Next Article in Journal
Raspberry Colloid Templated Catalysts Fabricated Using Spray Drying Method
Next Article in Special Issue
Effect of Metal Atom in Zeolitic Imidazolate Frameworks (ZIF-8 & 67) for Removal of Dyes and Antibiotics from Wastewater: A Review
Previous Article in Journal
Effect of Iron-Doping on the Structure and Photocatalytic Activity of TiO2 Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bismuth Vanadate (BiVO4) Nanostructures: Eco-Friendly Synthesis and Their Photocatalytic Applications

by
Hajar Q. Alijani
1,
Siavash Iravani
2,* and
Rajender S. Varma
3,*
1
Department of Biotechnology, Shahid Bahonar University of Kerman, Kerman 7616913439, Iran
2
Faculty of Pharmacy and Pharmaceutical Sciences, Isfahan University of Medical Sciences, Isfahan 8174673461, Iran
3
Regional Centre of Advanced Technologies and Materials, Czech Advanced Technology and Research Institute, Palacký University in Olomouc, Šlechtitelů 27, 783 71 Olomouc, Czech Republic
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(1), 59; https://doi.org/10.3390/catal13010059
Submission received: 20 November 2022 / Revised: 13 December 2022 / Accepted: 25 December 2022 / Published: 28 December 2022
(This article belongs to the Special Issue Advances in Photocatalysis for the Degradation of Organic Pollutants)

Abstract

:
Green nanotechnology plays an important role in designing environmentally-benign and sustainable synthesis techniques to provide safer products for human health and environments. In this context, the synthesis of bismuth vanadate (BiVO4) nanoparticles (NPs) based on green chemistry principles with the advantages of eco-friendliness, cost-effectiveness, and simplicity has been explored by researchers. Despite the advantages of these synthesis techniques, crucial aspects regarding their repeatability and large-scale production still need to be comprehensively explored. BiVO4 NPs have shown excellent potential in the pharmaceutical industry, cancer therapy, and photocatalysis. BiVO4 particles with monoclinic scheelite structures have been widely investigated for their environmental applications owing to their fascinating optical and electrical properties as well as their high stability and unique crystal structure properties. These NPs with good photostability and resistance to photocorrosion can be considered as promising nanophotocatalysts for degradation of pollutants including organic dyes and pharmaceutical wastes. However, additional explorations should be moved toward the optimization of reaction/synthesis conditions and associated photocatalytic mechanisms. Herein, recent developments regarding the environmentally-benign fabrication of BiVO4 NPs and their photocatalytic degradation of pollutants are deliberated, with a focus on challenges and future directions.

1. Introduction

Bismuth is a brittle white metal with a pink undertone and a rainbow matte that tarnishes yellow to dark gray. This diamagnetic metal has a hall effect. Accordingly, it has low electrical conductivity with stable electronic configuration, containing three electrons in the 6p orbital, along with high electrical resistance in a magnetic field. The most important ores of this element are bismuthinite (Bi2S3) and bismite (Bi2O3) [1,2]. Bismuth has many applications in industry and biomedicine [3,4]. For instance, bismuth subsalicylate (C7H5BiO4), under the brand name Pepto-bismol, is a colloidal drug used to treat the gastrointestinal diseases. Bismuth oxychloride (BiOCl) is a lustrous white powder that makes cosmetics shine. Bismuth compounds are also used in the production of synthetic fibers, rubber, nuclear reactors, transuranium elements, metal alloys, supercapacitor [5], among others [6]. Bismuth vanadate (BiVO4) is a yellow mineral solid that has garnered a lot of attention in recent years as a nanocatalyst. BiVO4 has three polymorphism structures—BiVO4 with reddish/yellowish brown color has a natural polymorph called pucherite with orthorhombic crystal system, and the other two polymorphs of BiVO4 are clinobisvanite and dreyerite. Clinobisvanite is an orange polymorph of BiVO4 with a monoclinic scheelite crystal system. The clinobisvanite bismuth vanadate often forms pseudo-tetragonal crystals, including tetragonal scheelite. The rarest polymorph of BiVO4 is dreyerite, the orange/brownish-yellow material with tetragonal zircon crystal system. However, monoclinic clinobisvanite is an excellent light-driven photocatalyst compared to other polymorphs [7,8,9].
The synthesis of BiVO4 nanoparticles (NPs) has been performed using different methods, namely sol-gel [10], hydrothermal technique [7], reverse-micro emulsion technique, co-precipitation [11], sonochemical [12,13], and solvothermal [14], among others [15]. The unique optical, electrical, catalytic and biocompatibility properties of BiVO4 NPs depend on their crystal structure. Meanwhile, green nanotechnology can be applied for developing eco-friendly preparative techniques for synthesizing nanomaterials with the advantages of cost-effectiveness, low toxicity, biocompatibility, and eco-friendly properties [16,17,18,19,20]. On the other hand, a wide variety of nanophotocatalysts such as ZnO, TiO2, Fe2O3, CdS, and ZnS have been deployed for environmental applications (especially photocatalytic degradation of pollutants), showing unique chemical, electrical, optical and physical features [21,22,23,24]. Among these, BiVO4 NPs with excellent photostability and resistance to photocorrosion can be considered as attractive photocatalysts for degradation of pollutants including organic dyes and pharmaceutical wastes [24,25]. Additionally, these NPs with the benefits of non-toxicity and eco-friendliness have displayed suitable antibacterial effects, making them potential candidates for environmental purposes [26]. In this context, designing composites containing BiVO4 NPs and other materials such as nickel ferrite (NiFe2O4) can help improve their photocatalytic features, thus enhancing the absorption region for efficiently degradation of pollutants [27]. For instance, BiVO4/NiFe2O4 composites were fabricated for the photodegradation of methylene blue under visible light. Accordingly, these composites could efficiently degrade methylene blue (~99%) after 90 min under visible light [28]. Lee et al. [29] reported the synthesis of copper (Cu)-doped BiVO4/graphitic carbon nitride (g-C3N4) nanocomposites with improved stability, light-harvesting efficiency, and electron/hole (e−/h+) pair separation compared to pristine g-C3N4 and BiVO4, showing enhanced photocatalytic performance [29]. Herein, the most recent advancements pertaining to the eco-friendly synthesis of BiVO4 NPs and their photocatalytic applications are cogitated, focusing on important challenges and future perspectives.

