitation (ultrasonic irradiation), and electrochemical and photocatalysis have been utilized effectively for the oxidation of pollutants [44,45]. Among these, photocatalytic processes have been effectively utilized in a series of oxidation and reduction reactions on the surface of semiconductor materials in the presence of light irradiation. A Web of Science bibliometrics (Figure 1) analysis showed that photocatalytic oxidation was an efficient process among AOPs for the degradation of hazardous pollutants, especially volatile organic compounds (VOCs).
When the semiconductor absorbs the photon at not less than the band gap energy of the semiconductor (Eg), the electrons (e) from the valence band (VB) are excited towards the conduction band (CB), and holes (h+) are left behind in the VB. The photogenerated electrons and holes pairs move into the surface of the semiconductor where they react with surface-adsorbed water or hydroxyl (OH) groups or dissolved oxygen in the reaction medium and produce reactive radical species, i.e., superoxide radical anions (O2•−) and OH radicals. The reactive radical species undergo redox reactions with surface-adsorbed pollutant molecules and completely degrade them. Furthermore, the presence of holes in the VB directly oxidizes the surface-adsorbed pollutants and electrons in the CB, contributing to the indirect oxidation of pollutants by OH radicals generated through the photo-splitting of hydrogen peroxide (H2O2) formed in situ. The corresponding steps involved in the photocatalytic oxidation process using semiconductor materials, and their schematic representation, are shown in Equations (1)–(12) and Figure 2.
Semiconductor + (λ ≥ band gap energy) → Semiconductor (h+) + Semiconductor (e)
h+ + H2O → Semiconductor + H+ + OH
Semiconductor (h+) + -OH → Semiconductor + OH
Semiconductor (e) + O2 → Semiconductor + O2• −
O2• −+ H+ → HO2
HO2 + e → HO2
HO2 + H+ → H2O2
HO2 + HO2 → H2O2 + O2
H2O2 + → 2 OH
H2O2 + Semiconductor (e) → Semiconductor + OH + OH
Semiconductor (h+) + pollutants → CO2 + H2O + other products
OH + O2•− + VOCs → intermediate products → CO2 + H2O + other products
Initially, various metal oxides (TiO2, ZnO, WO3, etc.) and metal sulfide-based (e.g., ZnS, CdS, etc.) semiconductor materials were utilized as catalysts for the photocatalytic oxidation of pollutants, as shown in Table 1. TiO2 is the best among the semiconductor materials as it has the characteristics of chemical and biological stability, low toxicity, high durability, resistance to photocorrosion, high photocatalytic activity, and low cost. Although TiO2 has these merits, it has some disadvantages, such as high bandgap energy (~3.2 eV), poor adsorption capacity, low surface area, and a high recombination rate of photogenerated charge carriers that limit the practical applicability of TiO2 materials. Furthermore, the high bandgap energy restricts its usage under simulated or natural solar light irradiation, and the high recombination rate of charge carriers reduces the photocatalytic degradation efficiency. Approaches such as metal or non-metal doping, coupling with high surface area adsorbents (e.g., activated carbon, graphene oxide, etc.) and other semiconductors (e.g., WO3, g-C3N4, etc.), reduction in size of the semiconductor, changes to the morphology and dye sensitization, etc., have been used to enhance the visible light response and the lifetime of photogenerated charge carriers, consequently improving the photocatalytic degradation efficiency of TiO2 materials. Nevertheless, modified TiO2 materials are limited in their industrial application because of the poor stability (leaching or natural degradation) and reusability of the materials, the high cost of dopants, harmful modifying agents, and poor re-production/re-synthesis of material with similar properties. In order to overcome this, various non-TiO2-based visible-light-responsive semiconductor materials have been developed for the photocatalytic oxidation of various pollutants, including VOCs.
Among non-TiO2 semiconducting materials, bismuth-based semiconducting materials [BiVO4, Bi2WO6, Bi2MoO6, BiOX (X–Cl, Br, I), etc.,] have received much attention because of their narrow band gap energy, excellent visible light absorption, considerable chemical and thermal stability, and suitable band potentials for the generation of reactive radical species [46,47]. BiVO4 is one of the best bismuth-based semiconducting materials due to its low band gap energy (Eg = 2.3–2.4 eV) and high visible light absorption. The discovery of BiVO4 by Kudo et al. [43] for O2 evolution under visible light irradiation has significantly influenced the development of BiVO4-based materials for various applications, including VOCs degradation. Nevertheless, the faster rate of photogenerated electron–hole pair recombination, poor charge carrier transport ability, and the water oxidation kinetics of BiVO4 limit its industrial application. A large number of modifications have been performed to improve the photocatalytic performance of BiVO4 so that commercial requirements would be fulfilled. These are described in the forthcoming sections.

3. Fundamental Aspects of BiVO4 Photocatalyst

BiVO4 is an n-type semiconductor and has been identified as one of the most efficient visible-light-responsive photocatalysts with a band gap energy of 2.4 eV. Naturally, BiVO4 occurs as the pucherite mineral with an orthorhombic crystal structure. However, laboratory-prepared BiVO4 crystallizes either in a scheelite or zircon-type structure. Furthermore, the scheelite structure has a monoclinic and tetragonal crystal system, and the zircon-type structure has a tetragonal crystal system [48]. The crystal structures are shown in Figure 3.
Similarly, the band structures of scheelite and zircon-type BiVO4 are shown in Figure 4. In zircon-type BiVO4, the valence and conduction bands are comprised of O 2p and V 3d orbitals, whereas in the case of scheelite-type BiVO4, the valence band consists of Bi 6s and O 2p orbitals and the conduction band consists of a V 3d orbital. Therefore, the monoclinic (m-BiVO4) scheelite-type system has a relatively smaller band gap energy (2.4 eV) than the tetragonal zircon-type system (2.9 eV); therefore, it shows high visible-light-driven activity [50].
However, the utilization of BiVO4 for catalytic activity is not impressive because it suffers from a high recombination rate of electron–hole pairs and poor charge transport properties. Therefore, the modification of BiVO4 materials, such as by doping of BiVO4 with metals and non-metals, to control its morphology and the synthesis of composite BiVO4 (heterojunction, S-scheme, and Z-scheme) materials has been performed to overcome the abovementioned shortcomings, which are discussed in forthcoming sections.

4. Synthesis Methodologies of BiVO4 Photocatalyst

It is clear that the crystallinity, particle size, and shape of photocatalysts have a significant impact on the photocatalytic activity of a catalyst. Furthermore, it is well known that photocatalytic processes take place on the photocatalyst’s surface. A substantial increase in a photocatalyst’s surface-to-volume ratio will increase its specific surface area, increasing the number of active sites that are available for photocatalytic reactions [50,51]. The morphology and particle size are directly proportional to the greater surface recombination of photogenerated charge carriers. The photocatalyst produces electrons and holes in the diffusion time; the diffusion time of the photocatalyst from the bulk to the surface is represented by Equation (13).
τ = r2/π2 D
where r is the grain radius and D is the diffusion coefficient of the charge carrier. Therefore, when the grain radius decreases, a large number of electrons and holes will travel to the surface for photocatalytic reaction. The synthesis of small and uniform particle sizes plays a vital role in enhancing the photocatalytic activity of semiconducting materials [49,50,51]. The various synthesis methodologies of BiVO4 with different morphologies and their photocatalytic activity and degradation efficiency are summarized in Table 2.

4.1. Hydrothermal Method

Kamble and Ling synthesized truncated square, 18-sided morphological BiVO4 nanomaterials using the hydrothermal method. Figure 5a–d show the different sizes and shapes of m-BiVO4 nanoparticles (NPs). Figure 5c shows the truncated square (18-sided) hexagonal bipyramidal shape with exposed {040} facets exhibiting a strongly revealed surface phenomenon and a facet effect for the visible-light-driven photocatalytic degradation of MB dye [42].
Sun et al. [51] synthesized a BiVO4 nanoplate-stacked star morphological structure via a hydrothermal method using ethylenediamine tetraacetic acid (EDTA). The molar ratio of EDTA to Bi3+ played an important role in the star morphology of BiVO4. The star-like BiVO4 structure showed a higher efficiency for the photodegradation of MB in 25 min under visible light irradiation. The dendritic structure of BiVO4 was fabricated by Lei et al. [52] using an additive-free hydrothermal method at various hydrothermal temperatures, including 100 °C, 140 °C, and 180 °C (Figure 6). The synthesized dendritic structure of BiVO4 was used as a photocatalyst for the degradation of RhB dye. The dendritic BiVO4 synthesized at 140 °C (99.3%) showed superior photocatalytic activity to material synthesized at 100 (58.3%) and 180 °C. The higher activity may be attributed to the high surface area (2.1 m2/g) and crystallinity compared to the other BiVO4 samples.
An olive-like BiVO4 hierarchical morphology via a template-free hydrothermal method was synthesized by Wang et al. [53]. The photocatalytic activities of olive-like BiVO4 was estimated against MB under visible light irradiation. The BiVO4 hierarchical morphology was prepared at different pH values. At a pH of 2.05, the BiVO4 appeared as an olive-like structure but steadily changed into a spherical structure when the pH value varied from 2.05 to 4.02, and at pH 6.00, it adopted a cuboid structure. The authors reported that the olive-like BiVO4 resulted in ∼95.7% degradation of MB within 1 h under visible light conditions.
S. Obregon et al. prepared BiVO4 hierarchical heterostructures using a surfactant-free hydrothermal method [54]. The m-BiVO4 prepared at pH 9 showed needle-like morphology with {110} and {002} planes. The synthesized m-BiVO4 showed good photoactivities for the degradation of methylene blue (MB) under UV–Vis irradiation.
Lu et al. [55] fabricated core–shell-structured (CSS) BiVO4 via a surfactant- and template-free hydrothermal method using bismuth nitrate/ammonium vanadate/ethanol/acetic acid as precursors. They also synthesized different morphologies using PVP and CTAB surfactants. Figure 7a–f show that the shell of the BiVO4 hollow spheres became thinner as the hydrothermal reaction time increased. The BiVO4 plate morphology and biscuit morphology showed ∼88% degradation of RhB at a very slow rate, while the BiVO4 biscuit morphology with a core shell structure showed ∼99% degradation of RhB in 4.5 h.
Meng et al. [56] fabricated various nanoparticles with polyhedral, rod-like, tubular, leaf-like, and spherical morphological structures using a hydrothermal method in the presence of triblock copolymer P123 as a surfactant (Figure 8). Their photocatalytic activities were estimated towards the decomposition of MB dye under visible light stimulation. At different pH = 1, 6, 9, or 10 conditions, the 2D nanoentities were annealed at 400 °C and showed the abovementioned different morphologies. Among the various BiVO4 morphologies, those synthesized hydrothermally with P123 at pH 6 or 10 showed excellent photocatalytic activity due to their greater surface areas and high concentrations of surface oxygen defects. Two-dimensional BiVO4 single-crystal nanosheets were prepared by Zhang et al. [57] with thicknesses of ∼10–40 nm via a hydrothermal route using sodium dodecyl benzene sulfonate (SDBS) as an anionic surfactant. SDBS formed micelles in aqueous solution and enabled the synthesis of BiVO4 nanoparticles with controlled growth. The as-synthesized BiVO4 nanoparticles were used for the photocatalytic degradation of the RhB dye.
A monoclinic BiVO4/sepiolite nanocomposite was fabricated by H. Naing et al. [58]. The nanocomposite BiVO4/sepiolite exhibited excellent visible light photocatalytic performance against antibiotic tetracyclines (TCs) and methylene blue (MB). In the monoclinic BiVO4-30%-sepiolite, fibrous or needle-like sepiolite was distributed on the peanut-shaped monoclinic BiVO4 surface. The photocatalytic efficacies of pure BiVO4, pure sepiolite, and serial monoclinic BiVO4/sepiolite (BVO/S) nanocomposites were studied using the remediation of MB dye and TCs in aqueous solution under visible light irradiation. About 96% of the MB pollutants and 78% of the antibiotic TCs were degraded by BVO-30% S after 4 h of visible light irradiation. A synergistic effect between sepiolite and monoclinic BiVO4 enhanced the separation of the photo-generation carriers, promoting high adsorption, and restrained the regrouping of electron–hole pairs, enhancing photocatalytic activity. Moreover, the hydrophobic nature of sepiolite nanofiber possibly enabled holes generated on the BiVO4/sepiolite nanocomposites to react with pollutants and degrade to smaller molecules.
Chen et al. [59] synthesized snow-like BiVO4 using a cetyltrimethylammonium bromide (CTAB)-assisted hydrothermal method. In the snow-like BiVO4 morphology, oxygen vacancies depended upon the concentration of CTAB. The snow-like BiVO4 morphology coexisted with counter-Br ions, inducing high-concentration surface oxygen defects, which produced more highly reactive oxygen species (ROS), i.e., superoxide and hydroxyl radicals, which resulted in the superfast degradation of ciprofloxacin (CIP).

4.2. Electro-Spinning Method

Various 1D nanostructured BiVO4 materials, such as nano-fibers [60] and micro-ribbon [61], have been synthesized for photocatalysis applications. Cheng et al. [60] prepared BiVO4 porous 1D nanofibers by an electro-spinning method using polyvinyl pyrrolidone (PVP)/acetic acid/ethanol/N, N-dimethylformamide/bismuth nitrate/vanadium (IV) oxy acetylacetonate as a precursor, and the photocatalytic efficiency was estimated towards the photodegradation of RhB dye. The authors reported that the 500 °C calcined BiVO4 sample displayed a higher photocatalytic efficiency for RhB than that for other calcinating temperatures. Liu et al. [61] reported micro-ribbon BiVO4 of an ∼2–3 μm width for the visible-light-driven photocatalytic degradation of MB dye. They also studied the impact of a calcinating temperature of 500 °C on the morphology of BiVO4.

4.3. Solvothermal Method

Red blood cell, flower-like microsphere and dendrite BiVO4 morphologies were prepared by Chen et al. [62] via a facile solvothermal method by adjusting the solution pH and using bismuth nitrate/ammonium vanadate/citric acid/ethylene glycol/ethanol/water as precursors. Figure 9 illustrates the various morphologies and microstructure of the BiVO4 samples using FE-SEM micrographs. Figure 9a,b show the morphology of BiVO4 red blood cell (S-BiVO4), which was achieved using Na2CO3 as a pH-controlling agent. The flower-like microsphere BiVO4 nanoparticles (A-BiVO4) were obtained using NH3·H2O, as shown in Figure 9c,d. Figure 9e shows the BiVO4 dendrite-like morphology (N-BiVO4), which was achieved without the addition of citric acid and Na2CO3 under similar conditions. The BiVO4 red blood cell (S-BiVO4) catalyst exhibited greater catalytic activity than the flower or dendritic morphologies of BiVO4.

4.4. Co-Precipitation Method

Mesoporous monoclinic BiVO4 photocatalysts with different morphologies were prepared by Suwanchawalit et al. [63]. In the synthesis of m-BiVO4, TX100 was used as a surfactant in the co-precipitation method. The TX100 molecules play an important role in the synthesis of BiVO4. An m-BiVO4 structure was prepared by Lai et al. [64] via a precipitation method using a visible light catalyst for the photocatalytic degradation of thiobencarb (TBC). TBC was efficiently degraded by approximately 97% within 5 h. The as-prepared BiVO4 photocatalyst had a polyhedral morphology with a 6–8 μm edge length (Figure 10).

