Nanotechnology provides a very good chance to research and develop multipurpose nanomaterials because of their smaller size, larger surface area, low cost, and nanoscale materials, which are auspicious tools for many biological applications. The AgVO3 nanoparticle synthesis employing plant extract has offered an eco-friendly alternative for the industry. Literature survey shows that no research has been performed on AgVO3 using jackfruit; hence, we prepared AgVO3 using jackfruit extract as a reducing agent by a simple, easy, and eco-friendly precipitation method. The characterization techniques used for synthesized AgVO3 nanoparticles were x-ray diffraction analysis, which determines the monoclinic structure of synthesized AgVO3 nanoparticles; Fourier transform infrared spectroscopy, which shows the bonding of V–O–V; and scanning electron microscopy and energy dispersive spectra (EDS), which confirm the size, shape, purity, and elemental composition. Brunauer–Emmett–Teller analysis confirms the pore size, pore volume, and surface area of synthesized AgVO3 nanoparticles, Raman studies show the crystalline property, and UV–Vis studies give information about the material’s formation and optical properties. The bandgap was calculated to be 2.54 eV. Furthermore, the photocatalytic studies show 98.14% degradation in 180 min using MB dye. We also performed scavenger studies for detection of OH· radicals and recyclability. Gram-negative (Klebsiella pneumonia and Pseudomonas aeruginosa) and gram-positive (Staphylococcus aureus and Bacillus subtilis) micro-organisms were used to determine the antimicrobial characteristics. The full analysis verifies AgVO3’s antibacterial activity against both gram-negative and gram-positive bacteria, as well as its excellent photocatalytic activities for the degradation of the organic dye methylene blue with a high degree of recyclability.

The expeditious growth of urbanization and industrialization has unfavorably contaminated our precious nature and natural resources, leading to health problems in living organisms, mainly in human beings. Primarily, water pollution due to waste effluents from the textile industry and pathogenic bacteria and other microbes results from the mixing of fecal materials, industrial and domestic sewage, pasture, and agricultural runoff with drinking water. Due to these, several infectious diseases, such as typhoid, cholera, and dysentery, affect human beings. Due to their toxic nature, organic dyes are considered one of the main poisons in industrial wastewater among chemical contaminants. Some of these dyes are extremely poisonous and pose serious harm to the ecosystem in the area.1 Thus, we need technology that can monitor, detect, and, if possible, clean the pollutants from the water. Research into the science and technology of diverse nanostructures has greatly increased as a result of the current circumstances, and this technology is rapidly evolving and will have a big impact on the future of materialization.2 Conventional wastewater treatment technologies have a number of major drawbacks, including cost, energy use, maintenance and operation costs, transportation and storage issues, time requirements, equipment handling, etc.

Nanotechnology has been incorporated into cleaner industrial processes and the creation of environmentally friendly products because it plays a priceless role in the purification of the environment through the detection, prevention, and removal of hazardous pollutants. For wastewater treatment, such as photocatalyst degradation of a variety of organic pollutants in wastewater, such as detergents, dyes, pesticides, etc., nano-catalysts and catalytic membranes, nano-adsorbents, nano-membranes, bioactive nanoparticles, and biomimetic membranes are used.3 

The synthesis of nanomaterials using green nanotechnology is an evolving research trend in current days because this technique is eco-friendly, non-toxic or less-toxic, and economical as compared to other current methods (physico-chemical). Various living entities, such as plants, fungi, bacteria, and algae, benefit from this nanoparticle production.4 However, researchers from all over the biosphere are now concentrating more on plant-based nanoparticle synthesis. Due to their great efficiency and biocompatibility, nanoparticles are now used in a number of characterization approaches for water and wastewater treatment. When wastewater is purified for reuse and recycled via green nanoparticle synthesis, heavy metals from waste water can be recovered and eliminated, as well as a range of organic pollutants. This helps address numerous global problems with water quality.5 These days, research has concentrated on the creation and evaluation of nanoparticles such as ZnO,6 Zr3(PO4)4,7 NiO,8 CuO,9 Cr2O3,10 Ag2O,11 BiOCl,12 etc., for their physicochemical and biological properties.