2. Eco-Friendly Synthesis of BiVO4 NPs

2.1. Biosynthesis Techniques

Biosynthesis of NPs is based on the employment of biological resources for the synthesis without the use of organic or inorganic chemicals as reducing and stabilizing agents (Table 1) [24,30]. However, crucial challenges such as the polydispersity of NPs, size/morphology controllability, and commercial/large-scale production of NPs still linger [31,32]. According to the literature, plant phytochemicals and some primary metabolites such as polyphenols, flavonoids, glycosides, tannins, proteins, terpenoids, and polysaccharides have a reducing, coating and stabilizing role in the synthesis of NPs [33,34,35,36]. In addition, each of these plant constituents has therapeutic properties; the multiple roles of plant compounds in green synthesis have garnered much attention lately where flavonoids and phenolic compounds have played a key role in the synthesis of BiVO4 NPs. The diol and hydroxyl functional groups in phenolic compounds and flavonoids reduce the metal ions [37]. It has been well established that flavonoid, terpenoids, antioxidants, and phenolic compounds could play an important role in the synthesis of BiVO4 NPs; largely free functional group in these phytochemicals being hydroxyl (OH) performing a crucial function in coating and reducing the metal ions in the biosynthesis of BiVO4 NPs [37,38].
The size of BiVO4 nanorods increased with the rise in the calcination temperature of NPs [39]; calcination of BiVO4 NPs eliminated the elemental and plant impurities. The crystallinity of BiVO4 NPs was decreased by tripling the amount of plant precursor. However, the antibacterial and antifungal activity of the NPs was increased while tripling the plant precursor [38]. The increase in plant precursor has led to the formation of bulk BiVO4 and distortion of the symmetry of the tetrahedron structure to the hexagonal of BiVO4 NPs. This increase could also enhance the absorption of electric charge, cavities and photocatalytic activity of BiVO4 NPs. These NPs exhibited efficient photocatalytic performance owing to their remarkable separation rate of photodegraded charge carriers, causing the degradation up to 98.3% under visible light irradiation for 120 min [41]. On the other hand, BiVO4 NPs containing these plant precursors had the highest inhibitory properties against MCF-7 cancer cells. Accordingly, the calcination and the amounts of plant extracts affect the crystal structure, purity, and photocatalytic and medicinal activity of BiVO4 NPs [41].

2.2. Microwave- and Ultrasonic-Assisted Synthesis

Microwave (MW)-assisted synthesis techniques have shown several advantages of cost-effectiveness, simple/time-saving purification, eco-friendliness, and fast reaction times [46,47]. These methods with significant reactivity, rapid heating, and non/low pollutions can be contemplated for safer synthesis of BiVO4 nanomaterials, reducing the energy and time consumption [48]. Spherical hollow BiVO4 nanocrystals with good optoelectronic properties were fabricated using MW-assisted combustion synthesis technique with the advantages of low energy consumption, rapidness, and simplicity. These NPs could be employed for the photocatalytic degradation of Alizarin Red S pollutant (~99.6%) after 180 min at natural pH. After the pore structure analysis, it was revealed that the lowest pore diameter of BiVO4 nanocrystals was ~4.41 nm. The synthesis routes and conditions can significantly affect the size and size distribution of these nanomaterials [49]. Tungsten (W)-doped BiVO4/WO3 heterojunctions were constructed using one-pot MW-assisted technique within 24 min. These heterojunctions exhibited improved photocatalytic performance and increased photogenerated charges lifetime, wherein W doping could reduce the recombination rates [50]. In addition, monoclinic BiVO4 structures were synthesized via a facile and rapid combined MW- and ultrasonic irradiation protocol to provide photocatalysts for the degradation of Rhodamine B under visible light irradiation. These NPs with a small crystal size and large band gap displayed excellent photocatalytic activity [51].
Souza et al. [52] synthesized BiVO4 nanoflowers decorated with gold (Au) NPs using MW irradiation. In these nanostructures, BiVO4 exhibited low band gap energy under visible light irradiation and Au NPs could serve as electron sinks and/or as electron sources via plasmon resonance, enhancing the charge separation of photogenerated electrons and holes. These photocatalysts with synergistic effects could be applied for the degradation of methylene blue (~95%) after 6 h under UV–visible light irradiation [52]. In addition, pure monoclinic BiVO4 NPs (~50 nm) were synthesized using a single-step, pH-controlled, MW-assisted technique at a temperature of 90 °C within a short reaction time (60 min) [53]. The introduced synthesis strategy can be considered as an up-scalable, low-temperature, and environmentally-benign alternative after the optimization of crucial factors such as pH, temperature, and reaction time controlling the morphology and crystal phase. These NPs could be applied for the photodegradation of Rhodamine B [53]. Despite the advantages of MW-assisted synthesis such as reduction in time and energy consumptions, the specific interaction of the microwaves with the reactive species is one of the important challenging issues. Notably, controlling the morphology and size of nanomaterials is another crucial aspect that needs suitable optimization of reaction conditions [54]. In ultrasonic-assisted synthesis of BiVO4 photocatalysts, it was revealed that ultrasonic irradiation time could affect the relevant features of visible-light-driven BiVO4 [55]. Notably, after the optimization of synthesis conditions, the enhanced photocatalytic activity could be obtained as exemplified in one study for the ultrasonic-assisted synthesis of BiVO4 photocatalysts with 60% pollutant degradation within 40 min of ultrasonic irradiation [55].