4.5. Sol–Gel Method

The sol–gel method is a promising approach for synthesising metal oxide/mixed oxide composites as this methodology is capable of controlling the morphological and surface properties of the materials at the nanoscale [65]. Min et al. synthesized La- and B-doped BiVO4 photocatalysts using the sol–gel method and utilized them for the photocatalytic degradation of MO dye [66]. Co-doped BiVO4 photocatalysts may enable the synergistic effects of lanthanum and boron to separate the photogenerated holes and electrons in BiVO4 composite. Although all synthesized bismuth vanadate composites have spherical structures, some La-doped composites show a decrease in particle size, and La-doping also inhibits particle growth. In this study, due to La and B doping, BiVO4 (La-B-BiVO4) composites showed a higher photocatalytic degradation of MO dye in 60 min than BiVO4 and B-BiVO4, and the specific surface area and surfaces for oxygen vacancies were also enhanced, reducing the crystallite size, and also reducing the band gap and the intensity of absorbed light in the visible region.
Mousavi-Kamazani synthesized composite nanostructures of copper oxides and bismuth (Cu/Cu2O/BiVO4/Bi7VO13) using the Pechini sol–gel method [67]. By altering the reaction conditions, different morphological structures were synthesised. The addition of ethylenediamine as a gelling agent, tannic acid as a chelating agent, and a 1:1:1 ratio for Cu:Bi:V enabled a rectangular cube-like morphology to be formed. When the gelling agent changed to polyethene glycol instead of ethylenediamine, plate-like microstructures were formed. By changing the chelating agent from tannic acid to fumaric acid, a pseudo-spherical morphology was obtained. In the absence of Cu and 0.5 mole Cu, platelike nanostructures, nanorods, and quasi-spherical structures were observed, respectively. Moreover, the composite structure (Cu/Cu2O/BiVO4/Bi7VO13)) with rectangular cube-like morphology (size of about 30–100 nm) exhibited excellent photocatalytic oxidative desulfurization of the oil derivatives under visible light (92%) than BiVO4 and Cu2V4O11. The results suggest that the addition of Cu and Cu2O species into the composites increases the electrical conductivity, capable of electron–hole separation, and alters the morphology and also particle size, which might be the reason for the enhanced photocatalytic activity.
Castaneda et al. synthesized BiVO4/TiO2 nanocomposites with different compositions (BiVO4: TiO2 = 1:0.6, 1:2.5, and 1:10) using a modified one-step sol–gel method [68]. The SEM studied before and after calcination shows diversity in the particle size and exhibits a spherical shape. The BiVO4/TiO2 nanocomposite with a mass ratio of (1:10) shows the highest photocatalytic efficiency compared to the other compositions and (photo) electrochemical responses for the degradation of azo dyes (Acid Blue-113, AB-113) (~99%) under visible light radiation.

5. Engineering/Modification Processes of BiVO4 Properties

As discussed earlier, BiVO4 is one of the promising photocatalysts for many practical applications ranging from water treatment, the removal of dyes and organic pollutants, H2 generation, cargo and biomedical deliveries, etc. [69,70]. The better photocatalytic performance of BiVO4 is accredited to their visible-light-responsive band gap (2.4 eV), layered structure, suitable valance band maximum, chemical stability, and nontoxicity. BiVO4 has four different polymorphic forms, including orthorhombic, zircon-tetragonal, monoclinic (m), and tetragonal (t) BiVO4 [71]. Among them, m-BiVO4 is more active for photo-related applications. Generally, a low temperature reaction yields zircon-tetragonal BiVO4 phase, which can be transformed into m-BiVO4 by inducing calcination reactions and reversibly back to t-BiVO4 phase at a temperature of 528° K. The band gaps energies of t-BiVO4 and m-BiVO4 were reported to be indirect of 2.3 eV and 2.4 eV, respectively [72]. Previous experimental studies with BiVO4 have some shortcomings, such as poor charge carrier transfer (bulk carrier mobility of 0.05–0.2 cm2 V−1s−1) and charge recombination before being captured by targeted molecules for photochemical reactions [73]. To overcome these issues, several strategies, for example, the nano-scaling [74], morphology engineering [75], crystal facet control [76], and crystal structure control [77] of BiVO4, have been demonstrated to improve the optical and electronic properties to some extent, resulting in high photocatalytic performances towards organic pollutant degradation.
Many recent studies have also proposed multiple advantages of doping and mixed-phase BiVO4 systems over single-phase BiVO4 photocatalysts [78,79,80]. This resulted in the extended light absorption capability, carrier mobility, and higher efficiency of BiVO4-based systems for photochemical reactions. This is mainly attributed to the effect of donor defects or through adding excess electrons to the BiVO4 model system through doping with metals (Mo, W, Sn) [69,81,82], nonmetals (S, F), and the creation of heterostructures with other semiconductors. In this section, an overview of different strategies such as (1) metal doping, (2) noble metal doping, (3) nanocomposite structures, (4) composite with carbon analogs, and (5) heterostructures related to BiVO4 photocatalyst have been briefly discussed.

5.1. Metal/Nonmetal-Doped BiVO4

Metal doping is one of the conventional and effective strategies for modifying the electronic properties of semiconductors, i.e., p- and n-type conductivity behaviors. In semiconductor photocatalysis, doping with metal/nonmetal can improve the charge carrier separation, tune the band gap energy, and enhance visible light absorption. For example, Liu et al. [83] synthesized Mo-doped BiVO4 using an electrospun method and studied its morphology, crystal structure, and optoelectronic properties. A stoichiometric amount of bismuth and vanadium solutions were prepared, and varied amounts of ammonium molybdate (0.4, 1, 1.5, 3 mol%) were incorporated and stirred (12 h) to obtain homogeneous solution. Then, an electrospinning reaction was performed at a temperature of 60 °C and 15 kV voltages. According to the formation mechanism (Figure 11A(i)), pure BiVO4 can form homogenous, well-dispersed particles together with many pores. When a small amount of Mo (1%) was incorporated into BiVO4, the particle size increased, and the existing pores disappeared (Figure 11A(iii–v)). When further increasing the Mo content (3%), the BiVO4 particle size increased and Mo saturated, forming secondary-phase MoO3 on the surface of BiVO4. As a result, photocatalytic tests revealed that 1% Mo-BiVO4 shows excellent photoactivity ~ three times higher than reference BiVO4 (Figure 11A(ii)), indicating that a small content of Mo dopant is crucial to improving the separation of charge carriers and electronic conductivity.
On the contrary, a higher Mo content led to crystal phase transformation from monoclinic to tetragonal together with secondary-phase formation, which might act as recombination centers causing a reduction in the photocatalytic activity. A theoretical study by Zhang et al. [84] purports the effects of Mo/W co-doping for the photocatalytic activity of monoclinic BiVO4. They found that Mo or W atom doping preferably occurs at the V site to generate continuum states directly above the conduction band (CB) level of BiVO4, and this decreases the band gap, which is beneficial for photochemical reactions. Particularly, they found that W-doped BiVO4 exhibits a smaller band gap than the Mo-doped BiVO4, and the electronic properties of BiVO4 are quite different. Additionally, Mo/W/Mo and W/Mo/W co-doping in BiVO4 requires low formation energies and reduced bandgaps compared to other doping systems, which may extend the light absorption and could be more suitable for visible-light-driven photocatalysis. Yao et al. [85] fabricated Mo-doped BiVO4 via a solid-state reaction by grinding stoichiometric amounts of Bi2O3, V2O5, and MoO3 in an agate mortar and heating at 600◦C for 5 h, followed by calcination at 800 °C for 2 h. The Mo-doped BiVO4 showed much more photocatalytic activity for water oxidation and MB dye degradation compared to pure BiVO4, due to the higher surface acidity of Mo-doped BiVO4 (2 atom%) stemmed from the existence of Lewis and Brønsted acidic sites associated with Mo6+ doping, which offer greater adsorption feasibility for water molecules and organics contaminants. Gao et al. [86] reported the synthesis of Ni-doped BiVO4 and Z-scheme BiVO4-Ni/AgVO3 nanofibers using a strategy which combined an electrospinning and hydrothermal strategy (Figure 11B(i)). First, a nanofibrous Ni-doped BiVO4 was obtained using an electrospinning precursor solution of Bi(NO3)3⋅5H2O, Vo(acac)2, and Ni(NO3)2⋅6H2O (prepared in mixed solvent of DMF, CH3COOH, and CH3CH2OH). PVP was used a matrix during the electrospinning process. Next, hydrothermally grown AgVO3 on the surface of Ni-doped BiVO4 forms a Z-scheme heterojunction of BiVO4-Ni/AgVO3. Due to the synergetic effect of Ni doping and AgVO3 assembly, the specific surface area and light absorption ability of BiVO4-Ni/AgVO3 was significantly improved compared to BiVO4. Ni-doping adds impurity energy levels and replaces V sites on the {121} plane (Figure 11B(ii)), producing the structural distortion of tetrahedral VO43+, which can be confirmed from a right shift in diffraction peaks [87]. The lowering of the diffraction peak intensity also suggests successful Ni doping, which leads to a reduction in crystallinity. A Z-scheme optimal BiVO4-Ni-1/AgVO3-25 photocatalyst showed superior photocatalytic Cr6+ reduction efficiency (99.7%) in 80 min. The comparison of apparent rate constants for Cr6+ reduction over different photocatalysts is shown in Figure 11B(iii). The inset shows the HRTEM image of BiVO4-Ni/AgVO3 and the presence of three different phases.
Figure 11. (A) (i) Schematic showing the electrospinning synthesis strategy for Mo-doped BiVO4 nanostructures. (ii) Photocatalytic degradation of MB using Mo−BiVO4 photocatalyst under light irradiation. (iiiv) TEM image of pure BiVO4, 1% Mo-BiVO4, and 3% Mo-BiVO4 (reproduced with permission from Ref. [83]). (B) (i) Synthetic protocol for nanofibrous BiVO4-Ni/AgVO3 [86], (ii) schematic representation of possible doping sites of Ni in BiVO4, (iii) photocatalytic reduction of Cr6+ over different photocatalysts and their apparent rate constants and inset shows the TEM image of BV-Ni-1/AV−25. Reproduced with permission from Ref. [87].
Figure 11. (A) (i) Schematic showing the electrospinning synthesis strategy for Mo-doped BiVO4 nanostructures. (ii) Photocatalytic degradation of MB using Mo−BiVO4 photocatalyst under light irradiation. (iiiv) TEM image of pure BiVO4, 1% Mo-BiVO4, and 3% Mo-BiVO4 (reproduced with permission from Ref. [83]). (B) (i) Synthetic protocol for nanofibrous BiVO4-Ni/AgVO3 [86], (ii) schematic representation of possible doping sites of Ni in BiVO4, (iii) photocatalytic reduction of Cr6+ over different photocatalysts and their apparent rate constants and inset shows the TEM image of BV-Ni-1/AV−25. Reproduced with permission from Ref. [87].
Nanomaterials 13 01528 g011
Bashir et al. [88] fabricated Gd-doped BiVO4 using a simple hydrothermal method. The resulting Gd-doped BiVO4 ultrasonically treated with rGO to form Gd-doped BiVO4/rGO of the nanocomposite nanostructures (Figure 12A(i)). According to SEM analysis, Gd doping does not change the surface morphology. As shown in TEM images (Figure 12A(ii,iii)), three different phases of components promote significantly improved electron/hole pair separation, excellent photocatalytic MB degradation efficiency (97%) than BiVO4 (53%), and Gd/BiVO4 (69%) within 100 min (Figure 12A(iv)). The higher photoactivity of rGO/Gd/BiVO4 is due to a heterojunction effect between Gd/BiVO4 and rGO sheets, which not only enhances the light absorption but also enlarges the surface area in the presence of rGO.
Unlike transition metals, doping with a rare earth metal is also found to improve the photocatalytic properties; however, this finding is rarely reported [89]. These metal ions possess excellent luminescence properties and therefore endow several benefits such as light absorption, modified surfaces, and acting as electron traps that can help minimize the recombination of photoinduced charge carriers [90]. Moscow et al. [91] reported erbium (Er) and yttrium (Y)-doped BiVO4 using a simple microwave-assisted approach (Figure 12B(i)). In synthesis, Bi and V precursors were dissolved under magnetic stirring. Later, different amounts of Er and Y precursors were introduced, followed by microwave irradiation forming Er3+- and Y3+-doped BiVO4 nanostructures. According to SEM and XRD results, Er3+ and Y3+ doping led to a reduction in the particle sizes of BiVO4 and formed mixed-phase BiVO4. Raman spectra analysis revealed (V–O) band shift from 820 cm−1 to 850 cm−1 and disappeared δ (VO4+) doublet due to the conversion of monoclinic BiVO4 to tetragonal BiVO4 phase (Figure 12B(ii)). Photocatalytic tests suggested that Y-doped BiVO4 reportedly has the highest degradation efficiencies of 93%, 85%, and 91% for MB, MO, and RhB dyes, respectively, in 180 min under light irradiation. The possible photocatalytic electron transfer mechanism is shown in Figure 12B(iii). A photocatalytic degradation of acetaldehyde was also achieved at an impressive rate using Y-doped BiVO4. Because of the formation of the inner energy state Er3+ and Y3+ metal, the band gap reduced, light absorption extended, and the recombination of electron–hole pairs suppressed. Next, the co-doping of Gd and Y into BiVO4 was successfully accomplished by simple hydrothermal synthesis [92]. Upon sunlight illumination (90 min), a Bi0.92Gd0.07Y0.01VO4 photocatalyst exhibited 94% degradation efficiency for methylene blue dye (MB), which is four times larger than pure BiVO4. Recently, Sudrajat and Hartuti [93] used a one-step hydrothermal preparation method to prepare B-doped BiVO4 (B-BiVO4) with an oval-shaped morphological structure. The B dopant acts as mid-gap-state electron donors, allowing more excitations of the band gap to be produced and the conduction band of BiVO4. The light-induced infrared absorption measurement confirmed that there were more electrons available for reduction reactions and more holes available for oxidation reactions, leading to greater photocatalytic activity for the mineralization of phenoxyacetic acid (PAA) in the presence of simulated sunlight.
Figure 12. (A) Gd-doped BiVO4. (i) Schematic illustration of the synthesis strategy, (ii) TEM images of rGO/Gd/BiVO4, (iii) HRTEM image of Gd/BiVO4, and (iv) comparison of dye removal efficiency with different photocatalysts (reprinted with permission from Ref. [88]). (B) Y- and Er-doped BiVO4. (i) Schematic representation of microwave-assisted synthesis, (ii) Raman spectra of BiVO4, Y-BiVO4, and Er-BiVO4, and (iii) schematic showing various dye degradation using Y−BiVO4 photocatalyst with sunlight. Reprinted with permission from Ref. [91].
Figure 12. (A) Gd-doped BiVO4. (i) Schematic illustration of the synthesis strategy, (ii) TEM images of rGO/Gd/BiVO4, (iii) HRTEM image of Gd/BiVO4, and (iv) comparison of dye removal efficiency with different photocatalysts (reprinted with permission from Ref. [88]). (B) Y- and Er-doped BiVO4. (i) Schematic representation of microwave-assisted synthesis, (ii) Raman spectra of BiVO4, Y-BiVO4, and Er-BiVO4, and (iii) schematic showing various dye degradation using Y−BiVO4 photocatalyst with sunlight. Reprinted with permission from Ref. [91].
Nanomaterials 13 01528 g012
Similarly, doping with nonmetals such as nitrogen (N) and fluorine (F) in BiVO4 has been adopted to promote light absorption, band gap tuning, and catalytically active surfaces. For instance, Wang et al. [94] prepared N-doped BiVO4 using a sol–gel technique with hexamethylene tetramine (C6H12N4) as a N source. N-doping was found to not change the morphology and surface area of BiVO4 significantly. However, N was found to be doped into crystal lattice O–Bi–N–V–O bonds, creating highly active V4+ species, oxygen vacancies, and a red shift in the absorption band. The photoactivity of BiVO4 in this system mainly depends on two factors: (i) the content N-doping and (ii) heat treatment temperature. The N-BiVO4 calcined at a temperature of 500 °C showed the highest activity. Li et al. [95] reported the synthesis of F-doped BiVO4 nanospheres via a simple two-step hydrothermal method. NaF was employed as the fluoride source which helped to modify the crystal lattice of BiVO4, suppressing the charge carrier recombination, resulting in high photoactivity. Wang et al. [96] synthesized B-doped-BiVO4 photocatalysts using a CS-template-assisted sol–gel method. According to these authors, B doping can form a monoclinic crystal structure, large surface areas, a smaller band gap value, and higher V4+ species. This results in the best photoactivity of 0.04B CS-BiVO4 for the degradation of MO dye, which is not only because of B doping but also due to the cellular morphology, which stems from the template as a similar 0.04B-BiVO4 sample prepared without a template showed much lower photoactivity.