In this situation, nanotechnology offers a bright spot and a glimmer of hope for resolving these problems and preserving our ecosystem. In current years, green synthesis of silver vanadate nanoparticles by plants such as Tagetes erecta,13, Selaginella bryopteris,14, Tabernaemontana divaricata,15, Phoenix dactylifera,16, Alchornea laxiflora,17, Trigonella foenum-graecum seed extract,18 capsicum Chinese plant,19, Conocarpus lancifolius plant extract,20, Lantana camara leaf extract,21  Aegle marmelos,7 and several others has been reported to have photodegradation and antimicrobial properties.

Green synthesis of iron,22 Ag@Au,23 Ag,24 ZnO,25 Bi2O3,26 and other nanoparticles has been performed using jackfruit extract. Among these, AgVO3 received more attention for its specific properties, such as photocatalysis,27 gas sensing,28 use in lithium-ion batteries,29 and antimicrobial30 and optoelectronic properties.31 

To the best of our knowledge and after a thorough literature review, no effort has been made so far to prepare silver vanadate nanoparticles using jackfruit extract by the precipitation method. The obtained silver vanadate nanoparticles were characterized by scanning electron microscopy (SEM), x-ray diffraction analysis (XRD), the Brunauer–Emmett–Teller (BET) method, a Fourier transmission infrared spectrometer (FTIR), energy dispersive spectra (EDS), and Raman spectra. Following production, AgVO3 nanoparticles were treated to evaluate their photodegradation capacity of the methylene blue dye by varying the different parameters, such as dye concentration, catalyst concentration, pH, and detection of ·OH radicals. Antimicrobial studies of gram-negative and gram-positive bacteria were also carried out. A comparison table of nanoparticles from green synthesis and their photodegradation efficiency with applications is shown in Table I.

TABLE I.

Comparative table of nanoparticles from green synthesis and their photodegradation efficiency and applications.

Degradation efficiency
NP’sGreen plantDyeTime%ApplicationsReferences
Ag Selaginella bryopteris MB 7 hr 100 Photocatalytic and antimicrobial 14  
Ag/Ag2Phoenix dactylifera Azo 50 min 84.5 Photodegradation 16  
Cango 
MB 
Ag Trigonella foenum-graecum Rhodamine-B 216 hr 93 Photocatalytic and antimicrobial 18  
AgVO3 Jackfruit extract MB 120 min 98.4 Photocatalytic and antimicrobial Present work 
Degradation efficiency
NP’sGreen plantDyeTime%ApplicationsReferences
Ag Selaginella bryopteris MB 7 hr 100 Photocatalytic and antimicrobial 14  
Ag/Ag2Phoenix dactylifera Azo 50 min 84.5 Photodegradation 16  
Cango 
MB 
Ag Trigonella foenum-graecum Rhodamine-B 216 hr 93 Photocatalytic and antimicrobial 18  
AgVO3 Jackfruit extract MB 120 min 98.4 Photocatalytic and antimicrobial Present work 

All chemicals were pure and of analytical reagent grade. NH4VO3, AgNO3, and NaOH were purchased from S.D. Fine Chemicals Ltd., and jackfruit extract was used to synthesize the AgVO3 nanoparticles by the precipitation method. Double-distilled water was used throughout the experiment.

Healthy, unripe, and unharmed jackfruits were identified and procured in Bengaluru from the Indian Institute of Horticultural Research (IIHR). After that, the fruit pulp was dried for 15 days in the shade. To obtain jackfruit powder, the resulting dried fruit pulp underwent additional processing by being homogenized in a blender. A Soxhlet apparatus was then used to extract from the 5 g of powder over the course of 15 cycles.