3. Photocatalytic Applications

Photocatalysts are semiconductor catalysts that absorb photons of light to create an electron–hole pair. Some important applications of semiconductor photocatalytic materials include:
Degradation of organic pollutants in industrial effluents,
Water treatment (the removal of stable organic compounds and microorganisms from municipal and laboratory wastewater),
Air purifier,
Paper industry,
Disinfection of surgical instruments, and
The removal of fingerprints from electrically and optically sensitive components.
BiVO4 NPs can be deployed for the photocatalytic hydrogen evolution and dye degradation [56]. Overall, ternary metal oxide nanomaterials have shown excellent photocatalytic abilities, since their electronic bands are generated by atomic orbitals of more than one element and the inflection of the stoichiometric ratio of the elements can superbly adjust the band gap energy and capabilities of valence and conduction bands [49]. BiVO4 as a direct band gap ternary metal oxide semiconductor with a band gap of 2.4 eV for solar light absorption can be deployed for photocatalytic applications. Because the conduction band exists closely to the empty Bi 6p orbitals, its overlap with the anti-bonding V 3d, O 2p states can additionally diminish the requirement for external bias for photocatalytic activity. In this context, the charge transport and interfacial charge transfer are crucial challenges in its performance [49,57]. BiVO4 photocatalysts can be employed for water oxidation owing to small band gap and appropriate band positions; however, for practical applications, short diffusion length should be resolved [58]. For instance, BiVO4/BiOx composites with long-term stability were prepared on conducting glasses through a hydrothermal fabrication technique and NaOH etching process, with improved photoelectrochemical catalytic capabilities towards water oxidation [58]. In addition, BiVO4 nanocatalysts were synthesized for the elimination of toxic organic pollutants from wastewater, electrochemical storage, and photoelectrochemical solar water oxidation. These NPs exhibited excellent photocatalytic performance for the degradation of methyl orange under visible light (~87.8%) within 80 min [59].
Heterogeneous photocatalysis has been performed for the degradation of organic dyes (Rhodamine 6G) using BiVO4 thin films under visible light irradiation. The sputtered BiVO4 films displayed an electronic band gap of 2.5 eV, making them suitable candidates for harvesting the visible light radiation [15]. In addition, monoclinic BiVO4 NPs (~50–70 nm) were synthesized for efficient visible light photocatalytic degradation of Rhodamine B and crystal violet as they exhibited improved photocatalytic performance for the degradation of pollutants under visible light [60]. BiVO4 NPs calcined at 400 °C exhibited excellent photocatalytic activity against methylene blue dye under solar irradiation, displaying good stability (~3 cycles) [61]. These photocatalysts could also be applied for the inhibition of pathogenic bacteria such as Pseudomonas aeruginosa, Acinetobacter baumannii, and Staphylococcus aureus. The suggested antibacterial effects of these photocatalysts included the bacterial cell wall damages, the disruption of DNA replications and bacterial metabolisms, the inhibition of formation of proteins, and the reactive oxygen species (ROS) [61]. Sharma et al. [62] fabricated monoclinic BiVO4 nanomaterials with suitable antimicrobial and photocatalytic performances. These nanostructures exhibited photocatalytic degradation efficiency towards methylene blue pollutant, along with the efficient antibacterial effects against Escherichia coli (Figure 1) [62].
BiVO4 NPs can be considered as light-driven photocatalysts due to their ferroelasticity, optical, and conductive nature [63]. These NPs have been deployed to remove phenolic compounds, methyl orange, methylene blue, and indigo carmine (Table 2). The photocatalytic performance of these NPs depends on the band structure of the electron, crystal structure, and their band gap energy. For an efficient photocatalyst, the gap band energy must be less than 3 eV to exploit light absorption in the visible region to utilize solar energy efficiently. Indirect and direct mechanisms as well as the photosensitization processes are effective in the photodegradation of pollutants [63]. According to the literature, the band gap energy caused the molecular excitation of the BiVO4 photocatalysts. Molecular excitation generates electrons in the higher conduction band (Ecb) and positive holes in the lower capacitance band (Evb) energies in BiVO4. The electron hole (h+) causes oxidative processes and traps e− for reduction processes; also, the formation of superoxide anion and hydrogen peroxide can be obtained from oxygen. During the photocatalytic oxidation processes, pollutants are completely degraded by ultraviolet radiation in the presence of BiVO4 catalysts and converted to CO2 and H2O. BiVO4 photocatalysts with environmental and energy applications have shown efficient light activation, high stability, low cost, and safety advantages for the environment and humans. In one study, BiVO4/graphene oxide/polyaniline composites were synthesized for photodegradation of methylene blue and safranin O upon the covering of polyaniline. The improvement in their photocatalytic performances was due to the formation of well-defined composite interfaces enhancing the charge separation efficacy [64].
Tahir et al. [65] constructed ternary WO3/g-C3N4@BiVO4 composites through an eco-friendly hydrothermal technique for the efficient production of hydrogen energy. These composites with active sites for photocatalytic reduction of water exhibited improved photocatalytic performance (432 μmol h−1g−1), offering great opportunities for energy harvesting. BiVO4 NPs with unique optical properties and photocatalytic performance could inhibit the recombination of photogenerated electron and holes and enhance the reduction reactions for H2 formation. The enhancement in photocatalytic efficiency of these photocatalysts could be due to the large surface area, efficient separation of electrons/holes pairs, and wide absorption region of visible light, because of the synergistic influences between WO3/g-C3N4 and BiVO4 NPs [65]. BiVO4 photocatalysts with low band gap energy have been applied for eco-friendly and sustainable H2O2 generation [66]. BiVO4 nanostructures were encapsulated with encapsulated for photocatalytic H2O2 formation; reduced graphene oxide was deployed for the stimulation of transporting charges and prevention of recombining photogenerated electron–hole pairs (Figure 2) [66]. The photocatalysts displayed efficient formation of H2O2 using oxalic acid, stimulating two-electron O2 reduction reaction with suitable cyclic stability; the photocatalytic flow reactor evaluation was applied for assessment of the feasibility of continuous generation of H2O2 [66].
BiVO4 nanophotocatalysts synthesized by an ultrasonic-assisted synthesis technique were deployed for photocatalytic degradation of organic dyes under visible light irradiation [67]. Besides, BiVO4 photocatalysts were applied for the photodegradation of Rhodamine B and 2,4-Dichlorophenol [68]. It was revealed the photocatalysts exhibited high photocatalytic performance at pH = 7 for 24 h for the photodegradation of Rhodamine B, while the best photodegradation of 2,4-Dichlorophenol could be achieved at pH = 0.5 for 24 h. Accordingly, the photocatalytic mechanism could be proposed by various charge carrier transfer pathways and active oxidation species in the heterostructured BiVO4 photocatalysts [68]. Similarly, monoclinic BiVO4 structures were investigated for the photocatalytic degradation of methyl orange under visible light irradiation [69]. The pH value could significantly affect the pore structure and morphology of these nanomaterials. As a result, spherical BiVO4 with porous structures along with the flower-cluster-like and flower-bundle-like BiVO4 structures were prepared at different pH levels. Notably, some important criteria such as surface area, bandgap energy, surface oxygen vacancy density, and porous architectures could highly affect the photocatalytic performance of these catalysts [69]. Several studies have focused on the related mechanisms of photocatalytic reactions by these photocatalysts which can help to enhance the catalytic efficiency for future practical applications [70]. Remarkably, noble metals can be deposited on the surface of BiVO4 for improving the photocatalytic efficiency by functioning as an electron trap because of the generation of the Schottky barrier, thereby decreasing the electron–hole recombination procedure. In one study, after the synthesis of Pt-BiVO4 catalysts, analysis was performed on trapping reactive oxygen species (OH and O2). As a result, the radicals like OH were generated on the surface of semiconductor as a robust oxidizing agents, which could attack the adsorbed organic molecules and participate in additional oxidation procedure [70]. Additionally, the study on dynamic of photogenerated holes in BiVO4 photoanodes for solar water oxidation revealed that two different recombination procedures limited the photocurrent formation in BiVO4 photoanodes, which included the recombination of surface-accumulated holes with bulk BiVO4 electrons along with the rapid electron/hole recombination [71].
In addition to photodegradation of dye pollutants, BiVO4 structures can be considered as promising alternatives for efficient degradation of pharmaceutical wastes [72]. For instance, spindle-shaped BiVO4/reduced graphene oxide/g-C3N4 nanocomposites with excellent solar-driven degradation activity were constructed as Z-scheme photocatalysts for the degradation of antibiotics (Figure 3) [73]; the photodegradation rates were ~81.10% and ~94.8% for tetracycline and ciprofloxacin in 60 min, respectively. These photocatalysts exhibited oriented carrier transport, photooxidation response, and superb optical activity, showing enhanced photogenerated electron–hole pairs and rapid carriers transfer under visible-light irradiation [73]. To design photocatalysts with improved photodegradation efficiency toward ciprofloxacin, the hybrid reduced graphene oxide-BiVO4 composite was designed using a facile MW-assisted synthesis technique [72]. Compared to the pure BiVO4 photocatalysts, this photocatalyst exhibited enhanced photodegradation capability toward ciprofloxacin under visible light. The composite displayed the highest ciprofloxacin degradation ratio (~68.2%) in 60 min, which was over three times than that (~22.7%) observed for pure BiVO4 photocatalysts. This enhancement in photocatalytic potential was due to the effective separation of electron–hole pairs rather than the increase in light absorption [72]. In another study, BiVO4 nanophotocatalysts with high recycling potential were synthesized for the removal of tetracycline and oxytetracycline antibiotics [74]. Accordingly, excellent performance of 72% and 83% degradation could be attained after 240 min under sunlight conditions for tetracycline and oxytetracycline, respectively. Mechanism studies revealed that the photogenerated electrons and holes could play crucial roles in the elimination of these pollutants [74].