5.2. Noble-Metal-Doped BiVO4 as Photocatalyst

Noble metals (Au, Ag, Pt) are very attractive in photocatalysis because of their surface plasmon resonance (LSPR), which can provoke the oscillation of conduction band electrons and plasmonic energy transfer. This mainly occurs via two different mechanisms: (i) direct electron transfer and (ii) plasmon-directed resonant transfer of energy. Cao et al. [97] have reported the synthesis of Au-BiVO4 nanosheets using hydrothermal and a cysteine-linking strategy. Au precursor in the presence of cysteine evolves Au-doped BiVO4, as shown in Figure 13A(i). Interestingly, the surface plasmon resonance (SPR) of Au enables excellent visible-light-driven photocatalytic activity related to pure BiVO4 for the degradation of MO dye (Figure 13A(ii)). In this system, Au acts as an electron sink retarding the recombination of photoinduced electrons and holes. The contribution from Au nanoparticles was clarified by comparing the experimental results with Pt-BiVO4 (prepared through a similar strategy), showing no SPR in the range of (500 ± 20 nm) as Au-BiVO4 under visible light illumination. Moreover, electron trapping on Au raises the Fermi level (Ef) of Au to more negative potentials (Ef*), leading to band alignment for an effective charge transfer. In addition, the photogenerated electrons transfer to Au nanoparticles can reduce the adsorbed sacrificial agent S2O82− to SO42− on the Au surface. As a result, the holes remain on the BiVO4 surface and have a considerably higher lifetime to perform the water oxidation process (Figure 13A(iii)). Reddy et al. [98] reported Au-doped BiVO4 photocatalyst synthesis via a sonication and calcination method and applied it as electrodes for water splitting and electrochemical storage. This study suggests that, due to Au doping, the photocurrent density increased 25 times related to reference BiVO4, demonstrating the synergistic role of Au, while BiVO4 increased the electrical conductivity and charge transfer at the interface. The synergistic effects of Ag nanoparticles (LSPR) and N-doped graphene (upconversion effect) with BiVO4 have also been reported for the degradation of tetracycline hydrochloride (TC•HCl), as shown in Figure 13B(i) [99]. A ternary photocatalyst N-GNDs/Ag/BiVO4 was prepared through the solvothermal and hydrothermal process, and it exhibits distinct behavior from reference BiVO4. According to the experimental results, possible charge transfer mechanisms using N-GNDs/Ag/BiVO4 have been schematically proposed (Figure 13B(ii)). Under light illumination, N-GNDs can absorb NIR light and contribute to the light upconversion phenomenon, which can help to extend light absorption. Since the potential of O2/O2 (−0.33 eV, NHE) is a higher negative value than the conduction band of BiVO4 (+0.47 eV, NHE), the conduction band electron of BiVO4 cannot reduce O2 to produce O2 radicals. As a result, the photoinduced electron from BiVO4 migrates to N-GNDs and Ag-NPs, leading to band alignment. Thus, a Schottky barrier is formed at the interface to facilitate more efficient electron transfer; N-GNDs serve as electron acceptors to capture photoexcited electrons from BiVO4, and hot electrons from Ag-NPs (LSPR effect) reduce O2 to O2, which later oxidizes TC•HCl to smaller molecules.

5.3. BiVO4-Based Nanocomposites

Single-phase BiVO4 is insufficient to achieve higher photocatalytic efficiency due to the poor electron transfer and low light absorption ability (band gap 2.4 eV). Recently, a concept of nanocomposite fabrication has been adopted, in which BiVO4 is coupled with other metals such as Au [98], Ag [100], Co [101], semiconductors TiO2 [102], ZnO [103], CdS [104], WO3 [105], MnO2 [106], Ag3VO4 [107], and carbon analogues graphene [99,108] g-C3N4 [109], carbon nanotubes [110], etc., enabling band gap alignment and efficient electron transfer across interfaces. The resulting composites could improve the overall efficiency by modifying the chemical, physical, and optical properties of BiVO4 and charge carriers and light absorption.
Wei et al. [111] fabricated N-doped Biochar (N-Biochar)@BiVO4 nanocomposite via an easy hydrothermal method and evaluated its photocatalytic performance for the degradation of triclosan (TCS). Figure 14A(i,ii) shows the fern-like morphology of BiVO4, while the composite shows well-dispersed N-biochar in intimate contact with BiVO4. Upon light irradiation (60 min), the BiVO4@N-Biochar catalysts showed 94.6% TCS degradation efficiency, which is much higher than pure BiVO4 (56.7%). As per LSMS and the E. coli (Escherichia coli) colony assessment studies, a detoxification efficiency of 72.3 ± 2.6% was determined, signifying a remarkable reduction in biotoxicity during photodegradation. The schematic representation of the possible charge transfer mechanism on BiVO4@N-Biochar is illustrated in Figure 14A(iii). Cao et al. [112] have reported the synthesis of Al-doped BiVO4 composites with the use of a simple hydrothermal method and evaluated for photocatalytic decomposition of MB dye (Figure 14B(i)). Different molar ratios of Bi to Al were used and calcined at a temperature of 500 °C under Ar gaseous environment (1h) to obtain Al-doped BiVO4. Based on optical spectroscopy, the band gap energies of pristine BiVO4, Al-0.03-BV, and Al-0.3-BV samples were determined to be 2.36, 2.40, and 2.41 eV, respectively. As shown in Figure 14B(ii), when isopropanol and potassium dichromate were used as scavengers, the degrading efficiency for MB was decreased, demonstrating that e and ·OH radicals are the active species in the degradation of dye molecules. Their experimental results confirm that optimal 30 mol% Al-BiVO4 (Al-0.3-BV) showed excellent photocatalytic activity for MB degradation due to the synergistic effect of appropriate Al doping and Al2O3 surface passivation. Based on transient photovoltage (TPV) and surface photocurrent (SPC) results, the coexistence of Al3+ and Al2O3 evolved, causing a synergistic effect for advancing e transfer and extending the lifetime of charge carriers. Al doping can result in transforming the surface morphology of BiVO4, as the polyhedron structure of BiVO4 becomes thinner and there is a reduced grain size (Figure 14B(iii–v)). Similarly, Wetchakun et al. [113] studied BiVO4/CeO2 nanocomposites through the co-precipitation and hydrothermal method. The different molar concentration of semiconductors’ constituent was fixed and evaluated for the degradation of dyes pollutants in water. The XRD results suggest that two different kinds of diffraction peaks confirmed the coexistence of mixed-phase, indicating BiVO4/CeO2 nanocomposite formation (Figure 14C(i)). Under light irradiation (>400 nm), the molar ratio 0.6:0.4 for BiVO4/CeO2 nanocomposite displayed the highest photocatalytic degradation activity for the removal of MB dye in water (Figure 14C(ii)).

5.3.1. Activated Carbon, Carbon, and Other Adsorbents-Based Composites

Patil et al. [73] reported the synthesis of BiVO4/Ag/rGO hybrid architectures using a cost-effective hydrothermal method exhibiting impressive reaction rates for water oxidation and organic pollutant degradation reactions. Fern-like BiVO4 nanostructures were prepared and decorated on the surface of reduced graphene oxide (rGO) sheets (Figure 15A(i)), which can offer a large surface area and excellent electron transfer properties. They discovered that Ag nanoparticles can be reduced during hydrothermal reaction and deposited on the surface of BiVO4, forming a Schottky junction between Ag and BiVO4. Figure 15A(iii,iv)) depicts SEM and TEM images of nanocomposites, demonstrating different phases of Ag, rGO, and BiVO4. As shown in Figure 15A(ii), a complete degradation of MB dye within 120 min was achieved using a hybrid BiVO4/Ag 2%/rGO catalyst under simulated light irradiation, which is ~2.18 and ~1.25 times larger than pure BiVO4 and BiVO4/Ag photocatalysts. On the basis of PL and PEC results, the highest photocatalytic performance in this system is attributed to the combined effect of Ag and rGO, enabling the efficient promotion of e/h+ separation across the interface and visible light absorption. The effect of graphene oxide on BiVO4 photocatalyst was also demonstrated by Zhang et al. by incorporating graphene oxide (GO) between the BiVO4 and NiOOH oxygen evolution catalysts (OEC). The results indicate that GO served as hole extraction layer due to its hole storage capability and improved the stability of the material. Meanwhile, GO employs the formation of p/n heterointerface with BiVO4 and encouraged the hole transfer from BiVO4 to NiOOH [114].
Graphitic carbon nitride (g-C3N4) is one of the best 2D semiconductor materials due to its large surface area and intriguing electronic properties. Incorporating g-C3N4 with BiVO4 can significantly modify the physicochemical properties and showed impressive photoactivity towards the degradation of organic pollutant dyes, CO2 reduction, and H2 generation reactions. Alhaddad et al. [115] reported a g-C3N4-incorporated Pt@BiVO4 nanocomposites catalyst for the detoxification of ciprofloxacin. A sol–gel reaction between Bi (NO3)3.5H2O and NH4VO3 was adopted using CH3COOH and HCl, forming BiVO4 nanoparticles. Then, using C6H14 as a solvent, solid dispersions of BiVO4 in different mass contents of 1.0, 2.0, 3.0, and 4.0 wt% and g-C3N4 were prepared, where it was agitated for 4h, resulting in the formation of heterostructures. At last, solid dispersions were then prepared via photoreduction. As shown in the TEM image (Figure 15B(i)), Pt nanoparticles can be photoproduced and deposited on the surface of BiVO4 and g-C3N4. A comparison study revealed that the 0.5 wt% Pt@4 wt% BiVO4-g-C3N4 heterojunction is optimal, displaying 5.0- and 3.7-times higher photocatalytic efficiency than pure g-C3N4 and BiVO4 for the decomposition of ciprofloxacin (Figure 15B(ii)). The plausible electron transfer mechanism over Pt@BiVO4-g-C3N4 during the photocatalytic removal of ciprofloxacin is illustrated in Figure 15B(iii). The creation of the p-n heterojunction can facilitate charge carrier separation, while Pt nanoparticles further assist in the increase in the light absorption and photocatalytic efficiency.

5.3.2. Heterojunction Construction

Interestingly, monoclinic (m) BiVO4 and tetragonal zircon (tz) BiVO4 can form a heterojunction. For example, Dabodiya et al. [116] prepared mixed-phase BiVO4 (m:tz-60:40) using a microwave–hydrothermal method, which displays a 95% degradation efficiency for Rhodamine B dye. As shown in Figure 16A(ii), the effect of phase transition from tz-BiVO4 to m-BiVO4 was investigated in terms of its photocatalytic efficiency for the decomposition of RhB dye. On the basis of UV-reflectance and PL results, they discovered a reduction in the bandgap energy and facilitated e-/h+ separation at the m-BiVO4/tz-BiVO4 interfaces and enhanced photoactivity under visible light irradiation. Similarly, Patil et al. [117] reported the controlled synthesis of BiVO4 (pillars-like, dendrite-like, and microgranule-like) and the m-BiVO4/tz-BiVO4 heterojunction using simple a hydrothermal–solvothermal and solid-state reaction (Figure 16B(i)). The mixed-phase m-BiVO4/tz-BiVO4 heterojunction prepared through solvothermal reaction displayed the highest photodegradation efficiency of 95% for MB dye related to single-phase BiVO4 prepared through a hydrothermal (BVO-HDR; 79%) and solid-state reaction (BVO-SSR; 88%). Experimental results confirmed that temperature plays a critical role in the phase transformation of BiVO4. On the basis of PEC and EIS results, the high photocurrent density and reduction in internal resistance is confirmed, demonstrating that special dendritic architectures and the heterojunction effect is crucial to promote the e-/h+ separation and utilization. Figure 16B(ii) shows the possible electron transfer mechanism over the single-phase BiVO4 and m-BiVO4/tz-BiVO4 heterojunction for the photocatalytic degradation of organic dyes under light irradiation.
Recently, Li et al. [103] developed a ZnO/BiVO4 heterojunction thin films catalyst using a simple chemical bath deposition and electrodeposition method. The as-formed three-dimension choral-like ZnO/BiVO4 displayed an excellent photoelectrocatalytic tetracycline degradation efficiency of 84.5%. Figure 17A(i,ii) shows the schematic representation of the BiVO4/ZnO electrode and charge transfer process during photoelectrocatalysis. According to the radical scavenger’s test, •O2 and •OH were found to be major active species responsible for tetracycline degradation. As shown in Figure 17A(iii), ZnO/BiVO4 showed a tetracycline degradation efficiency 84.5%.
Yan et al. [118] fabricated the BiVO4/Ag3VO4 heterojunction using a simple hydrothermal and coprecipitation method and evaluated its photocatalytic performance towards RhB dye degradation. Figure 17B(i) shows the TEM image of hybrid BiVO4/Ag3VO4 and two sets of lattice fringes, indicating an intimate interface between the two semiconductor and heterojunction creation. As shown in Figure 17(Bii), the difference in the molar ratio showed varied photocatalytic activity. Among them, a 10:1 mol ratio of the BiVO4:Ag3VO4 heterojunction displayed the highest 95% degradation efficiency for RhB dye, which is 10- and 3.4-times higher than pure BiVO4 and Ag3VO4, respectively. The schematic representation of the photocatalytic reaction mechanism is illustrated in Figure 17B(iii). On the basis of electrochemical impedance analysis (Figure 17B(iv)), a small semicircle for BiVO4:Ag3VO4 suggests a lowest internal resistance and faster electron transfer process due to the heterojunction effect and the effective separation of the photo-induced charges carriers.
Bao et al. [119] designed a facet–heterojunction Z-scheme photocatalyst AgBr-Ag-BiVO4 {010} to increase the photoactivity of BiVO4 for the inactivation of pathogenic bacteria and the degradation of organic dyes from wastewater. First, facet-controlled BiVO4 {010} was obtained using the facile hydrothermal method, and then Ag nanoparticles were deposited through photoreduction, while in situ chemical treatment in the presence of KBr and Fe(NO3)3 enabled the transformation of the outermost layer of Ag nanoparticles to AgBr. Interestingly, Ag nanoparticles can be selectively deposited on the BiVO4 {010} facets and transform. Figure 18i shows the FESEM image of the AgBr-Ag-BiVO4 {010} heterojunction and highly dispersed AgBr nanoparticles, which showed an increase in particle sizes upon transformation. AgBr-Ag-BiVO4 {010} displayed the highest photocatalytic inactivation for Escherichia coliK-12, which is ∼4 and 15 times the reference Ag-BiVO4 {010} and BiVO4, respectively (Figure 18iii). The possible charge transfer mechanism during the photocatalytic inactivation of bacterium is illustrated in Figure 18ii. The photoluminescence (PL) spectroscopy and PEC results suggested a suppression in the charge recombination upon the heterojunction, and the electron paramagnetic resonance (EPR) results revealed that h+, OH, and O2, are the major active species for the degradation of RhB dyes in wastewater.

5.3.3. BiVO4-Based S-Scheme and Z-Scheme Nanocomposite Materials

As mentioned above, it is difficult for single BiVO4 semiconducting materials to offer a strong visible light response and high redox capability simultaneously. So, the coupling of BiVO4 with another suitable semiconducting material to synthesize S-scheme- and Z-scheme-type composite materials has received significant attention. The band structure of both the S-scheme- and Z-scheme-type composite materials is shown in Figure 19.
In both composite materials, the high CB semiconductor materials combine with high VB semiconducting materials, whereas they follow two different pathways for photo-generated electron–hole pairs separation and transfer. In Z-scheme-type materials, the photo-generated electron–hole pairs are separated by an internal electric field between the two semiconductor interfaces, whereas in the case of S-scheme-type materials, the charge carrier separation and transfer occur through an internal electric field, energy band bending, and Coulomb gravity. Furthermore, there are two types of Z-scheme-type materials, such as direct Z-scheme-type materials and mediator (indirect)-based Z-scheme-type materials. In indirect Z-scheme-type materials, metals and non-metals, carbon materials, and quantum dots, etc., are used as mediators. For example, Li et al. coupled high-valence-band-edge BiVO4 with high-conduction-band-edge g-C3N4 material through a wet impregnation-calcination method which yielded Z-scheme BiVO4/g-C3N4 materials. The calculated CB and VB potentials of g-C3N4 and BiVO4 are 1.20 and 1.54 eV and 0.46 and 2.86 eV, respectively. Under visible light irradiation, the electron–hole pairs are generated in both the semiconductors. Subsequently, the electrons present in the CB of BiVO4 combines with holes present in the VB of g-C3N4, so the electrons in the CB of g-C3N4 and holes in the VB of BiVO4 are efficiently separated and possess a higher potential than the generation potential of the reactive radicals (OH and O2•–). Consequently, a higher concentration of reactive radicals is generated, which showed higher photocatalytic activity in the degradation of malachite green (MG) dye in the presence of visible light irradiation and H2O2 [120]. Similarly, Hu et al. developed a g-C3N4/BiVO4-based S-scheme system using hydrothermal methods for the degradation of paraben preservative in the presence of visible light and natural solar light irradiation. The CB and CB of g-C3N4 were located at–1.3 and + 1.44 V vs. RHE at pH 0, respectively, and the CB and VB of BiVO4 were sited at + 0.09 and + 2.4 V vs. RHE, respectively. The energy difference present in the mixed g-C3N4/BiVO4 system would allow the transfer of electrons of g-C3N4 to BiVO4, which leads to the positively charged region on the g-C3N4 side and negatively charged region on the BiVO4 side. Therefore, there is a generation of an inner electric field at the interface of g-C3N4/BiVO4, with the direction from BiVO4 to g-C3N4. Under irradiation, the photoexcited electrons on the CB of BiVO4 combined with the holes on the VB of g-C3N4 lead to the efficient separation of electrons and holes on the CB of g-C3N4 and on the VB of BiVO4 for the production of reactive radicals, which could participate in the degradation of paraben [121]. Similarly, various researchers have developed BiVO4-based Z-scheme and S-scheme materials for the degradation of dyes and other compounds; however, this has been scarcely studied for the application of VOC degradation.