To make AgVO3 nanoparticles by precipitation, ammonium metavanadate (0.1M), a silver source, and aqueous jackfruit extract were used. The synthesis was started by adding 5 ml of jackfruit extract to 20 ml of NH4VO3 solution (0.1M) in a beaker. At a temperature of 50 °C, the mixture was gently stirred for 30 min. The beaker was then filled with 20 ml of silver nitrate solution (0.1M). NaOH was cautiously applied in order to maintain a pH level of 6. The resulting precipitate was extensively washed with distilled water to remove any impurities. The AgVO3 precipitate was then dried at 100 °C for ∼5 h. The resulting AgVO3 nanoparticles were subsequently used for further characterization and other applications. pH was controlled and the acidic condition was maintained to prevent the reduction of Ag+ to Ag. The synthesis pathway was adopted to get silver vanadate by controlling the pH, metal concentration, aging time, and drying temperature. Therefore, the methodology detailed here describes the optimum conditions to get the coveted metalate, limiting the polymerization of vanadium ions and the precipitation of silver oxides. The secondary phase of these potential implications does not affect the photocatalytic and antimicrobial applications of synthesized AgVO3.

The crystallinity and structure of the AgVO3 nanoparticles were assessed using an x-ray diffractometer (Rigaku SmartLab) furnished with monochromatic high-intensity Cu-Kα radiation, where λ (lambda) equals 1.5418 Å. Using a TESCAN Vega 3LMV instrument, SEM was carried out to study the material’s surface morphology. Through energy-dispersive x-ray diffraction analysis coupled with SEM, the purity and elemental composition were ascertained. In addition, an Agilent Cary 60 UV–Visible spectrophotometer was used to record the UV–Visible spectra of the degraded solution. Using nitrogen adsorption at −196 °C, the BET-specific surface area of the Ag vanadate was determined using Quanta Chrome Instruments Version 10.01. Using an Ar+ ion laser (514.5 nm laser line) with a power-varying laser density, measurements of Raman spectroscopy were performed. To assess the scattered light, a Jobin Yvon T64000 spectrometer with a cooled charge-coupled device (CCD) detector was utilized.

The photocatalytic activity of AgVO3 as a catalyst for degrading the organic dye methylene blue in water was investigated using visible light. A HEBER scientific photo-reactor was used to increase the degradation’s effectiveness. This reactor has fans to dissipate the heat produced by the light source and a water jacket to block out visible light, which may affect how quickly the methylene blue dye degrades. 100 ml of a 5 ppm methylene blue solution was placed in a 9-in. quartz tube, along with 50 mg of photocatalyst, for the degradation process. The mixture was given 30 min to equilibrate in the dark in order to reach adsorption–desorption equilibrium. A water-cooled condenser surrounds a high-pressure mercury lamp (250 W, 365 nm) that is installed inside a quartz chamber and, at about 12 cm, separates the solution from the light source. After a 20-min interval, samples were removed from the reactor and centrifuged to separate the photocatalyst and the solution. After each time interval, the amount of absorption was measured with a UV–Visible spectrophotometer to determine the level of degradation. It is important to note that a number of factors, including catalyst load, pH variations, and varying dye concentrations, affect the degradation efficiency. In order to achieve high degradation efficiency, all of these restrictions were methodically investigated and improved in this study.

The effectiveness of antimicrobial screening against gram-negative (Klebsiella pneumonia and Pseudomonas aeruginosa) and gram-positive (Staphylococcus aureus and Bacillus subtilis) bacterial strains has been investigated using three different dosages of AgVO3 nanoparticles.