4. Conclusions and Perspectives

BiVO4 as a narrow-band-gap semiconductor has shown excellent optical features, non-toxicity, and significant chemical stability. BiVO4 nanomaterials with attractive photocatalytic performances have been widely explored for the degradation of dye pollutants and pharmaceutical wastes as well as for photocatalytic antibacterial applications. Their photocatalytic activity is related to their band gap, particle size, and crystalline phase. Notably, the pH value, morphology, and crystalline phase with significant effects on photocatalytic activities of BiVO4 nanomaterials ought to be further explored. Future explorations should focus on the associated photocatalytic mechanisms and optimization of reaction/synthesis conditions. A wide variety of biosynthesis techniques have been introduced for the synthesis of nanocatalysts with the added benefits of safety, inexpensiveness, simplicity, and environmentally-benign properties. However, additional efforts are still required pertaining to their large-scale/commercial production, optimized reaction/synthesis conditions, stability of the ensued NPs, size distribution, and the adequate control of size/morphology. In this context, understanding the related metabolic pathways, reducing/capping agents, and understanding the underlying mechanisms can help to better control the properties of NPs.

Author Contributions

H.Q.A., writing—review; S.I. and R.S.V., conceptualization, writing—review, and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

There is no conflict of interest.

References

  1. Mohan, R. Green bismuth. Nat. Chem. 2010, 2, 336. [Google Scholar] [CrossRef] [PubMed]
  2. Salvador, J.A.; Figueiredo, S.A.; Pinto, R.M.; Silvestre, S.M. Bismuth compounds in medicinal chemistry. Future Med. Chem. 2012, 4, 1495–1523. [Google Scholar] [CrossRef] [PubMed]
  3. Aguilera-Ruiz, E.; Zambrano-Robledo, P.; Vazquez-Arenas, J.; Cruz-Ortiz, B.; Peral, J.; García-Pérez, U.M. Photoactivity of nanostructured spheres of BiVO4 synthesized by ultrasonic spray pyrolysis at low temperature. Mater. Res. Bull. 2021, 143, 111447. [Google Scholar] [CrossRef]
  4. Koventhan, C.; Pandiyarajan, S.; Chen, S.-M. Simple sonochemical synthesis of flake-ball shaped bismuth vanadate for voltammetric detection of furazolidone. J. Alloys Compd. 2022, 895, 162315. [Google Scholar] [CrossRef]
  5. Packiaraj, R.; Devendran, P.; Asath Bahadur, S.; Nallamuthu, N. Structural and electrochemical studies of Scheelite type BiVO4 nanoparticles: Synthesis by simple hydrothermal method. J. Mater. Sci. Mater. Electron. 2018, 29, 13265–13276. [Google Scholar] [CrossRef]
  6. Liu, X.; Xiao, M.; Xu, L.; Miao, Y.; Ouyang, R. Characteristics, applications and determination of bismuth. J. Nanosci. Nanotechnol. 2016, 16, 6679–6689. [Google Scholar] [CrossRef]
  7. Zhang, A.; Zhang, J. Hydrothermal processing for obtaining of BiVO4 nanoparticles. Mater. Lett. 2009, 63, 1939–1942. [Google Scholar] [CrossRef]
  8. Trinh, D.T.T.; Khanitchaidecha, W.; Channei, D.; Nakaruk, A. Synthesis, characterization and environmental applications of bismuth vanadate. Res. Chem. Intermed. 2019, 45, 5217–5259. [Google Scholar] [CrossRef]
  9. Zhao, Z.; Li, Z.; Zou, Z. Electronic structure and optical properties of monoclinic clinobisvanite BiVO4. Phys. Chem. Chem. Phys. 2011, 13, 4746–4753. [Google Scholar] [CrossRef]
  10. Pookmanee, P.; Kojinok, S.; Puntharod, R.; Sangsrichan, S.; Phanichphant, S. Preparation and characterization of BiVO4 powder by the sol-gel method. Ferroelectrics 2013, 456, 45–54. [Google Scholar] [CrossRef]
  11. Josephine, A.J.; Dhas, C.R.; Venkatesh, R.; Arivukarasan, D.; Christy, A.J.; Monica, S.E.S.; Keerthana, S. Effect of pH on visible-light-driven photocatalytic degradation of facile synthesized bismuth vanadate nanoparticles. Mater. Res. Express 2020, 7, 015036. [Google Scholar] [CrossRef]
  12. Xu, H.; Zeiger, B.W.; Suslick, K.S. Sonochemical synthesis of nanomaterials. Chem. Soc. Rev. 2013, 42, 2555–2567. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Liu, W.; Cao, L.; Su, G.; Liu, H.; Wang, X.; Zhang, L. Ultrasound assisted synthesis of monoclinic structured spindle BiVO4 particles with hollow structure and its photocatalytic property. Ultrason. Sonochem. 2010, 17, 669–674. [Google Scholar] [CrossRef] [PubMed]
  14. Nguyen, T.D.; Hong, S.-S. Facile solvothermal synthesis of monoclinic-tetragonal heterostructured BiVO4 for photodegradation of rhodamine B. Catal. Commun. 2020, 136, 105920. [Google Scholar] [CrossRef]
  15. Venkatesan, R.; Velumani, S.; Ordon, K.; Makowska-Janusik, M.; Corbel, G.; Kassiba, A. Nanostructured bismuth vanadate (BiVO4) thin films for efficient visible light photocatalysis. Mater. Chem. Phys. 2018, 205, 325–333. [Google Scholar] [CrossRef]