6. Volatile Organic Compounds Degradation Application

As described in the above sections, volatile organic compounds (VOCs) pose a serious threat to environment and human health. VOCs are mainly BTEX (benzene, toluene, ethylbenzene, xylene), acetylene, acetone, ethylene, trichloroethylene, benzaldehyde, acetaldehyde, isopropanol, hexane, etc., and are mainly released by human activities through outdoor and indoor sources. VOCs create significant health issues; they specifically cause allergies, cancer, and they slow down and damage the nervous and respiratory system. Therefore, significant research activities are put forward for controlling and degradation of VOCs. Among those, the photocatalytic oxidation/degradation of VOCs has received great attention because of the simple operation and reaction conditions, low cost, and the fact that it completely degrades VOCs, and renewable solar energy can be used for their degradation. Among photocatalysts, BiVO4 is a better visible-light-responsive system for VOCs degradation, which are described in this section and Table 3.
Hu et. al. synthesized a BiVO4/TiO2 heterojunction photocatalyst using the sol–gel method and evaluated the photocatalytic oxidation of gaseous benzene under UV light and simulated solar light irradiation. The BiVO4 with a loading percentage of 0.5% presented a higher photocatalytic oxidation of benzene (66.8% conversion, Figure 20a) and a high amount of CO2 production (Figure 20b) compared to the other percentage loaded materials and bare TiO2 and BiVO4 materials. The improved visible light absorption by the introduction of BiVO4 and the activated species (•OH) are responsible for the high activity of the BiVO4/TiO2 heterojunction photocatalyst [123]. Furthermore, Zhao et al. developed CuO/BiVO4 hollow nanospheres using the sol–gel method followed by the impregnation method and demonstrated visible light photocatalytic activity by the degradation of gaseous toluene. The results revealed that 5% CuO-loaded BiVO4 hollow nanospheres showed higher visible light photocatalytic activity (85%) in toluene degradation compared to other percentage-loaded composites and bare materials [124]. Chen et al. hydrothermally synthesized BiVO4/α-Fe2O3 composites and calcined them at various temperatures (250, 350, 450, and 550 °C) and evaluated their efficiency by the degradation of benzene under UV light irradiation. The composites calcined at 350 °C exhibited higher benzene removal efficiency (66.87%) than other composites and bare α-FeOOH, Fe2O3, and BiVO4 materials [131]. However, still the binary composites showed low degradation efficiency due to the low separation and transfer rate of photogenerated charge carriers. In order to further enhance the photocatalytic efficiency of BiVO4, BiVO4-based Z-scheme and ternary nanocomposites were prepared, and also metal nanoparticles were introduced into BiVO4. For example, a coral-like direct Z-scheme BiVO4/g-C3N4 was synthesized by Sun et al. for the degradation of toluene under visible light irradiation. The visible light absorption of BiVO4 was significantly increased after g-C3N4 loading and also promoted the separation of photogenerated electron hole pairs. The improved separation of photogenerated electron hole pairs on the direct Z-scheme BiVO4/g-C3N4 material led to a higher toluene degradation efficiency compared to bare materials [130]. Furthermore, Li et al. studied automobile exhaust gas purification by improving the adsorption capacity of volatile compounds (automobile exhaust gases HC, NO, and CO) onto g-C3N4/BiVO4 composites through introducing tourmaline powder into g-C3N4/BiVO4 composites. The introduction of tourmaline powder considerably increases the adsorption capacity of the automobile exhaust gas molecules by releasing negative ions, which enhances the contact between the automobile exhaust gas molecules and the g-C3N4/BiVO4 composite material. The enhancement in the adsorption capacity improves the hydrocarbon, CO, and NO purification efficiency by 1.73, 1.74, and 2.52 times compared to pure g-C3N4 [134]. In addition, the charge carrier’s separation and transport efficiency was enhanced by introducing an oxygen vacancy (OVs) on BiVO4 using the electrochemical reduction process. The ternary nanocomposite (BiVO4/WO3/TiO2) was prepared by coupling OVs-BiVO4 with WO3/TiO2 nanotubes for toluene gas degradation. The OVs-BiVO4/WO3/TNTs displayed a 28-times higher photocurrent intensity compared to pristine BiVO4/WO3/TNTs, which leads to higher photocatalytic toluene gas degradation. Furthermore, the stability of the composite materials was also enhanced by introducing OVs [135]. Recently, Zhu et al. developed triangular Ag nanoplates (AgNPs)-loaded BiVO4 for the degradation of gaseous formaldehyde (HCHO) under the irradiation of daylight lamp as a visible light source. The loading of triangular Ag nanoplates significantly decreases the recombination rate of photogenerated electron–hole pairs that extends the lifetime of charge carriers, which leads to a high HCHO oxidation efficiency. The plasmonic effect of AgNPs was also the reason for the enhancement of the catalytic activity [137]. Similarly, Shi et al. prepared activated carbon from semi-coke waste generated during the processing of coal and was loaded into a ternary BiVO4–BiPO4–g-C3N4 Z-scheme heterojunction photocatalyst using a one-step sol–gel method for the degradation of toluene under visible light irradiation. The activated carbon-loaded ternary composites showed a 2.43-times higher photocatalytic activity (85.6%) than the pure photocatalyst in the degradation of toluene under 60% relative humidity and 0.5 g/cm3 of composite material. The enhanced adsorption of toluene by activated carbon loading and the improved visible light response leads to higher activity in toluene VOC degradation [139]. Likewise, the development of BiVO4-based visible light active materials for VOC decomposition is still growing; however, there are some limitations that affect the possible utilization of BiVO4 on a commercial scale, which are described in the forthcoming sections.

7. Summary and Outlook

BiVO4 has been identified as one of the most promising visible-light-responsive photocatalytic materials (low bandgap energy, ~2.4 eV) for the degradation of various pollutant molecules, including VOCs. However, the materials are facing significant issues such as the high recombination rate of photogenerated charge carriers, the inappropriate position of CB of BiVO4, and the low redox ability of charge carriers. So, various engineering modifications have been performed to improve these limitations, which leads to high catalytic efficiency. Therefore, in summary, we have reviewed the synthesis methodologies of BiVO4 and various engineering modifications of BiVO4, such as changes in the morphology, metal and non-metal loading, heterojunction formation, Z-scheme- and S-scheme-type materials development and support on high-surface-area adsorbents, etc., followed by their application in VOCs degradation, which was reviewed in detail in this paper, for which it showed a greater performance.
Though the modified BiVO4 material showed an efficient visible light photocatalytic performance, it is still lagging behind in terms of commercial applications because of its poor reusability and low lifetime. Furthermore, for the better commercial utilization and recovery of materials, there is a need for an immobilized photocatalytic reactor. However, the presently available BiVO4-based immobilized reactor systems have a low mass transfer effect (both external and internal). In addition, morphologies-controlled synthesis generally results in a relatively larger size of BiVO4 material, which is easier to agglomerate and it decreases the active surface sites, thereby reducing the catalytic performance. Hence, future research needs to focus on advanced synthesis techniques, such as atomic layer deposition methods, plasma treatment, and other micro techniques, to achieve a greater precision in the control of the morphology and the production of an appropriate size of the BiVO4. This should be followed by smaller size BiVO4-based materials as these could be effectively used to make an efficient photocatalytic surface with a high mass-transfer effect. Future research also needs to be focused on the production of concrete evidence for the photogenerated separation of electron–hole pairs and transfer pathways/mechanism in heterojunction and Z-scheme- and S-scheme-type materials. Finally, the toxicity assessment of the synthesized materials as well as degraded solution requires increased attention in the near future for the commercial utilization of a developed photocatalytic system. We hope that this review article has created a solid foundation for the development of BiVO4-based composite materials with a high performance, as well as their associated technologies, for the decomposition of VOCs.

Author Contributions

All authors have contributed equally in writing review article. G.S.K.: conceptualization, methodology, data curation, writing—original draft, writing—review and editing, formal analysis, T.S.N.: conceptualization, methodology, writing—original draft, writing—review and editing, formal analysis, S.S.P.: writing—original draft, writing—review and editing, M.T.: writing—review and editing, R.K.C.: writing—original draft, P.D.S.: writing—original draft, U.S.S.: writing—original draft, writing—review and editing, Y.-C.L.: data curation, project administration and supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

G.S.K. is very grateful to postdoctoral research advisor Yong-Chien Ling for valuable guidance, moral support, and constant encouragement. G.S.K. is highly thankful of T.S.N., S.S.P., and Molly Thomas (M.T.) for providing their valuable contribution in writing and revising the review article.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Abbreviations

Advanced oxidation processes (AOPs), volatile organic compounds (VOCs), bismuth oxyhalides (BiOX), conduction band (CB), valence band (VB), graphene oxide (GO), reduced graphene oxide (rGO), ultraviolet (UV), graphitic carbon nitride (g-C3N4), O-dichlorobenzene (O-DCB), formaldehyde (HCHO), ozone (O3), hydrocarbons (HC), Ag nanoplates (AgNPs), oxygen vacancies (OVs), methylene blue (MB), rhodamine B (RhB), tetracyclines (TCs), methyl orange (MO), ciprofloxin (CIP), acid blue-113, AB-113, thiobencarb (TBC), indigo carmine (IC), cetyltrimethylammonium ammonia bromide (CTAB), sodium dodecyl benzene sulfonate (SDBS), ethylenediamine tetraacetic acid (EDTA), polyvinyl pyrrolidone (PVP).