  16. Nath, D.; Banerjee, P. Green nanotechnology—A new hope for medical biology. Environ. Toxicol. Pharmacol. 2013, 36, 997–1014. [Google Scholar] [CrossRef]
  17. Iravani, S.; Varma, R.S. Biofactories: Engineered nanoparticles via genetically engineered organisms. Green Chem. 2019, 21, 4583–4603. [Google Scholar] [CrossRef]
  18. Iravani, S.; Varma, R.S. Sustainable synthesis of cobalt and cobalt oxide nanoparticles and their catalytic and biomedical applications. Green Chem. 2020, 22, 2643–2661. [Google Scholar] [CrossRef]
  19. Iravani, S.; Varma, R.S. Greener synthesis of lignin nanoparticles and their applications. Green Chem. 2020, 22, 612–636. [Google Scholar] [CrossRef]
  20. Iravani, S.; Varma, R.S. Green synthesis, biomedical and biotechnological applications of carbon and graphene quantum dots. A review. Environ. Chem. Lett. 2020, 18, 703–727. [Google Scholar] [CrossRef]
  21. Ngullie, R.C.; Alaswad, S.O.; Bhuvaneswari, K.; Shanmugam, P.; Pazhanivel, T.; Arunachalam, P. Synthesis and Characterization of Efficient ZnO/g-C3N4 Nanocomposites Photocatalyst for Photocatalytic Degradation of Methylene Blue. Coatings 2020, 10, 500. [Google Scholar] [CrossRef]
  22. Nasrollahzadeh, M.; Sajjadi, M.; Iravani, S.; Varma, R.S. Trimetallic Nanoparticles: Greener Synthesis and Their Applications. Nanomaterials 2020, 10, 1784. [Google Scholar] [CrossRef] [PubMed]
  23. Nasrollahzadeh, M.; Sajjadi, M.; Iravani, S.; Varma, R.S. Green-synthesized nanocatalysts and nanomaterials for water treatment: Current challenges and future perspectives. J. Hazard. Mater. 2021, 401, 123401. [Google Scholar] [CrossRef] [PubMed]
  24. Khatami, M.; Iravani, S. Green and eco-friendly synthesis of nanophotocatalysts: An overview. Comments Inorg. Chem. 2021, 41, 133–187. [Google Scholar] [CrossRef]
  25. Tammina, S.K.; Mandal, B.K.; Kadiyala, N.K. Photocatalytic degradation of methylene blue dye by nonconventional synthesized SnO2 nanoparticles. Environ. Nanotechnol. Monit. Manag. 2018, 10, 339–350. [Google Scholar] [CrossRef]
  26. Mahanthappa, M.; Kottam, N.; Yellappa, S. Enhanced photocatalytic degradation of methylene blue dye using CuSCdS nanocomposite under visible light irradiation. Appl. Surf. Sci. 2019, 475, 828–838. [Google Scholar] [CrossRef]
  27. Paul, A.; Dhar, S.S. Construction of hierarchical MnMoO4/NiFe2O4 nanocomposite: Highly efficient visible light driven photocatalyst in the degradation of different polluting dyes in aqueous medium. Colloids Surf. A Physicochem. Eng. Asp. 2020, 585, 124090. [Google Scholar] [CrossRef]
  28. Fatima, U.; Khalid, N.R.; Nawaz, T.; Tahir, M.B.; Fatima, N.; Kebaili, I.; Alrobei, H.; Alzaid, M.; Shahzad, K.; Ali, A.M. Synthesis of BiVO4/NiFe2O4 composite for photocatalytic degradation of methylene blue. Appl. Nanosci. 2021, 11, 2793–2800. [Google Scholar] [CrossRef]
  29. Lee, G.-J.; Lee, X.-Y.; Lyu, C.; Liu, N.; Andandan, S.; Wu, J.J. Sonochemical Synthesis of Copper-doped BiVO4/g-C3N4 Nanocomposite Materials for Photocatalytic Degradation of Bisphenol A under Simulated Sunlight Irradiation. Nanomaterials 2020, 10, 498. [Google Scholar] [CrossRef] [Green Version]
  30. Letchumanan, D.; Sok, S.P.; Ibrahim, S.; Nagoor, N.H.; Arshad, N.M. Plant-Based Biosynthesis of Copper/Copper Oxide Nanoparticles: An Update on Their Applications in Biomedicine, Mechanisms, and Toxicity. Biomolecules 2021, 11, 564. [Google Scholar] [CrossRef]
  31. Pasula, R.R.; Lim, S. Engineering nanoparticle synthesis using microbial factories. Eng. Biol. 2017, 1, 12–17. [Google Scholar] [CrossRef]
  32. Luo, C.-H.; Shanmugam, V.; Yeh, C.-S. Nanoparticle biosynthesis using unicellular and subcellular supports. NPG Asia Mater. 2015, 7, e209. [Google Scholar] [CrossRef] [Green Version]
  33. Varma, R.S. Greener approach to nanomaterials and their sustainable applications. Curr. Opin. Chem. Eng. 2012, 1, 123–128. [Google Scholar] [CrossRef]
  34. Varma, R.S. Journey on greener pathways: From the use of alternate energy inputs and benign reaction media to sustainable applications of nano-catalysts in synthesis and environmental remediation. Green Chem. 2014, 16, 2027–2041. [Google Scholar] [CrossRef]
  35. Varma, R.S. Greener and sustainable chemistry. Appl. Sci. 2014, 4, 493–497. [Google Scholar] [CrossRef] [Green Version]
  36. Varma, R.S. Greener and Sustainable Trends in Synthesis of Organics and Nanomaterials. ACS Sustain. Chem. Eng. 2016, 4, 5866–5878. [Google Scholar] [CrossRef]
  37. Karatoprak, G.Ş.; Aydin, G.; Altinsoy, B.; Altinkaynak, C.; Koşar, M.; Ocsoy, I. The Effect of Pelargonium endlicherianum Fenzl. root extracts on formation of nanoparticles and their antimicrobial activities. Enzym. Microb. Technol. 2017, 97, 21–26. [Google Scholar] [CrossRef]