References

  1. Jing, L.; Zhou, W.; Tian, G.; Fu, H. Surface tuning for oxide-based nanomaterials as efficient photocatalysts. Chem. Soc. Rev. 2013, 42, 9509–9549. [Google Scholar] [CrossRef] [PubMed]
  2. Tian, Q.; Yu, X.; Zhang, L.; Yu, D. Monodisperse raspberry-like multi hollow polymer/Ag nanocomposite microspheres for rapid catalytic degradation of methylene blue. J. Colloid Interface Sci. 2017, 491, 294–304. [Google Scholar] [CrossRef] [PubMed]
  3. Zhao, Y.; Tao, C.; Xiao, G.; Su, H. Controlled synthesis, and wastewater treatment of Ag2O/TiO2 modified chitosan-based photocatalytic film. RSC Adv. 2017, 7, 11211–11221. [Google Scholar] [CrossRef]
  4. Huang, L.; Yu, Y.; Fu, C.; Guo, H.; Li, X. Photocatalytic degradation of imidazolium ionic liquids using dye sensitized TiO2/SiO2 composites. RSC Adv. 2017, 7, 732120–732125. [Google Scholar] [CrossRef]
  5. Rahmawati, F.; Fadillah, I.; Mudjijono, M. Composite of nano-TiO2 with cellulose acetate membrane from nata de coco (Nano-TiO2/CA(NDC)) for methyl orange degradation. J. Mater. Environ. Sci. 2017, 8, 289–297. [Google Scholar]
  6. You, J.; Guo, Y.; Guo, R.; Liu, X. A review of visible light-active photocatalysts for water disinfection: Features and prospects. Chem. Eng. J. 2019, 373, 624–641. [Google Scholar] [CrossRef]
  7. Vijitha, R.; Nagaraja, K.; Hanafiah, M.M.; Rao, K.M.; Venkateswarlu, K.; Lakkaboyana, S.K.; Rao, K.S.V. Fabrication of eco-friendly polyelectrolyte membranes based on sulfonate grafted sodium alginate for drug delivery, toxic metal ion removal and fuel cell applications. Polymers 2021, 13, 3293. [Google Scholar] [CrossRef]
  8. Pirhashemi, M.; Habibi-Yangjeh, A.; Rahim Pouran, S. Review on the criteria anticipated for the fabrication of highly efficient ZnO-based visible-light-driven photocatalysts. J. Ind. Eng. Chem. 2018, 62, 1–25. [Google Scholar] [CrossRef]
  9. Liu, L.; Zhou, H. Investigation and assessment of volatile organic compounds in water sources in China. Environ. Monit. Assess. 2011, 173, 825–836. [Google Scholar] [CrossRef]
  10. Kostopoulou, M.N.; Golfinopoulos, S.K.; Nikolaou, A.D.; Xilourgidis, N.K.; Lekkas, T.D. Volatile organic compounds in the surface waters of Northern Greece. Chemosphere 2000, 40, 527–532. [Google Scholar] [CrossRef]
  11. Almaie, S.; Vatanpour, V.; Rasoulifard, M.H.; Koyuncu, I. Volatile organic compounds (VOCs) removal by photocatalysts: A review. Chemosphere 2022, 306, 135655. [Google Scholar] [CrossRef] [PubMed]
  12. Jin, X.; Wu, Y.; Santhamoorthy, M.; Le, T.T.N.; Le, V.T.; Yuan, Y.; Xia, C. Volatile organic compounds in water matrices: Recent progress, challenges, and perspective. Chemosphere 2022, 308, 136182. [Google Scholar] [CrossRef] [PubMed]
  13. Tan, L.; Yu, C.; Wang, M.; Zhang, S.; Sun, J.; Dong, S.; Sun, J. Synergistic effect of adsorption and photocatalysis of 3D g-C3N4-agar hybrid aerogels. Appl. Surf. Sci. 2019, 467–468, 286–292. [Google Scholar] [CrossRef]
  14. Alharbi, N.S.; Hu, B.W.; Hayat, T.; Rabah, S.O.; Alsaedi, A.; Zhuang, L.; Wang, X.K. Efficient elimination of environmental pollutants through sorption reduction and photocatalytic degradation using nanomaterials. Front. Chem. Sci. Eng. 2020, 14, 1124–1135. [Google Scholar] [CrossRef]
  15. Qu, S.; Wang, W.; Pan, X.; Li, C. Improving the Fenton catalytic performance of FeOCl using an electron mediator. J. Hazard. Mater. 2020, 384, 121494. [Google Scholar] [CrossRef]
  16. Zhang, L.; Wang, H.; Chen, Z.; Wong, P.K.; Liu, J. Bi2WO6 micro/nano-structures: Synthesis, modifications and visible-light-driven photocatalytic applications. Appl. Catal. B Environ. 2011, 106, 1–13. [Google Scholar] [CrossRef]
  17. Yang, J. Evaluation Procedure of Photocatalysts for VOCs Degradation Based on Density Functional Theory: G-C3N4 Dots/Graphene. Acta Phys.-Chim. Sin. 2021, 37, 2011039. [Google Scholar] [CrossRef]
  18. Wang, Y.; Long, Y.; Zhang, D. Novel bifunctional V2O5/BiVO4 nanocomposite materials with enhanced antibacterial activity. Taiwan Inst. Chem. Eng. 2016, 68, 387–395. [Google Scholar] [CrossRef]
  19. Lai, Y.; Meng, M.; Yu, Y. One-step synthesis, characterizations and mechanistic study of nanosheets-constructed fluffy ZnO and Ag/ZnO spheres used for Rhodamine B photodegradation. Appl. Catal. B 2010, 100, 491–501. [Google Scholar] [CrossRef]
  20. Pirhashemi, M.; Habibi-Yangjeh, A. ZnO/NiWO4/Ag2CrO4 nanocomposites with p–n–n heterojunctions: Highly improved activity for degradations of water contaminants under visible light. Sep. Purif. Technol. 2018, 193, 69–80. [Google Scholar] [CrossRef]
  21. Choi, Y.I.; Jeon, K.H.; Kim, H.S.; Lee, J.H.; Park, S.J.; Roh, J.E.; Khan, M.M.; Sohn, Y. TiO2/BiOX (X = Cl, Br, I) hybrid microspheres for artificial waste water and real sample treatment under visible light irradiation. Sep. Purif. Technol. 2016, 160, 28–42. [Google Scholar] [CrossRef]
  22. Akhundi, A.; Habibi-Yangjeh, A. Graphitic carbon nitride nanosheets decorated with CuCr2O4 nanoparticles: Novel photocatalysts with high performances in visible light degradation of water pollutants. J. Colloid Interface Sci. 2017, 504, 697–710. [Google Scholar] [CrossRef] [PubMed]
  23. Xu, C.; Ravi Anusuyadevi, P.; Aymonier, C.; Luque, R.; Marre, S. Nanostructured materials for photocatalysis. Chem. Soc. Rev. 2019, 48, 3868–3902. [Google Scholar] [CrossRef] [PubMed]
  24. Feizpoor, S.; Habibi-Yangjeh, A. Integration of Ag2WO4 and AgBr with TiO2 to fabricate ternary nanocomposites: Novel plasmonic photocatalysts with remarkable activity under visible light. Mater. Res. Bull. 2018, 99, 93–102. [Google Scholar] [CrossRef]
  25. Deng, W.; Zhao, H.; Pan, F.; Feng, X.; Jung, B.; Wahab, A.A.; Batchelor, B.; Li, Y. Visible-light-driven photocatalytic degradation of organic water pollutants promoted by sulfate addition. Environ. Sci. Technol. 2017, 51, 13372–13379. [Google Scholar] [CrossRef] [PubMed]
  26. Wenderich, K.; Mul, G. Methods, Mechanism, and applications of photo deposition in photocatalysis: A review. Chem. Rev. 2016, 116, 14587–14619. [Google Scholar] [CrossRef]
  27. Moreira, F.C.; Boaventura, R.A.R.; Brillas, E.; Vilar, V.J.P. Electrochemical advanced oxidation processes: A review on their application to synthetic and real wastewaters. Appl. Catal. B 2017, 202, 217–261. [Google Scholar] [CrossRef]
  28. Malathi, A.; Vasanthakumar, V.; Arunachalam, P.; Madhavan, J.; Ghanem, M.A. A low-cost additive-free facile synthesis of BiFeWO6/BiVO4 nanocomposite with enhanced visible-light induced photocatalytic activity. J. Colloid Interface Sci. 2017, 506, 553–563. [Google Scholar] [CrossRef]
  29. Qiao, Z.; Yan, T.; Li, W.; Huang, B. In situ anion exchange synthesis of In2S3/In (OH)3 heterostructures for efficient photocatalytic degradation of MO under solar light. New J. Chem. 2017, 41, 3134–3142. [Google Scholar] [CrossRef]
  30. Malathi, A.; Arunachalam, P.; Grace, A.N.; Madhavan, J.; Al-Mayouf, A.M. A robust visible-light driven BiFeWO6/BiOI nanohybrid with efficient photocatalytic and photoelectrochemical performance. Appl. Surf. Sci. 2017, 412, 85–95. [Google Scholar] [CrossRef]
  31. Palmai, M.; Zahran, E.M.; Angaramo, S.; Balint, S.; Paszti, Z.; Knecht, M.R.; Bachas, L.G. Pd-decorated m-BiVO4/BiOBr ternary composite with dual heterojunction for enhanced photocatalytic activity. J. Mater. Chem. A 2017, 5, 529–534. [Google Scholar] [CrossRef]
  32. Chen, L.; Meng, D.; Wu, X.; Wang, A.; Wang, J.; Yu, M.; Liang, Y. Enhanced Visible Light Photocatalytic Performances of Self-Assembled Hierarchically Structured BiVO4/Bi2WO6 Heterojunction Composites with Different Morphologies. RSC Adv. 2016, 6, 52300–52309. [Google Scholar] [CrossRef]
  33. Huang, H.; Liu, L.; Zhang, Y.; Tian, N. Novel BiIO4/BiVO4 composite photocatalyst with highly improved visible-light-induced photocatalytic performance for rhodamine B degradation and photocurrent generation. RSC Adv. 2015, 5, 1161–1167. [Google Scholar] [CrossRef]
  34. Shang, M.; Wang, W.; Ren, J.; Sun, S.; Zhang, L. A Novel BiVO4 Hierarchical Nanostructure: Controllable Synthesis, Growth Mechanism, and Application in Photocatalysis. CrystEngComm 2010, 12, 1754–1758. [Google Scholar] [CrossRef]
  35. Fan, T.; Chen, C.; Tang, Z. Hydrothermal Synthesis of Novel BiFeO3/BiVO4 Heterojunctions with Enhanced Photocatalytic Activities under Visible Light Irradiation. RSC Adv. 2016, 6, 9994–10000. [Google Scholar] [CrossRef]
  36. Su, C.; Li, R.; Li, C.; Wang, W. Piezo-promoted regeneration of Fe2+ boosts peroxydisulfate activation by Bi2Fe4O9 nanosheets. Appl. Catal. B Environ. 2022, 310, 121330. [Google Scholar] [CrossRef]
  37. Zhou, L.; Wang, W.; Zhang, L.; Xu, H.; Zhu, W. Single-Crystalline BiVO4 Microtubes with Square Cross-Sections: Microstructure, Growth Mechanism, and Photocatalytic Property. J. Phys. Chem. C 2007, 111, 13659–13664. [Google Scholar] [CrossRef]
  38. Walsh, A.; Yan, Y.; Huda, M.N.; Al-Jassim, M.M.; Wei, S.-H. Band Edge Electronic Structure of BiVO4: Elucidating the Role of the Bi s and V d Orbitals. Chem. Mater. 2009, 21, 547–551. [Google Scholar] [CrossRef]
  39. Guo, M.; Wang, Y.; He, Q.; Wang, W.; Wang, W.; Fu, Z.; Wang, H. Enhanced Photocatalytic Activity of S-Doped BiVO4 Photocatalysts. RSC Adv. 2015, 5, 58633–58639. [Google Scholar] [CrossRef]
  40. Hu, Y.; Fan, J.; Pu, C.; Li, H.; Liu, E.; Hu, X. Facile Synthesis of Double Cone-Shaped Ag4V2O7/BiVO4 Nanocomposites with Enhanced Visible Light Photocatalytic Activity for Environmental Purification. J. Photochem. Photobiol. A Chem. 2017, 337, 172–183. [Google Scholar] [CrossRef]
  41. Lv, D.; Zhang, D.; Pu, X.; Kong, D.; Lu, Z.; Shao, X.; Ma, H.; Dou, J. One-Pot Combustion Synthesis of BiVO4/BiOCl Composites with Enhanced Visible-Light Photocatalytic Properties. Sep. Purif. Technol. 2017, 174, 97–103. [Google Scholar] [CrossRef]
  42. Kamble, G.S.; Ling, Y.-C. Solvothermal Synthesis of Facet-Dependent BiVO4 Photocatalyst with Enhanced Visible-Light-Driven Photocatalytic Degradation of Organic Pollutant: Assessment of Toxicity by Zebrafish Embryo. Sci. Rep. 2020, 10, 12993. [Google Scholar] [CrossRef]
  43. Kudo, A.; Omori, K.; Kato, H. A Novel Aqueous Process for Preparation of Crystal Form-Controlled and Highly Crystalline BiVO4 Powder from Layered Vanadates at Room Temperature and Its Photocatalytic and Photophysical Properties. J. Am. Chem. Soc. 1999, 121, 11459–11467. [Google Scholar] [CrossRef]
  44. Deng, Y.; Zhao, R. Advanced Oxidation Processes (AOPs) in Wastewater Treatment. Curr. Pollut. Rep. 2015, 1, 167–176. [Google Scholar] [CrossRef]
  45. Pandis, P.K.; Kalogirou, C.; Kanellou, E.; Vaitsis, C.; Savvidou, M.G.; Sourkouni, G.; Zorpas, A.A.; Argirusis, C. Key Points of Advanced Oxidation Processes (AOPs) for Wastewater, Organic Pollutants and Pharmaceutical Waste Treatment: A Mini Review. Chem. Eng. 2022, 6, 8. [Google Scholar] [CrossRef]
  46. Liu, X.; Gu, S.; Zhao, Y.; Zhou, G.; Li, W. BiVO4, Bi2WO6 and Bi2MoO6 Photocatalysis: A Brief Review. J. Mater. Sci. Technol. 2020, 56, 45–68. [Google Scholar] [CrossRef]
  47. Zhong, X.; Li, Y.; Wu, H.; Xie, R. Recent Progress in BiVO4-Based Heterojunction Nanomaterials for Photocatalytic Applications. Mater. Sci. Eng. B 2023, 289, 116278. [Google Scholar] [CrossRef]
  48. Tokunaga, S.; Kato, H.; Kudo, A. Selective Preparation of Monoclinic and Tetragonal BiVO4 with Scheelite Structure and Their Photocatalytic Properties. Chem. Mater. 2001, 13, 4624–4628. [Google Scholar] [CrossRef]
  49. Park, Y.; McDonald, K.J.; Choi, K.-S. Progress in bismuth vanadate photoanodes for use in solar water oxidation. Chem. Soc. Rev. 2013, 42, 2321–2337. [Google Scholar] [CrossRef]
  50. Nagabhushana, G.P.; Nagaraju, G.; Chandrappa, G.T. Synthesis of Bismuth Vanadate: Its Application in H2 Evolution and Sunlight-Driven Photodegradation. J. Mater. Chem. A Mater. 2013, 1, 388–394. [Google Scholar] [CrossRef]
  51. Sun, S.; Wang, W.; Zhou, L.; Xu, H. Efficient Methylene Blue Removal Over Hydrothermally Synthesized Starlike BiVO4. Ind. Eng. Chem. Res. 2009, 48, 1735–1739. [Google Scholar] [CrossRef]
  52. Lei, B.-X.; Zhang, P.; Wang, S.-N.; Li, Y.; Huang, G.-L.; Sun, Z.-F. Additive-Free Hydrothermal Synthesis of Novel Bismuth Vanadium Oxide Dendritic Structures as Highly Efficient Visible-Light Photocatalysts. Mater. Sci. Semicond. Process. 2015, 30, 429–434. [Google Scholar] [CrossRef]
  53. Wang, B.; Guo, L.; He, T. Fabrication of an Olive-like BiVO4 Hierarchical Architecture with Enhanced Visible-Light Photocatalytic Activity. RSC Adv. 2016, 6, 30115–30124. [Google Scholar] [CrossRef]
  54. Obregón, S.; Caballero, A.; Colón, G. Hydrothermal Synthesis of BiVO4: Structural and Morphological Influence on the Photocatalytic Activity. Appl. Catal. B 2012, 117, 59–66. [Google Scholar] [CrossRef]
  55. Lu, Y.; Luo, Y.-S.; Xiao, H.-M.; Fu, S.-Y. Novel Core–Shell Structured BiVO4 Hollow Spheres with an Ultra-High Surface Area as Visible-Light-Driven Catalyst. CrystEngComm 2014, 16, 6059–6065. [Google Scholar] [CrossRef]
  56. Meng, X.; Zhang, L.; Dai, H.; Zhao, Z.; Zhang, R.; Liu, Y. Surfactant-Assisted Hydrothermal Fabrication and Visible-Light-Driven Photocatalytic Degradation of Methylene Blue over Multiple Morphological BiVO4 Single-Crystallites. Mater. Chem. Phys. 2011, 125, 59–65. [Google Scholar] [CrossRef]
  57. Zhang, L.; Chen, D.; Jiao, X. Monoclinic Structured BiVO4 Nanosheets: Hydrothermal Preparation, Formation Mechanism, and Coloristic and Photocatalytic Properties. J. Phys. Chem. B 2006, 110, 2668–2673. [Google Scholar] [CrossRef]
  58. Naing, H.H.; Wang, K.; Li, Y.; Mishra, A.K.; Zhang, G. Sepiolite Supported BiVO4 Nanocomposites for Efficient Photocatalytic Degradation of Organic Pollutants: Insight into the Interface Effect towards Separation of Photogenerated Charges. Sci. Total Environ. 2020, 722, 137825. [Google Scholar] [CrossRef]
  59. Chen, Z.; Mi, N.; Huang, L.; Wang, W.; Li, C.; Teng, Y.; Gu, C. Snow-like BiVO4 with rich oxygen defects for efficient visible light photocatalytic degradation of ciprofloxacin. Sci. Total Environ. 2022, 808, 152083. [Google Scholar] [CrossRef]
  60. Cheng, Y.; Chen, J.; Yan, X.; Zheng, Z.; Xue, Q. Preparation of Porous BiVO4 Fibers by Electrospinning and Their Photocatalytic Performance under Visible Light. RSC Adv. 2013, 3, 20606–20612. [Google Scholar] [CrossRef]
  61. Liu, G.; Liu, S.; Lu, Q.; Sun, H.; Xu, F.; Zhao, G. Synthesis of Monoclinic BiVO4 Microribbons by Sol–Gel Combined with Electrospinning Process and Photocatalytic Degradation Performances. J. Sol-Gel Sci. Technol. 2014, 70, 24–32. [Google Scholar] [CrossRef]
  62. Chen, L.; Meng, D.; Wu, X.; Wang, J.; Wang, Y.; Liang, Y. Shape-Controlled Synthesis of Novel Self-Assembled BiVO4 Hierarchical Structures with Enhanced Visible Light Photocatalytic Performances. Mater. Lett. 2016, 176, 143–146. [Google Scholar] [CrossRef]
  63. Suwanchawalit, C.; Buddee, S.; Wongnawa, S. Triton X-100 Induced Cuboid-like BiVO4 Microsphere with High Photocatalytic Performance. J. Environ. Sci. 2017, 55, 257–265. [Google Scholar] [CrossRef] [PubMed]
  64. Lai, H.-F.; Chen, C.-C.; Chang, Y.-K.; Lu, C.-S.; Wu, R.-J. Efficient Photocatalytic Degradation of Thiobencarb over BiVO4 Driven by Visible Light: Parameter and Reaction Pathway Investigations. Sep. Purif. Technol. 2014, 122, 78–86. [Google Scholar] [CrossRef]
  65. Tseng, T.K.; Lin, Y.S.; Chen, Y.J.; Chu, H. A Review of Photocatalysts Prepared by Sol-Gel Method for VOCs Removal. Int. J. Mol. Sci. 2010, 11, 2336–2361. [Google Scholar] [CrossRef]
  66. Wang, M.; Che, Y.; Niu, C.; Dang, M.; Dong, D. Lanthanum and boron co-doped BiVO4 with enhanced visible light photocatalytic activity for degradation of methyl orange. J. Rare Earths 2013, 31, 878–884. [Google Scholar] [CrossRef]
  67. Mousavi-Kamazani, M. Cube-like Cu/Cu2O/BiVO4/Bi7VO13 composite nanoparticles: Facile sol-gel synthesis for photocatalytic desulfurization of thiophene under visible light. J. Alloy. Comp. 2020, 823, 153786. [Google Scholar] [CrossRef]
  68. Drisya, K.T.; Solís-López, M.; Ríos-Ramírez, J.J.; Durán-Álvarez, J.C.; Rousseau, A.; Velumani, S.; Asomoza, R.; Kassiba, A.; Jantrania, A.; Castaneda, H. Electronic and optical competence of TiO2/BiVO4 nanocomposites in the photocatalytic processes. Sci. Rep. 2020, 10, 13507. [Google Scholar] [CrossRef]
  69. Malathi, A.; Madhavan, J.; Ashokkumar, M.; Arunachalam, P. A Review on BiVO4 Photocatalyst: Activity Enhancement Methods for Solar Photocatalytic Applications. Appl. Catal. A Gen. 2018, 555, 47–74. [Google Scholar] [CrossRef]
  70. Villa, K.; Novotný, F.; Zelenka, J.; Browne, M.P.; Ruml, T.; Pumera, M. Visible-Light-Driven Single-Component BiVO4 Micromotors with the Autonomous Ability for Capturing Microorganisms. ACS Nano 2019, 13, 8135–8145. [Google Scholar] [CrossRef]
  71. Kim, T.W.; Ping, Y.; Galli, G.A.; Choi, K.S. Simultaneous Enhancements in Photon Absorption and Charge Transport of Bismuth Vanadate Photoanodes for Solar Water Splitting. Nat. Commun. 2015, 6, 8769. [Google Scholar] [CrossRef] [PubMed]
  72. Laraib, I.; Carneiro, M.A.; Janotti, A. Effects of Doping on the Crystal Structure of BiVO4. J. Phys. Chem. C 2019, 123, 26752–26757. [Google Scholar] [CrossRef]
  73. Patil, S.S.; Mali, M.G.; Hassan, M.A.; Patil, D.R.; Kolekar, S.S.; Ryu, S.W. One-Pot in Situ Hydrothermal Growth of BiVO4/Ag/RGO Hybrid Architectures for Solar Water Splitting and Environmental Remediation. Sci. Rep. 2017, 7, 8404. [Google Scholar] [CrossRef] [PubMed]
  74. Xi, G.; Ye, J. Synthesis of Bismuth Vanadate Nanoplates with Exposed {001} Facets and Enhanced Visible-Light Photocatalytic Properties. Chem. Comm. 2010, 46, 1893–1895. [Google Scholar] [CrossRef]
  75. Rodrigues, B.S.; Branco, C.M.; Corio, P.; Souza, J.S. Controlling Bismuth Vanadate Morphology and Crystalline Structure through Optimization of Microwave-Assisted Synthesis Conditions. Cryst. Growth Des. 2020, 20, 3673–3685. [Google Scholar] [CrossRef]
  76. Li, R.; Zhang, F.; Wang, D.; Yang, J.; Li, M.; Zhu, J.; Zhou, X.; Han, H.; Li, C. Spatial Separation of Photogenerated Electrons and Holes among {010} and {110} Crystal Facets of BiVO4. Nat. Commun. 2013, 4, 1432. [Google Scholar] [CrossRef]
  77. Zhao, Y.; Li, R.; Mu, L.; Li, C. Significance of Crystal Morphology Controlling in Semiconductor-Based Photocatalysis: A Case Study on BiVO4 Photocatalyst. Cryst. Growth Des. 2017, 17, 2923–2928. [Google Scholar] [CrossRef]
  78. Yin, X.; Sun, X.; Mao, Y.; Wang, R.; Li, D.; Xie, W.; Liu, Z.; Liu, Z. Synergistically enhanced photocatalytic degradation of tetracycline hydrochloride by Z-scheme heterojunction MT-BiVO4 microsphere/P-doped g-C3N4 nanosheet composite. J. Environ. Chem. Eng. 2023, 11, 109412. [Google Scholar] [CrossRef]
  79. Sun, H.; Hua, W.; Li, Y.; Wang, J.-G. Promoting Photoelectrochemical Activity and Stability of WO3/BiVO4 Heterojunctions by Coating a Tannin Nickel Iron Complex. ACS Sustain. Chem. Eng. 2020, 8, 12637–12645. [Google Scholar] [CrossRef]
  80. Qin, C.; Tang, X.; Chen, J.; Liao, H.; Zhong, J.; Li, J. In-Situ Fabrication of Bi/BiVO4 Heterojunctions with N-Doping for Efficient Elimination of Contaminants. Colloids Surf. A Physicochem. Eng. Asp. 2021, 617, 126224. [Google Scholar] [CrossRef]
  81. Kahng, S.; Kim, J.H. Heterojunction Photoanode of SnO2 and Mo-Doped BiVO4 for Boosting Photoelectrochemical Performance and Tetracycline Hydrochloride Degradation. Chemosphere 2022, 291, 132800. [Google Scholar] [CrossRef]
  82. Sun, L.; Sun, J.; Han, N.; Liao, D.; Bai, S.; Yang, X.; Luo, R.; Li, D.; Chen, A. rGO Decorated W Doped BiVO4 Novel Material for Sensing Detection of Trimethylamine. Sens. Actuators B Chem. 2019, 298, 126749. [Google Scholar] [CrossRef]
  83. Liu, B.; Yan, X.; Yan, H.; Yao, Y.; Cai, Y.; Wei, J.; Chen, S.; Xu, X.; Li, L. Preparation and Characterization of Mo Doped in BiVO4 with Enhanced Photocatalytic Properties. Materials 2017, 10, 976. [Google Scholar] [CrossRef] [PubMed]
  84. Zhang, J.; Deng, M.; Ren, F.; Wu, Y.; Wang, Y. Effects of Mo/W Codoping on the Visible-Light Photocatalytic Activity of Monoclinic BiVO4 within the GGA + U Framework. RSC Adv. 2016, 6, 12290–12297. [Google Scholar] [CrossRef]
  85. Yao, W.; Iwai, H.; Ye, J. Effects of Molybdenum Substitution on the Photocatalytic Behavior of BiVO4. Dalton Trans. 2008, 11, 1426–1430. [Google Scholar] [CrossRef] [PubMed]
  86. Gao, Y.; Liu, F.; Chi, X.; Tian, Y.; Zhu, Z.; Guan, R.; Song, J. A Mesoporous Nanofibrous BiVO4-Ni/AgVO3 Z-Scheme Heterojunction Photocatalyst with Enhanced Photocatalytic Reduction of Cr6+ and Degradation of RhB under Visible Light. Appl. Surf. Sci. 2022, 603, 154416. [Google Scholar] [CrossRef]
  87. Regmi, C.; Kshetri, Y.K.; Kim, T.-H.; Pandey, R.P.; Ray, S.K.; Lee, S.W. Fabrication of Ni-Doped BiVO4 Semiconductors with Enhanced Visible-Light Photocatalytic Performances for Wastewater Treatment. Appl. Surf. Sci. 2017, 413, 253–265. [Google Scholar] [CrossRef]
  88. Bashir, S.; Jamil, A.; Khan, M.S.; Alazmi, A.; Abuilaiwi, F.A.; Shahid, M. Gd-Doped BiVO4 Microstructure and Its Composite with a Flat Carbonaceous Matrix to Boost Photocatalytic Performance. J. Alloy. Compd. 2022, 913, 165214. [Google Scholar] [CrossRef]
  89. Seabold, J.A.; Choi, K.-S. Efficient and Stable Photo-Oxidation of Water by a Bismuth Vanadate Photoanode Coupled with an Iron Oxyhydroxide Oxygen Evolution Catalyst. J. Am. Chem. Soc. 2012, 134, 2186–2192. [Google Scholar] [CrossRef]
  90. Labhane, P.K.; Sonawane, G.H.; Sonawane, S.H. Influence of Rare-Earth Metal on the Zinc Oxide Nanostructures: Application in the Photocatalytic Degradation of Methylene Blue and p-Nitro Phenol. Green Process. Synth. 2018, 7, 360–371. [Google Scholar] [CrossRef]
  91. Moscow, S.; Kavinkumar, V.; Sriramkumar, M.; Jothivenkatachalam, K.; Saravanan, P.; Rajamohan, N.; Vasseghian, Y.; Rajasimman, M. Impact of Erbium (Er) and Yttrium (Y) Doping on BiVO4 Crystal Structure towards the Enhancement of Photoelectrochemical Water Splitting and Photocatalytic Performance. Chemosphere 2022, 299, 134343. [Google Scholar] [CrossRef] [PubMed]
  92. Noor, M.; Sharmin, F.; Mamun, M.A.A.; Hasan, S.; Hakim, M.A.; Basith, M.A. Effect of Gd and Y Co-Doping in BiVO4 Photocatalyst for Enhanced Degradation of Methylene Blue Dye. J. Alloy. Compd. 2022, 895, 162639. [Google Scholar] [CrossRef]
  93. Sudrajat, H.; Hartuti, S. Increased photocatalytic activity caused by B doping into BiVO4. Res. Chem. Intermed. 2019, 45, 2179–2195. [Google Scholar] [CrossRef]
  94. Wang, M.; Liu, Q.; Che, Y.; Zhang, L.; Zhang, D. Characterization and Photocatalytic Properties of N-Doped BiVO4 Synthesized via a Sol–Gel Method. J Alloy. Compd. 2013, 548, 70–76. [Google Scholar] [CrossRef]
  95. Li, J.-Q.; Guo, Z.-Y.; Liu, H.; Du, J.; Zhu, Z.-F. Two-Step Hydrothermal Process for Synthesis of F-Doped BiVO4 Spheres with Enhanced Photocatalytic Activity. J Alloy. Compd. 2013, 581, 40–45. [Google Scholar] [CrossRef]
  96. Wang, M.; Zheng, H.; Liu, J.; Dong, D.; Che, Y.; Yang, C. Enhanced Visible-Light-Driven Photocatalytic Activity of B-Doped BiVO4 Synthesized Using a Corn Stem Template. Mater. Sci. Semicond. Process. 2015, 30, 307–313. [Google Scholar] [CrossRef]
  97. Cao, S.W.; Yin, Z.; Barber, J.; Boey, F.Y.C.; Loo, S.C.J.; Xue, C. Preparation of Au-BiVO4 Heterogeneous Nanostructures as Highly Efficient Visible-Light Photocatalysts. ACS Appl. Mater. Interfaces 2012, 4, 418–423. [Google Scholar] [CrossRef]
  98. Venkata Reddy, C.; Neelakanta Reddy, I.; Ravindranadh, K.; Raghava Reddy, K.; Shim, J.; Cheolho, B. Au-Doped BiVO4 Nanostructure-Based Photoanode with Enhanced Photoelectrochemical Solar Water Splitting and Electrochemical Energy Storage Ability. Appl. Surf. Sci. 2021, 545, 149030. [Google Scholar] [CrossRef]
  99. Ma, C.; Din, S.T.U.; Seo, W.C.; Lee, J.; Kim, Y.; Jung, H.; Yang, W. BiVO4 Ternary Photocatalyst Co-Modified with N-Doped Graphene Nanodots and Ag Nanoparticles for Improved Photocatalytic Oxidation: A Significant Enhancement in Photoinduced Carrier Separation and Broad-Spectrum Light Absorption. Sep. Purif. Technol. 2021, 264, 118423. [Google Scholar] [CrossRef]
  100. Yalçın, E.; Dükkancı, M. Ternary CuS@Ag/BiVO4 Composite for Enhanced Photo-Catalytic and Sono-Photocatalytic Performance under Visible Light. J. Solid State Chem. 2022, 313, 123319. [Google Scholar] [CrossRef]
  101. Yang, L.; Chen, H.; Xu, Y.; Qian, R.; Chen, Q.; Fang, Y. Synergetic Effects by Co2+ and PO43− on Mo-Doped BiVO4 for an Improved Photoanodic H2O2 Evolution. Chem. Eng. Sci. 2022, 251, 117435. [Google Scholar] [CrossRef]
  102. Resasco, J.; Zhang, H.; Kornienko, N.; Becknell, N.; Lee, H.; Guo, J.; Briseno, A.L.; Yang, P. TiO2/BiVO4 Nanowire Heterostructure Photoanodes Based on Type II Band Alignment. ACS Cent. Sci. 2016, 2, 80–88. [Google Scholar] [CrossRef]
  103. Li, Y.; Sun, X.; Tang, Y.; Ng, Y.H.; Li, L.; Jiang, F.; Wang, J.; Chen, W.; Li, L. Understanding Photoelectrocatalytic Degradation of Tetracycline over Three-Dimensional Coral-like ZnO/BiVO4 Nanocomposite. Mater. Chem. Phys. 2021, 271, 124871. [Google Scholar] [CrossRef]
  104. Hemmatpour, P.; Nezamzadeh-Ejhieh, A. A Z-Scheme CdS/BiVO4 Photocatalysis towards Eriochrome Black T: An Experimental Design and Mechanism Study. Chemosphere 2022, 307, 135925. [Google Scholar] [CrossRef] [PubMed]
  105. Wei, P.; Wen, Y.; Lin, K.; Li, X. 2D/3D WO3/BiVO4 Heterostructures for Efficient Photoelectrocatalytic Water Splitting. Int. J. Hydrogen Energy 2021, 46, 27506–27515. [Google Scholar] [CrossRef]
  106. Liaqat, M.; Khalid, N.R.; Tahir, M.B.; Znaidia, S.; Alrobei, H.; Alzaid, M. Visible Light Induced Photocatalytic Activity of MnO2/BiVO4 for the Degradation of Organic Dye and Tetracycline. Ceram. Int. 2022, 49, 10455–10461. [Google Scholar] [CrossRef]
  107. Gao, L.; Long, X.; Wei, S.; Wang, C.; Wang, T.; Li, F.; Hu, Y.; Ma, J.; Jin, J. Facile Growth of AgVO3 Nanoparticles on Mo-Doped BiVO4 Film for Enhanced Photoelectrochemical Water Oxidation. Chem. Eng. J 2019, 378, 122193. [Google Scholar] [CrossRef]
  108. Iwase, A.; Yoshino, S.; Takayama, T.; Ng, Y.H.; Amal, R.; Kudo, A. Water Splitting and CO2 Reduction under Visible Light Irradiation Using Z-Scheme Systems Consisting of Metal Sulfides, CoOx-Loaded BiVO4, and a Reduced Graphene Oxide Electron Mediator. J. Am. Chem. Soc. 2016, 138, 10260–10264. [Google Scholar] [CrossRef]
  109. Li, Y.; Qin, T.; Chen, W.; Huang, M.; Xu, J.; Lv, J. Construction of a Switchable G-C3N4/BiVO4 Heterojunction from the Z-Scheme to the Type II by Incorporation of Pyromellitic Diimide. Cryst. Growth Des. 2022, 22, 1645–1653. [Google Scholar] [CrossRef]
  110. Zhao, D.; Zong, W.; Fan, Z.; Xiong, S.; Du, M.; Wu, T.; Fang, Y.-W.; Ji, F.; Xu, X. Synthesis of Carbon-Doped BiVO4@multi-Walled Carbon Nanotubes with High Visible-Light Absorption Behavior, and Evaluation of Their Photocatalytic Properties. CrystEngComm 2016, 18, 9007–9015. [Google Scholar] [CrossRef]
  111. Wei, X.; Xu, X.; Yang, X.; Liu, Z.; Naraginti, S.; Sen, L.; Weidi, S.; Buwei, L. Novel Assembly of BiVO4@N-Biochar Nanocomposite for Efficient Detoxification of Triclosan. Chemosphere 2022, 298, 134292. [Google Scholar] [CrossRef]
  112. Cao, Y.; Yang, Y.; Zhang, R.; Li, S.; Zhang, J.; Xie, T.; Wang, D.; Lin, Y. Investigation on Al3+and Al2O3 Coexist in BiVO4 for Efficient Methylene Blue Degradation: Insight into Surface States and Charge Separation. Langmuir 2021, 37, 7617–7624. [Google Scholar] [CrossRef]