  38. Pramila, S.; Nagaraju, G.; Mallikarjunaswamy, C.; Latha, K.; Chandan, S.; Ramu, R.; Rashmi, V.; Lakshmi Ranganatha, V. Green Synthesis of BiVO4 nanoparticles by microwave method using Aegle marmelos juice as a fuel: Photocatalytic and antimicrobial study. Anal. Chem. Lett. 2020, 10, 298–306. [Google Scholar] [CrossRef]
  39. Mohamed, H.; Sone, B.; Khamlich, S.; Coetsee-Hugo, E.; Swart, H.; Thema, T.; Sbiaa, R.; Dhlamini, M. Biosynthesis of BiVO4 nanorods using Callistemon viminalis extracts: Photocatalytic degradation of methylene blue. Mater. Today Proc. 2021, 36, 328–335. [Google Scholar] [CrossRef]
  40. Baliga, M.S.; Shivashankara, A.R.; Haniadka, R.; Dsouza, J.; Bhat, H.P. Phytochemistry, nutritional and pharmacological properties of Artocarpus heterophyllus Lam (jackfruit): A review. Food Res. Int. 2011, 44, 1800–1811. [Google Scholar] [CrossRef]
  41. Mallikarjunaswamy, C.; Pramila, S.; Nagaraju, G.; Ramu, R.; Ranganatha, V.L. Green synthesis and evaluation of antiangiogenic, photocatalytic, and electrochemical activities of BiVO4 nanoparticles. J. Mater. Sci. Mater. Electron. 2021, 32, 14028–14046. [Google Scholar] [CrossRef]
  42. Manjunatha, A.S.; Pavithra, N.S.; Marappa, S.; Prashanth, S.A.; Nagaraju, G. Green synthesis of flower-like BiVO4 nanoparticles by solution combustion method using lemon (Citrus limon) juice as a fuel: Photocatalytic and electrochemical study. ChemistrySelect 2018, 3, 13456–13463. [Google Scholar] [CrossRef] [Green Version]
  43. Farag, M.A.; Paré, P.W. Phytochemical analysis and anti-inflammatory potential of Hyphaene thebaica L. fruit. J. Food Sci. 2013, 78, C1503–C1508. [Google Scholar] [CrossRef] [PubMed]
  44. Mohamed, H.E.A.; Afridi, S.; Khalil, A.T.; Zohra, T.; Alam, M.M.; Ikram, A.; Shinwari, Z.K.; Maaza, M. Phytosynthesis of BiVO4 nanorods using Hyphaene thebaica for diverse biomedical applications. AMB Express 2019, 9, 200. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Bhardwaj, R.; Nandal, U. Nutritional and therapeutic potential of bael (Aegle marmelos Corr.) fruit juice: A review. Nutr. Food Sci. 2015, 45, 895–919. [Google Scholar] [CrossRef]
  46. Burakova, E.A.; Dyachkova, T.P.; Rukhov, A.V.; Tugolukov, E.N.; Galunin, E.V.; Tkachev, A.G.; Basheer, A.A.; Ali, I. Novel and economic method of carbon nanotubes synthesis on a nickel magnesium oxide catalyst using microwave radiation. J. Mol. Liq. 2018, 253, 340–346. [Google Scholar] [CrossRef]
  47. Kumar, A.; Kuang, Y.; Liang, Z.; Sun, X. Microwave chemistry, recent advancements, and eco-friendly microwave-assisted synthesis of nanoarchitectures and their applications: A review. Mater. Today Nano 2020, 11, 100076. [Google Scholar] [CrossRef]
  48. Chen, M.; Chen, C.; Shi, S.; Chen, C. Low-temperature synthesis multiwalled carbon nanotubes by microwave plasma chemical vapor deposition using CH4–CO2 gas mixture. Jpn. J. Appl. Phys. 2003, 42, 614–619. [Google Scholar] [CrossRef]
  49. Daniel Abraham, S.; Theodore David, S.; Biju Bennie, R.; Joel, C.; Sanjay Kumar, D. Eco-friendly and green synthesis of BiVO4 nanoparticle using microwave irradiation as photocatalayst for the degradation of Alizarin Red S. J. Mol. Struct. 2016, 1113, 174–181. [Google Scholar] [CrossRef]
  50. Claudino, C.H.; Kuznetsova, M.; Rodrigues, B.S.; Chen, C.; Wang, Z.; Sardela, M.; Souza, J.S. Facile one-pot microwave-assisted synthesis of tungsten-doped BiVO4/WO3 heterojunctions with enhanced photocatalytic activity. Mater. Res. Bull. 2020, 125, 110783. [Google Scholar] [CrossRef]
  51. Zhang, Y.; Li, G.; Yang, X.; Yang, H.; Lu, Z.; Chen, R. Monoclinic BiVO4 micro-/nanostructures: Microwave and ultrasonic wave combined synthesis and their visible-light photocatalytic activities. J. Alloys Compd. 2013, 551, 544–550. [Google Scholar] [CrossRef]
  52. Souza, J.S.; Hirata, F.T.H.; Corio, P. Microwave-assisted synthesis of bismuth vanadate nanoflowers decorated with gold nanoparticles with enhanced photocatalytic activity. J. Nanopart. Res. 2019, 21, 35. [Google Scholar] [CrossRef]
  53. Pingmuang, K.; Nattestad, A.; Kangwansupamonkon, W.; Wallace, G.G.; Phanichphant, S.; Chen, J. Phase-controlled microwave synthesis of pure monoclinic BiVO4 nanoparticles for photocatalytic dye degradation. Appl. Mater. Today 2015, 1, 67–73. [Google Scholar] [CrossRef]
  54. 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]
  55. Kansaard, T.; Bangbai, C.; Jayasankar, C.K.; Pecharapa, W. Effect of Ultrasonic Irradiation Time on Physical Properties and Photocatalytic Performance of BiVO4 Nanoparticles Prepared via Sonochemical Process. Integr. Ferroelectr. 2021, 214, 123–132. [Google Scholar] [CrossRef]
  56. Tahir, M.B.; Iqbal, T.; Kiran, H.; Hasan, A. Insighting role of reduced graphene oxide in BiVO4 nanoparticles for improved photocatalytic hydrogen evolution and dyes degradation. Int. J. Energy Res. 2019, 43, 2410–2417. [Google Scholar] [CrossRef]
  57. 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]
  58. Lai, B.-R.; Lin, L.-Y.; Xiao, B.-C.; Chen, Y.-S. Facile synthesis of bismuth vanadate/bismuth oxide heterojunction for enhancing visible light-responsive photoelectrochemical performance. J. Taiwan Inst. Chem. Eng. 2019, 100, 178–185. [Google Scholar] [CrossRef]
  59. Reddy, C.V.; Reddy, I.N.; Koutavarapu, R.; Reddy, K.R.; Kim, D.; Shim, J. Novel BiVO4 nanostructures for environmental remediation, enhanced photoelectrocatalytic water oxidation and electrochemical energy storage performance. Sol. Energy 2020, 207, 441–449. [Google Scholar] [CrossRef]
  60. Sajid, M.M.; Amin, N.; Shad, N.A.; Bashir khan, S.; Javed, Y.; Zhang, Z. Hydrothermal fabrication of monoclinic bismuth vanadate (m-BiVO4) nanoparticles for photocatalytic degradation of toxic organic dyes. Mater. Sci. Eng. B 2019, 242, 83–89. [Google Scholar] [CrossRef]
  61. Ganeshbabu, M.; Kannan, N.; Sundara Venkatesh, P.; Paulraj, G.; Jeganathan, K.; MubarakAli, D. Synthesis and characterization of BiVO4 nanoparticles for environmental applications. RSC Adv. 2020, 10, 18315–18322. [Google Scholar] [CrossRef] [PubMed]
  62. Sharma, R.; Uma; Singh, S.; Verma, A.; Khanuja, M. Visible light induced bactericidal and photocatalytic activity of hydrothermally synthesized BiVO4 nano-octahedrals. J. Photochem. Photobiol. B Biol. 2016, 162, 266–272. [Google Scholar] [CrossRef] [PubMed]
  63. Lopes, O.F.; Carvalho, K.T.; Macedo, G.K.; de Mendonca, V.R.; Avansi, W.; Ribeiro, C. Synthesis of BiVO 4 via oxidant peroxo-method: Insights into the photocatalytic performance and degradation mechanism of pollutants. New J. Chem. 2015, 39, 6231–6237. [Google Scholar] [CrossRef]
  64. Dowla Biswas, M.R.U.; Ho, B.S.; Oh, W.-C. Eco-friendly conductive polymer-based nanocomposites, BiVO4/graphene oxide/polyaniline for excellent photocatalytic performance. Polym. Bull. 2020, 77, 4381–4400. [Google Scholar] [CrossRef]
  65. Tahir, M.B.; Riaz, K.N.; Asiri, A.M. Boosting the performance of visible light-driven WO3/g-C3N4 anchored with BiVO4 nanoparticles for photocatalytic hydrogen evolution. Int. J. Energy Res. 2019, 43, 5747–5758. [Google Scholar] [CrossRef]
  66. Dhabarde, N.; Carrillo-Ceja, O.; Tian, S.; Xiong, G.; Raja, K.; Subramanian, V.R. Bismuth Vanadate Encapsulated with Reduced Graphene Oxide: A Nanocomposite for Optimized Photocatalytic Hydrogen Peroxide Generation. J. Phys. Chem. C 2021, 125, 23669–23679. [Google Scholar] [CrossRef]
  67. Kansaard, T.; Pecharapa, W. Characterization of BiVO4 nanoparticles prepared by sonochemical process. Ferroelectrics 2019, 552, 140–147. [Google Scholar] [CrossRef]
  68. Tian, H.; Wu, H.; Fang, Y.; Li, R.; Huang, Y. Hydrothermal synthesis of m-BiVO4/t-BiVO4 heterostructure for organic pollutants degradation: Insight into the photocatalytic mechanism of exposed facets from crystalline phase controlling. J. Hazard. Mater. 2020, 399, 123159. [Google Scholar] [CrossRef]
  69. Jiang, H.; Dai, H.; Meng, X.; Zhang, L.; Deng, J.; Liu, Y.; Au, C.T. Hydrothermal fabrication and visible-light-driven photocatalytic properties of bismuth vanadate with multiple morphologies and/or porous structures for Methyl Orange degradation. J. Environ. Sci. 2012, 24, 449–457. [Google Scholar] [CrossRef]
  70. Choe, H.R.; Kim, J.H.; Ma, A.; Jung, H.; Kim, H.Y.; Nam, K.M. Understanding Reaction Kinetics by Tailoring Metal Co-catalysts of the BiVO4 Photocatalyst. ACS Omega 2019, 4, 16597–16602. [Google Scholar] [CrossRef]
  71. Ma, Y.; Pendlebury, S.R.; Reynal, A.; Formal, F.L.; Durrant, J.R. Dynamics of photogenerated holes in undoped BiVO4 photoanodes for solar water oxidation. Chem. Sci. 2014, 5, 2964–2973. [Google Scholar] [CrossRef] [Green Version]
  72. Yan, Y.; Sun, S.; Song, Y.; Yan, X.; Guan, W.; Liu, X.; Shi, W. Microwave-assisted in situ synthesis of reduced graphene oxide-BiVO4 composite photocatalysts and their enhanced photocatalytic performance for the degradation of ciprofloxacin. J. Hazard. Mater. 2013, 250–251, 106–114. [Google Scholar] [CrossRef] [PubMed]
  73. Li, Z.; Bao, Z.; Yao, F.; Cao, H.; Wang, J.; Qiu, L.; Lv, J.; Sun, X.; Zhang, Y.; Wu, Y. One-dimensional bismuth vanadate nanostructures constructed Z-scheme photocatalyst for highly efficient degradation of antibiotics. J. Water Process Eng. 2022, 46, 102599. [Google Scholar] [CrossRef]