  113. Wetchakun, N.; Chaiwichain, S.; Inceesungvorn, B.; Pingmuang, K.; Phanichphant, S.; Minett, A.I.; Chen, J. BiVO4/CeO2 Nanocomposites with High Visible-Light-Induced Photocatalytic Activity. ACS Appl. Mater. Interfaces 2012, 4, 3718–3723. [Google Scholar] [CrossRef] [PubMed]
  114. Guo, M.; Wan, S.; Li, C.; Zhang, K. Graphene oxide as a hole extraction layer loaded on BiVO4 photoanode for highly efficient photoelectrochemical water splitting. Rare Met. 2022, 41, 3795–3802. [Google Scholar] [CrossRef]
  115. Alhaddad, M.; Amin, M.S. Removal of Ciprofloxacin Applying Pt@BiVO4-g-C3N4 Nanocomposite under Visible Light. Opt. Mater. 2022, 124, 111976. [Google Scholar] [CrossRef]
  116. Dabodiya, T.S.; Selvarasu, P.; Murugan, A.V. Tetragonal to Monoclinic Crystalline Phases Change of BiVO4 via Microwave-Hydrothermal Reaction: In Correlation with Visible-Light-Driven Photocatalytic Performance. Inorg. Chem. 2019, 58, 5096–5110. [Google Scholar] [CrossRef]
  117. Patil, S.S.; Lee, J.; Kim, T.; Nagappagari, L.R.; Lee, K. Controlled Synthesis and Structural Modulation to Boost Intrinsic Photocatalytic Activity of BiVO4. CrystEngComm 2022, 24, 2686–2696. [Google Scholar] [CrossRef]
  118. Yan, M.; Wu, Y.; Yan, Y.; Yan, X.; Zhu, F.; Hua, Y.; Shi, W. Synthesis and Characterization of Novel BiVO4/Ag3VO4 Heterojunction with Enhanced Visible-Light-Driven Photocatalytic Degradation of Dyes. ACS Sustain. Chem. Eng. 2016, 4, 757–766. [Google Scholar] [CrossRef]
  119. Bao, S.; Wang, Z.; Zhang, J.; Tian, B. Facet-Heterojunction-Based Z-Scheme BiVO4{010} Microplates Decorated with AgBr-Ag Nanoparticles for the Photocatalytic Inactivation of Bacteria and the Decomposition of Organic Contaminants. ACS Appl. Nano Mater. 2020, 3, 8604–8617. [Google Scholar] [CrossRef]
  120. Li, Y.; Wang, X.; Wang, X.; Xia, Y.; Zhang, A.; Shi, J.; Gao, L.; Wei, H.; Chen, W. Z-scheme BiVO4/g-C3N4 heterojunction: An efficient, stable and heterogeneous catalyst with highly enhanced photocatalytic activity towards Malachite Green assisted by H2O2 under visible light. Colloids Surf. A Physicochem. Eng. Asp. 2021, 618, 126445. [Google Scholar] [CrossRef]
  121. Hu, C.; Tian, M.; Wu, L.; Chen, L. Enhanced photocatalytic degradation of paraben preservative over designed g-C3N4/BiVO4 S-scheme system and toxicity assessment. Ecotoxicol. Environ. Saf. 2022, 231, 113175. [Google Scholar] [CrossRef]
  122. Huang, C.-M.; Pan, G.-T.; Peng, P.-Y.; Yang, T.C.-K. In Situ DRIFT Study of Photocatalytic Degradation of Gaseous Isopropanol over BiVO4 under Indoor Illumination. J. Mol. Catal. A Chem. 2010, 327, 38–44. [Google Scholar] [CrossRef]
  123. Hu, Y.; Li, D.; Zheng, Y.; Chen, W.; He, Y.; Shao, Y.; Fu, X.; Xiao, G. BiVO4/TiO2 Nanocrystalline Heterostructure: A Wide Spectrum Responsive Photocatalyst towards the Highly Efficient Decomposition of Gaseous Benzene. Appl. Catal. B 2011, 104, 30–36. [Google Scholar] [CrossRef]
  124. Zhao, W.R.; Zeng, W.Y.; Xi, H.P.; Yu, X.X. Photocatalytic Degradation of Gas-Phase Toluene over CuO Loaded BiVO4 Hollow Nanospheres under Visible-Light Irradiation. Wuli Huaxue Xuebao/Acta Phys.-Chim. Sin. 2014, 30, 761–767. [Google Scholar] [CrossRef]
  125. Sun, J.; Li, X.; Zhao, Q.; Ke, J.; Zhang, D. Novel V2O5/BiVO4/TiO2 Nanocomposites with High Visible-Light-Induced Photocatalytic Activity for the Degradation of Toluene. J. Phys. Chem. C 2014, 118, 10113–10121. [Google Scholar] [CrossRef]
  126. Sun, J.; Li, X.; Zhao, Q.; Tadé, M.O.; Liu, S. Quantum-Sized BiVO4 Modified TiO2 Microflower Composite Heterostructures: Efficient Production of Hydroxyl Radicals towards Visible Light-Driven Degradation of Gaseous Toluene. J. Mater. Chem. A Mater. 2015, 3, 21655–21663. [Google Scholar] [CrossRef]
  127. Shi, Q.; Zhao, W.; Xie, L.; Chen, J.; Zhang, M.; Li, Y. Enhanced Visible-Light Driven Photocatalytic Mineralization of Indoor Toluene via a BiVO4/Reduced Graphene Oxide/Bi2O3 All-Solid-State Z-Scheme System. J. Alloy. Compd. 2016, 662, 108–117. [Google Scholar] [CrossRef]
  128. Song, X.; Li, Y.; Wei, Z.; Ye, S.; Dionysiou, D.D. Synthesis of BiVO4/P25 Composites for the Photocatalytic Degradation of Ethylene under Visible Light. Chem. Eng. J. 2017, 314, 443–452. [Google Scholar] [CrossRef]
  129. Sun, J.; Li, X.; Zhao, Q.; Tadé, M.O.; Liu, S. Construction of p-n heterojunction β-Bi2O3/BiVO4 nanocomposite with improved photoinduced charge transfer property and enhanced activity in degradation of ortho-dichlorobenzene. Appl. Catal. B 2017, 219, 259–268. [Google Scholar] [CrossRef]
  130. Sun, R.; Shi, Q.; Zhang, M.; Xie, L.; Chen, J.; Yang, X.; Chen, M.; Zhao, W. Enhanced Photocatalytic Oxidation of Toluene with a Coral-like Direct Z-Scheme BiVO4/g-C3N4 Photocatalyst. J. Alloys. Compd. 2017, 714, 619–626. [Google Scholar] [CrossRef]
  131. Chen, R.; Zhu, C.; Lu, J.; Xiao, J.; Lei, Y.; Yu, Z. BiVO4/α-Fe2O3 Catalytic Degradation of Gaseous Benzene: Preparation, Characterization and Photocatalytic Properties. Appl. Surf. Sci. 2018, 427, 141–147. [Google Scholar] [CrossRef]
  132. Hu, Y.; Chen, W.; Fu, J.; Ba, M.; Sun, F.; Zhang, P.; Zou, J. Hydrothermal Synthesis of BiVO4/TiO2 Composites and Their Application for Degradation of Gaseous Benzene under Visible Light Irradiation. Appl. Surf. Sci. 2018, 436, 319–326. [Google Scholar] [CrossRef]
  133. Yang, J.; Shi, Q.; Zhang, R.; Xie, M.; Jiang, X.; Wang, F.; Cheng, X.; Han, W. BiVO4 Quantum Tubes Loaded on Reduced Graphene Oxide Aerogel as Efficient Photocatalyst for Gaseous Formaldehyde Degradation. Carbon 2018, 138, 118–124. [Google Scholar] [CrossRef]
  134. Li, Y.; Xing, X.; Pei, J.; Li, R.; Wen, Y.; Cui, S.; Liu, T. Automobile Exhaust Gas Purification Material Based on Physical Adsorption of Tourmaline Powder and Visible Light Catalytic Decomposition of g-C3N4/BiVO4. Ceram. Int. 2020, 46, 12637–12647. [Google Scholar] [CrossRef]
  135. Sun, M.; Wang, X.; Chen, Z.; Murugananthan, M.; Chen, Y.; Zhang, Y. Stabilized Oxygen Vacancies over Heterojunction for Highly Efficient and Exceptionally Durable VOCs Photocatalytic Degradation. Appl. Catal. B 2020, 273, 119061. [Google Scholar] [CrossRef]
  136. Shi, Q.; Zhang, M.; Zhang, Z.; Li, Y.; Qu, Y.; Liu, Z.; Yang, J.; Xie, M.; Han, W. Energy and Separation Optimization of Photogenerated Charge in BiVO4 Quantum Dots by Piezo-Potential for Efficient Gaseous Pollutant Degradation. Nano Energy 2020, 69, 104448. [Google Scholar] [CrossRef]
  137. Zhu, Z.; Lin, Y.-C.; Chung, C.-L.; Wu, R.-J.; Huang, C.-L. A Novel Composite of Triangular Silver Nanoplates on BiVO4 for Gaseous Formaldehyde Degradation. Appl. Surf. Sci. 2021, 543, 148784. [Google Scholar] [CrossRef]
  138. Yang, R.; Chen, Q.; Ma, Y.; Zhu, R.; Fan, Y.; Huang, J.; Niu, H.; Dong, Y.; Li, D.; Zhang, Y.; et al. Highly Efficient Photocatalytic Hydrogen Evolution and Simultaneous Formaldehyde Degradation over Z-Scheme ZnIn2S4-NiO/BiVO4 Hierarchical Heterojunction under Visible Light Irradiation. Chem. Eng. J. 2021, 423, 130164. [Google Scholar] [CrossRef]
  139. Shi, L.; Xue, J.; Xiao, W.; Wang, P.; Long, M.; Bi, Q. Efficient Degradation of VOCs Using Semi-Coke Activated Carbon Loaded Ternary Z-Scheme Heterojunction Photocatalyst BiVO4–BiPO4–g-C3N4 under Visible Light Irradiation. Phys. Chem. Chem. Phys. 2022, 24, 22987–22997. [Google Scholar] [CrossRef] [PubMed]
  140. Luo, Y.; Sun, G.; Tian, B.; Zhang, J. Facet-Heterojunction-Based Photothermocatalyst CdS-Au-{010}BiVO4{110}-MnOx with Excellent Synergetic Effect for Toluene Degradation. Chem. Eng. J. 2022, 442, 135835. [Google Scholar] [CrossRef]
Figure 1. Number of publications on photocatalytic degradation of volatile organic compounds. (Source: Web of Science, searched keywords: photocatalytic oxidation of volatile organic compounds, photocatalytic degradation of volatile organic compounds, nitrogen-doped TiO2, and volatile organic compounds (VOCs)).
Figure 1. Number of publications on photocatalytic degradation of volatile organic compounds. (Source: Web of Science, searched keywords: photocatalytic oxidation of volatile organic compounds, photocatalytic degradation of volatile organic compounds, nitrogen-doped TiO2, and volatile organic compounds (VOCs)).
Nanomaterials 13 01528 g001
Figure 2. Schematic representation of mechanism of photocatalytic oxidation.
Figure 2. Schematic representation of mechanism of photocatalytic oxidation.
Nanomaterials 13 01528 g002
Figure 3. (A) The crystal structures of (S) tetragonal scheelite (red: V, purple: Bi, and gray: O) and (B) zircon-type BiVO4 (red: V, purple: Bi, and gray: O) (reprinted with permission from Ref. [49]; 2013, copyright from Elsevier).
Figure 3. (A) The crystal structures of (S) tetragonal scheelite (red: V, purple: Bi, and gray: O) and (B) zircon-type BiVO4 (red: V, purple: Bi, and gray: O) (reprinted with permission from Ref. [49]; 2013, copyright from Elsevier).
Nanomaterials 13 01528 g003
Figure 4. (A) Band structures of the tetragonal zircon-type BiVO4 and (B) band structure of monoclinic scheelite-type BiVO4 (reprinted with permission from Ref. [43]; 1999, copyright from American Chemical Society).
Figure 4. (A) Band structures of the tetragonal zircon-type BiVO4 and (B) band structure of monoclinic scheelite-type BiVO4 (reprinted with permission from Ref. [43]; 1999, copyright from American Chemical Society).
Nanomaterials 13 01528 g004
Figure 5. FE-SEM images of pure m-BiVO4 photocatalyst (a) at 10 µm scale bar, (b) at 1 µm scale bar. Inset at 200 µm scale bar showing the hexagonal structure of a single NP, (c) at 500 µm scale bar showing uniformly neighboring NPs, and (d) at 200 µm scale bar showing the rough diameter and thickness of a single NP. (Reprinted with permission from [42]; copyright from Nature).
Figure 5. FE-SEM images of pure m-BiVO4 photocatalyst (a) at 10 µm scale bar, (b) at 1 µm scale bar. Inset at 200 µm scale bar showing the hexagonal structure of a single NP, (c) at 500 µm scale bar showing uniformly neighboring NPs, and (d) at 200 µm scale bar showing the rough diameter and thickness of a single NP. (Reprinted with permission from [42]; copyright from Nature).
Nanomaterials 13 01528 g005
Figure 6. FE-SEM images of the BiVO4 samples attained from (a,b) 100 °C, (c,d) 140 °C, and (e,f) 180 °C hydrothermal temperatures. (Adopted with permission from [52]; copyright from Elsevier.
Figure 6. FE-SEM images of the BiVO4 samples attained from (a,b) 100 °C, (c,d) 140 °C, and (e,f) 180 °C hydrothermal temperatures. (Adopted with permission from [52]; copyright from Elsevier.
Nanomaterials 13 01528 g006
Figure 7. Close-up sequence of the evolution of the CSS BiVO4 hollow spheres: (a) 1 h, (b) 3 h, (c) 5 h, (d) 8 h, (e) 10 h, and (f) 15 h. (Reprinted with permission from [55]; copyright permission from the Royal Society of Chemistry).
Figure 7. Close-up sequence of the evolution of the CSS BiVO4 hollow spheres: (a) 1 h, (b) 3 h, (c) 5 h, (d) 8 h, (e) 10 h, and (f) 15 h. (Reprinted with permission from [55]; copyright permission from the Royal Society of Chemistry).
Nanomaterials 13 01528 g007
Figure 8. Schematic diagram of growth mechanisms of the BiVO4 particles with numerous morphologies under dissimilar hydrothermal conditions. (Reprinted with permission from [56]; copyright from Elsevier).
Figure 8. Schematic diagram of growth mechanisms of the BiVO4 particles with numerous morphologies under dissimilar hydrothermal conditions. (Reprinted with permission from [56]; copyright from Elsevier).
Nanomaterials 13 01528 g008
Figure 9. FE-SEM images of (a,b) S-BiVO4, (c,d) A-BiVO4, and (e) N-BiVO4. (Reprinted with permission from [62]; copyright from Elsevier).
Figure 9. FE-SEM images of (a,b) S-BiVO4, (c,d) A-BiVO4, and (e) N-BiVO4. (Reprinted with permission from [62]; copyright from Elsevier).
Nanomaterials 13 01528 g009
Figure 10. SEM image of the as-prepared BiVO4 photocatalyst. (Reprinted with permission from [64]; copyright from Elsevier).
Figure 10. SEM image of the as-prepared BiVO4 photocatalyst. (Reprinted with permission from [64]; copyright from Elsevier).
Nanomaterials 13 01528 g010
Figure 13. (A) Au-doped BiVO4 heterogeneous nanostructures. (i) TEM image of Au−BiVO4 nanoplates. (ii) Photocatalytic degradation performance of MO dye with Au−BiVO4 under light irradiation. (iii) Schematic representation of mechanism of photocatalytic water oxidation using Au−BiVO4 using sacrificial agent S2O82−. Facile transfer of electron from BiVO4 to Au shifts fermi energy level (Ef) of Au to more negative potentials. Reproduced with permission from Ref. [97]. (B) N−GNDs/Ag/BiVO4 photocatalyst for degradation of TC•HCl. (i) HRTEM image of N−GNDs/Ag/BiVO4. (ii) Schematic illustration for electron transfer mechanism for degradation of TC•HCl. Reproduced with permission from Ref. [99].
Figure 13. (A) Au-doped BiVO4 heterogeneous nanostructures. (i) TEM image of Au−BiVO4 nanoplates. (ii) Photocatalytic degradation performance of MO dye with Au−BiVO4 under light irradiation. (iii) Schematic representation of mechanism of photocatalytic water oxidation using Au−BiVO4 using sacrificial agent S2O82−. Facile transfer of electron from BiVO4 to Au shifts fermi energy level (Ef) of Au to more negative potentials. Reproduced with permission from Ref. [97]. (B) N−GNDs/Ag/BiVO4 photocatalyst for degradation of TC•HCl. (i) HRTEM image of N−GNDs/Ag/BiVO4. (ii) Schematic illustration for electron transfer mechanism for degradation of TC•HCl. Reproduced with permission from Ref. [99].
Nanomaterials 13 01528 g013
Figure 14. (A) BiVO4@N-Biochar nanocomposite photocatalyst for detoxification of triclosan. (i) SEM image of BiVO4 and (ii) SEM image of BiVO4@N-Biochar. (iii) Schematics for degradation mechanism of TCS using BiVO4@N-biochar under light irradiation. (Reprinted with permission from Ref. [111]). (B) Al-Doped BiVO4 composites, (i) schematic representation of hydrothermal assisted synthesis, and (ii) comparison of photodegrading efficiency of Al-BiVO4 composites in presence of different scavengers. SEM image of (iii) pure BiVO4 (iv) 0.03-BV and (v) 0.3-BV composite photocatalysts. Reprinted with permission from Ref. [112]. (C) BiVO4/CeO2 nanocomposites as visible light photocatalysts, (i) XRD patterns of pure BiVO4, pure CeO2, and BiVO4/CeO2 nanocomposites with various mole ratios, and (ii) comparison of photoactivity of different photocatalysts for MB degradation. Reprinted with permission from Ref. [113].
Figure 14. (A) BiVO4@N-Biochar nanocomposite photocatalyst for detoxification of triclosan. (i) SEM image of BiVO4 and (ii) SEM image of BiVO4@N-Biochar. (iii) Schematics for degradation mechanism of TCS using BiVO4@N-biochar under light irradiation. (Reprinted with permission from Ref. [111]). (B) Al-Doped BiVO4 composites, (i) schematic representation of hydrothermal assisted synthesis, and (ii) comparison of photodegrading efficiency of Al-BiVO4 composites in presence of different scavengers. SEM image of (iii) pure BiVO4 (iv) 0.03-BV and (v) 0.3-BV composite photocatalysts. Reprinted with permission from Ref. [112]. (C) BiVO4/CeO2 nanocomposites as visible light photocatalysts, (i) XRD patterns of pure BiVO4, pure CeO2, and BiVO4/CeO2 nanocomposites with various mole ratios, and (ii) comparison of photoactivity of different photocatalysts for MB degradation. Reprinted with permission from Ref. [113].
Nanomaterials 13 01528 g014
Figure 15. (A) BiVO4/Ag/rGO nanocomposite architectures. (i) Schematic illustration of growth of BiVO4 and Ag Nps on rGO nanosheets, (ii) photocatalytic degradation performance of BiVO4/Ag/rGO for MB dye under light, (iii) SEM image of BiVO4/Ag/rGO, and (iv) HRTEM image showing the three components rGO, Ag, and rGO. Reprinted with permission from Ref. [73]. (B) (i) TEM image of Pt@BiVO4-g-C3N4 nanocomposite, (ii) degradation ratio of ciprofloxacin using composite photocatalysts, and (iii) photocatalytic electron transfer mechanism. Reprinted with permission from Ref. [115].
Figure 15. (A) BiVO4/Ag/rGO nanocomposite architectures. (i) Schematic illustration of growth of BiVO4 and Ag Nps on rGO nanosheets, (ii) photocatalytic degradation performance of BiVO4/Ag/rGO for MB dye under light, (iii) SEM image of BiVO4/Ag/rGO, and (iv) HRTEM image showing the three components rGO, Ag, and rGO. Reprinted with permission from Ref. [73]. (B) (i) TEM image of Pt@BiVO4-g-C3N4 nanocomposite, (ii) degradation ratio of ciprofloxacin using composite photocatalysts, and (iii) photocatalytic electron transfer mechanism. Reprinted with permission from Ref. [115].
Nanomaterials 13 01528 g015
Figure 16. (A) Heterojunctions of BiVO4 emerged from mixed of tetragonal and monoclinic crystalline phase. (i) HRTEM image of BiVO4 indicating the mixed-phase and (ii) schematic showing the effect of phase transition on the photocatalytic degradation for RhB dye. Reprinted with permission from Ref. [116]. (B) (i) Controlled synthesis of BiVO4, heterojunctions, and photoactivity thereof. (ii) Photocatalytic dye degradation mechanism showing electron transfer in single-phase BVO and tz-BVO/m-BVO heterojunction under light illumination. Reprinted with permission from Ref. [117].
Figure 16. (A) Heterojunctions of BiVO4 emerged from mixed of tetragonal and monoclinic crystalline phase. (i) HRTEM image of BiVO4 indicating the mixed-phase and (ii) schematic showing the effect of phase transition on the photocatalytic degradation for RhB dye. Reprinted with permission from Ref. [116]. (B) (i) Controlled synthesis of BiVO4, heterojunctions, and photoactivity thereof. (ii) Photocatalytic dye degradation mechanism showing electron transfer in single-phase BVO and tz-BVO/m-BVO heterojunction under light illumination. Reprinted with permission from Ref. [117].
Nanomaterials 13 01528 g016
Figure 17. (A) BiVO4/ZnO heterojunction. (i) Schematic representation of coral-like BVO/ZnO nanostructures, (ii) schematic for possible electron transfer mechanism in BiVO4/ZnO junction catalyst, and (iii) photoelectrocatalytic removal efficiency of aqueous tetracycline (50 mL, 20 mg/L, 1.2 V vs. Ag/AgCl) in the presence of varied samples. Reprinted with permission from Ref. [103]. (B) Visible-light-driven BiVO4/Ag3VO4 heterojunction, (i) TEM and HRTEM image of BiVO4/Ag3VO4 heterojunction, and (ii) comparison of photocatalytic RhB dye using different catalysts. (iii) Photocatalytic reaction mechanism and (iv) comparison of electron impedance spectroscopy (EIS) results. Reprinted with permission from Ref. [118].
Figure 17. (A) BiVO4/ZnO heterojunction. (i) Schematic representation of coral-like BVO/ZnO nanostructures, (ii) schematic for possible electron transfer mechanism in BiVO4/ZnO junction catalyst, and (iii) photoelectrocatalytic removal efficiency of aqueous tetracycline (50 mL, 20 mg/L, 1.2 V vs. Ag/AgCl) in the presence of varied samples. Reprinted with permission from Ref. [103]. (B) Visible-light-driven BiVO4/Ag3VO4 heterojunction, (i) TEM and HRTEM image of BiVO4/Ag3VO4 heterojunction, and (ii) comparison of photocatalytic RhB dye using different catalysts. (iii) Photocatalytic reaction mechanism and (iv) comparison of electron impedance spectroscopy (EIS) results. Reprinted with permission from Ref. [118].
Nanomaterials 13 01528 g017
Figure 18. Z-Scheme BiVO4{010} microplates deposited with AgBr−Ag nanoparticles as photocatalyst. (i) SEM image of AgBrAg−BiVO4 {010} and (ii) schematics for mechanism of bacterial inactivation using AgBr−Ag−BiVO4 {010} under light irradiation. (iii) Photocatalytic inactivation for Escherichia coliK-12 (reprinted with permission from Ref. [119].
Figure 18. Z-Scheme BiVO4{010} microplates deposited with AgBr−Ag nanoparticles as photocatalyst. (i) SEM image of AgBrAg−BiVO4 {010} and (ii) schematics for mechanism of bacterial inactivation using AgBr−Ag−BiVO4 {010} under light irradiation. (iii) Photocatalytic inactivation for Escherichia coliK-12 (reprinted with permission from Ref. [119].
Nanomaterials 13 01528 g018
Figure 19. The band structure of (A) Z-scheme- and (B) S-scheme-type composite materials.
Figure 19. The band structure of (A) Z-scheme- and (B) S-scheme-type composite materials.
Nanomaterials 13 01528 g019
Figure 20. Photocatalytic oxidation of benzene (a) and the amount of CO2 production (b) using BiVO4/TiO2 and bare nanomaterials under visible light irradiation (reprinted with permission from [123]; 2011, copyright from Elsevier.
Figure 20. Photocatalytic oxidation of benzene (a) and the amount of CO2 production (b) using BiVO4/TiO2 and bare nanomaterials under visible light irradiation (reprinted with permission from [123]; 2011, copyright from Elsevier.
Nanomaterials 13 01528 g020
Table 1. Basic semiconductor materials used for photocatalytic applications.
Table 1. Basic semiconductor materials used for photocatalytic applications.
S. NoSemiconductorBandgap Energy
(Eg), eV
VB PositionCB Position
1.TiO23.2+3.1−0.1
2.ZnO3.2+3.0−0.2
3.WO32.8+3.0+0.4
4.ZrO25.2+5.0+0.2
5.SnO23.8+4.1+0.3
6.SrTiO33.2+3.1−0.1
7.CuO2.1+2.36+0.26
8.ZnS3.7+1.4−2.3
9.CdS2.5+2.1−0.4
10.BiVO42.3–2.4+2.7+0.3
11.Bi2MoO62.63+2.9+0.27
12.Bi2WO62.77+3.05+0.28
13.BiOF3.64+4.24+0.6
14.BiOCl3.22+3.40.18
15.BiOBr2.64+3.0+0.36
16.BiOI1.77+2.32+0.55
17.Bi2O32.1–2.8+2.48+0.38
Table 2. Synthesis methodologies, morphology, the photocatalytic activity of pollutant(s) and degradation efficiency of BiVO4-based materials.
Table 2. Synthesis methodologies, morphology, the photocatalytic activity of pollutant(s) and degradation efficiency of BiVO4-based materials.
Method of SynthesisMorphologyPhotocatalytic Activity of Pollutant(s) and Degradation EfficiencyRefs.
Surfactant- and template-free hydrothermal methodTruncated square, 18-sidedPollutant: MB Dye (20 ppm)
Light source: 1000 W xenon lamp
% degradation: 91% after 60 min
[42]
Hydrothermal method using EDTA as a chelating agent2D star-like crystalsPollutant: MB Dye (15 ppm)
Light source: 500-W xenon lamp
% degradation: 99.3% after 25 min
[51]
Additive-free hydrothermal methodDendritic structure of BiVO4Pollutant: RhB Dye (10 ppm)
Light source: 500 W xenon lamp
% degradation: 91% after 210 min
[52]
Template-free hydrothermal methodOlive-like BiVO4Pollutant: MB Dye (10 μM)
Light source: 300 W xenon lamp
% degradation: 84.1–95.7% (Different pH value) after 180 min
[53]
Surfactant-free hydrothermal methodOctahedralPollutant: MB Dye (10 ppm)
Light source: low power xenon lamp.
% degradation: 50–60% after 120 min
[54]
Surfactant- and template-free hydrothermal methodPlate morphology and biscuit morphologyPollutant: RhB Dye (10−5 mol/L)
Light source: 500 W xenon lamp
% degradation: 99% after 270 min
[55]
Hydrothermal method in the presence of triblock copolymer P123 as a surfactantPolyhedral, rod-like, tubular, leaf-like, and spherical Pollutant: MB Dye 1.0 × 10−5 mol/L)
Light source: 300 W xenon lamp
% degradation: 90% after 120 min
[56]
Hydrothermal route using of SDBS as an anionic surfactant2D single-crystal nanosheetsPollutant: RhB Dye (2.09 10−4 mol dm−3)
Light source: Sunlight
% degradation: - 95% after 100 min
[57]
Hydrothermal processFibrous or needle-like sepiolite distributed peanut-shape monoclinic BiVO4 surfacePollutant: TCs (5 ppm); MB Dye (10 ppm)
Light source: LED lamp
% degradation: TCs 78%; MB Dye 96% after 240 min
[58]
CTAB-assisted hydrothermal methodSnow-like Pollutant: CIP (10 ppm)
Light source: 500 W xenon lamp
% degradation: 98.5% after 70 min
[59]
Electro-spinning method 1D nanofibersPollutant: RhB Dye (10 ppm)
Light source: 300 W xenon-illuminator
% degradation: 100% after 120 min
[60]
Electro-spinning method1D micro-ribbonsPollutant: RhB Dye (20 ppm)
Light source: 500 W xenon lamp
% degradation: 93.3% after 300 min
[61]
Solvothermal method through adjusting the solution pHRed blood cell, flower-like microsphere, and dendrite morphologies Pollutant: MB Dye (10 ppm)
Light source: 500 W xenon lamp
% degradation: dendrite-like < flower-like microsphere < red-blood-cell-like morphology after 180 min
[62]
Coprecipitation (500 °C for 5 h)CuboidsPollutant: IC Dye (50 mL)
Light source: fluorescence light 18 W
% degradation: ~90% after 300 min
[63]
Precipitation (450 °C for 15 min)PolyhedralPollutant: TBC (5 ppm)
Light source: two visible lamps (15 W)
% degradation: 97% after 300 min
[64]
Sol–gel methodSpherical structuresPollutant: MO (15 ppm)
Light source: 250 W halogen lamp
% degradation: 98% after 50 min
[65]
Pechini sol–gel methodRectangular cube-like, plate-like microstructures, plate-like nanostructures, nanorods, quasi-spherical structuresPollutant: Thiophen (800 ppm)
Light source: 400 W Osram lamp
% degradation: 92% after 150 min
[66]
Modified one-step sol–gel method SphericalPollutant: AB-113 (40 ppm)
Light source: 1.6 kW xenon arc ozone-free lamp
% degradation: (~99%) after 120 min
[67]
Table 3. BiVO4-based semiconductor nanomaterials for VOCs degradation.
Table 3. BiVO4-based semiconductor nanomaterials for VOCs degradation.
S. No.MaterialsVOCs DegradedVOCs ConcentrationDegradation of VOCs (%)Degradation Time of VOCs (h)Ref.
1BiVO4Isopropanol160 ppm8812[122]
2BiVO4/TiO2Benzene260 ppm848[123]
3BiVO4/CuOToluene75 μL/L856[124]
4V2O5/BiVO4/TiO2Toluene120 ppm916[125]
5Quantum-sized BiVO4/TiO2 microflowerToluene-896[126]
6BiVO4/RGO/Bi2O3Toluene25 ppm95.66[127]
7BiVO4/P25Ethylene0.15 ppm11.26[128]
8β-Bi2O3/BiVO4o-DCB-706[129]
9Coral-like Z-scheme BiVO4/g-C3N4Toluene25 ppm68.28[130]
10BiVO4/α-Fe2O3Benzene100 ppm66.873.5[131]
11BiVO4/TiO2Benzene260 ppm418[132]
12BiVO4 quantum tubes/rGOHCHO50 ppm6015 min[133]
13g-C3N4/BiVO4/tourmaline powderAutomobile exhaust gas (HC, NO, CO)300–400 ppm (HC), 2.5–4% (NO), and 45–65 ppm (CO)6.9 (HC)
7.2 (NO)
46.7 (CO)
1[134]
14Oxygen vacancies (OVs) introduced BiVO4/WO3/TiO2 nanotubesToluene100 ppm1001[135]
15BiVO4 quantum dots/ZnO nanorodHCHO50 ppm1001[136]
16Ag/BiVO4HCHO10 ppm845[137]
17ZnIn2S4-NiO/BiVO4HCHO1.5 mol/L17 mmol/h3[138]
18semi-coke activated carbon/
BiVO4–BiPO4–g-C3N4
Toluene200 ppm85.62[139]
19CdS-Au-{010}BiVO4{110}-MnOxToluene4723 mg/m397.2100 min[140]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kamble, G.S.; Natarajan, T.S.; Patil, S.S.; Thomas, M.; Chougale, R.K.; Sanadi, P.D.; Siddharth, U.S.; Ling, Y.-C. BiVO4 As a Sustainable and Emerging Photocatalyst: Synthesis Methodologies, Engineering Properties, and Its Volatile Organic Compounds Degradation Efficiency. Nanomaterials 2023, 13, 1528. https://doi.org/10.3390/nano13091528

AMA Style

Kamble GS, Natarajan TS, Patil SS, Thomas M, Chougale RK, Sanadi PD, Siddharth US, Ling Y-C. BiVO4 As a Sustainable and Emerging Photocatalyst: Synthesis Methodologies, Engineering Properties, and Its Volatile Organic Compounds Degradation Efficiency. Nanomaterials. 2023; 13(9):1528. https://doi.org/10.3390/nano13091528

Chicago/Turabian Style

Kamble, Ganesh S., Thillai Sivakumar Natarajan, Santosh S. Patil, Molly Thomas, Rajvardhan K. Chougale, Prashant D. Sanadi, Umesh S. Siddharth, and Yong-Chein Ling. 2023. "BiVO4 As a Sustainable and Emerging Photocatalyst: Synthesis Methodologies, Engineering Properties, and Its Volatile Organic Compounds Degradation Efficiency" Nanomaterials 13, no. 9: 1528. https://doi.org/10.3390/nano13091528

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

Article Metrics

Back to TopTop