  74. Hemavibool, K.; Sansenya, T.; Nanan, S. Enhanced Photocatalytic Degradation of Tetracycline and Oxytetracycline Antibiotics by BiVO4 Photocatalyst under Visible Light and Solar Light Irradiation. Antibiotics 2022, 11, 761. [Google Scholar] [CrossRef]
Figure 1. The mechanism of antibacterial effects of monoclinic BiVO4 nanomaterials against pathogenic bacteria. Adapted from [62] with permission. Copyright 2016 Elsevier.
Figure 1. The mechanism of antibacterial effects of monoclinic BiVO4 nanomaterials against pathogenic bacteria. Adapted from [62] with permission. Copyright 2016 Elsevier.
Catalysts 13 00059 g001
Figure 2. (A) The encapsulation of BiVO4 (BVO) with reduced graphene oxide (rGO). (B) Digital photographs of the commercial BVO and the prepared composites. (C) Mechanism of photocatalytic generation of H2O2. Adapted from [66] with permission. Copyright 2021 American Chemical Society.
Figure 2. (A) The encapsulation of BiVO4 (BVO) with reduced graphene oxide (rGO). (B) Digital photographs of the commercial BVO and the prepared composites. (C) Mechanism of photocatalytic generation of H2O2. Adapted from [66] with permission. Copyright 2021 American Chemical Society.
Catalysts 13 00059 g002
Figure 3. The proposed mechanism for antibiotic photodegradation (tetracycline and ciprofloxacin) using BiVO4/reduced graphene oxide (RGO)/g-C3N4 Z-scheme photocatalysts. Adapted from [73] with permission. Copyright 2022 Elsevier.
Figure 3. The proposed mechanism for antibiotic photodegradation (tetracycline and ciprofloxacin) using BiVO4/reduced graphene oxide (RGO)/g-C3N4 Z-scheme photocatalysts. Adapted from [73] with permission. Copyright 2022 Elsevier.
Catalysts 13 00059 g003
Table 1. Some selected examples of BiVO4 NPs synthesized using eco-friendly techniques.
Table 1. Some selected examples of BiVO4 NPs synthesized using eco-friendly techniques.
SaltsChelating AgentsCrystal StructureShape of NPsPlant PhytochemicalsApplicationsRefs.
Bi(NO3)3
and VOSO4
Flower extract of Callistemon viminalis (bottlebrush)Monoclinic scheeliteNanorods (the basal and longitudinal dimensions of the nanorods ranged from 350 to 450 nm and 1.2–2 μm, respectively)Flavonoids, saponins, alkaloids, steroids and triterpenoidsPhotocatalytic activity (methylene blue)[39]
NH4VO3
and Bi(NO3)3
Fruit extract of Unripe jackfruitMonoclinic scheeliteThe asymmetrically arranged BiVO4 nanostructures (the shapes were nearly spherical and hexagonal); the size being ~90–250 nmCarotenoids, flavonoids, volatile acids sterols and
tannins
stilbenoids, and arylbenzofurons
Anticancer (breast cancer cell lines), Photocatalytic activity (Methylene blue) and electrochemical sensor[40,41]
-Citrus Limon
(lemon)
Pure monoclinicFlower-like structures Citric acidPhotocatalytic activity (Indigo Carmine) and electrochemical sensor[42]
Bi(NO3)3
and VOSO4
Fruit extract of Hyphaene thebaicaClinobisvanite monoclinicNanorods (well-aligned rod shaped)Cinnamic acid, flavonoids, vanillic acid, epicatechin, glycosides, stilbene and sugars composition Antibacterial, antifungal, and antiviral activity [43,44]
NH4VO3
and Bi(NO3)3
Fruit extract of Aegle marmelos (bael)Monoclinic scheelite-Flavonoids,
polyphenols tannins, alkaloids, coumarins, steroids and natural sugar
Antibacterial, antifungal, and photocatalytic activity (methylene blue)[38,45]
Table 2. Selected examples of photocatalytic degradation of pollutants using biosynthesized BiVO4 NPs under visible light irradiation.
Table 2. Selected examples of photocatalytic degradation of pollutants using biosynthesized BiVO4 NPs under visible light irradiation.
Shape and Crystal Structure of NPsPlant ExtractsPollutantsDegradation Efficiency (%)Absorption Peak (nm)Gap Band (eV)Refs.
Monoclinic scheelite and nanorodsCallistemon viminalisMethylene blue82.636612.59[39]
Monoclinic scheeliteAegle marmelosMethylene blue906632.5[38]
Monoclinic scheelite and quasispherical-like structureUnripe jackfruitMethylene blue98.3250–3002.4[41]
Flower-like
and monoclinic scheelite
Citrus LimonIndigo Carmine90.6610∼2.6–2.8[42]
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

Q. Alijani, H.; Iravani, S.; Varma, R.S. Bismuth Vanadate (BiVO4) Nanostructures: Eco-Friendly Synthesis and Their Photocatalytic Applications. Catalysts 2023, 13, 59. https://doi.org/10.3390/catal13010059

AMA Style

Q. Alijani H, Iravani S, Varma RS. Bismuth Vanadate (BiVO4) Nanostructures: Eco-Friendly Synthesis and Their Photocatalytic Applications. Catalysts. 2023; 13(1):59. https://doi.org/10.3390/catal13010059

Chicago/Turabian Style

Q. Alijani, Hajar, Siavash Iravani, and Rajender S. Varma. 2023. "Bismuth Vanadate (BiVO4) Nanostructures: Eco-Friendly Synthesis and Their Photocatalytic Applications" Catalysts 13, no. 1: 59. https://doi.org/10.3390/catal13010059

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