1. XRD studies

The XRD pattern of the AgVO3 nanoparticles presented in Fig. 1 confirms the monoclinic crystal structure of AgVO3 within a range of 5°–80°. The diffraction angle values observed at 28.8°, 30°, 32°, 33°, 34°, 35°, and 51° indicate the presence of AgVO3 (−211), (501), (−411), (−112), (−602), (112), and (020) nanoparticles (JCPDS No. 29–1154).32 The intensity and crystallinity of the diffraction peaks were slightly affected by pH. The presence of the Ag phase (JCPDS 870597) was also detected. The full width at half maximum (FWHM) of the AgVO3 nanoparticles was investigated to identify if their regular crystallite size matches the strongest diffraction angle at 30°. The crystalline size was then estimated by combining this knowledge with the Debye–Scherrer equation, which is shown below:
where k is the dimensionless shape factor (0.89), D is the crystallite size of the samples, β is the full width at the half maximum of the line in radians, and λ represents the x-ray wavelength and gives the diffraction angle; therefore, the prepared AgVO3 sample crystallite size is calculated as 87.29 nm.
FIG. 1.

(a) Standard curve of AgVO3 and (b) XRD pattern of the standard curve of AgVO3 nanoparticles prepared using jackfruit extract.

FIG. 1.

(a) Standard curve of AgVO3 and (b) XRD pattern of the standard curve of AgVO3 nanoparticles prepared using jackfruit extract.

Close modal

Figure 2 shows how the functional groups in the AgVO3 nanoparticles were detected using the Fourier transform infrared (FTIR) approach. The study was performed between 4000 and 500 cm−1. Notably, the appearance of three separate peaks at 960, 750, and 600 cm−1 indicates the presence of (V–O–V) bonds, in which two vanadium atoms are connected to the central oxygen atom. Asymmetric stretching modes are linked to these peaks. In addition, new peaks were found at 2100 and 2300 cm−1, which indicate the presence of C–C double bonds, and at 1400 cm−1, which indicate the presence of a C–C single bond. These results help characterize the molecular structure of the AgVO3 nanoparticles by offering important insights into the chemical composition and bonding inside them.

FIG. 2.

FTIR spectra of the AgVO3 nanoparticle.

FIG. 2.

FTIR spectra of the AgVO3 nanoparticle.

Close modal

1. Scanning electron microscopy (SEM)

Scanning electron microscopy (SEM) was used to examine the morphology of the synthetic AgVO3 nanoparticles. The SEM analysis yielded nanographs of the AgVO3 nanoparticles, all of which consistently revealed a distinctive feature: rod-shaped structures. Through SEM analysis, we were able to determine that the Ag vanadate nanoparticles exhibited a heterogeneous morphology composed of Ag vanadate rods 100 nm in width. These SEM images of AgVO3 nanoparticles are presented in Fig. 3(a), providing clear visual evidence of their characteristic morphology.

FIG. 3.

(a) SEM and (b) EDS images of AgVO3 nanoparticles prepared using jackfruit. (c) Elemental mapping of AgVO3 nanoparticles.

FIG. 3.

(a) SEM and (b) EDS images of AgVO3 nanoparticles prepared using jackfruit. (c) Elemental mapping of AgVO3 nanoparticles.

Close modal

2. Elemental mapping

The EDS spectra presented in Figs. 3(b) and 3(c) confirms the existence of signals corresponding to silver, vanadium, and oxygen. This observation indicates the purity of the prepared AgVO3 sample. Moreover, the quantitative analysis reveals a stoichiometric ratio for AgVO3 of 11.72:12.03:76.25, which closely aligns with the theoretical value.

An important analysis method for measuring a material’s specific surface area is based on the Brunauer–Emmett–Teller (BET) hypothesis, which attempts to explain the physical adsorption of gas molecules on a solid surface.33 Surface parameters such as pore size, volume, and area have a significant impact on nanomaterial qualities. The BET technique is very important in elucidating the surface parameters of the nanoparticles. The BET surface area was evaluated by adsorption or physisorption of N2, utilizing a surface analysis device (Beishide, 3H–2000PSI, China). Figures 4(a)−4(c) depict the N2 adsorption/desorption isotherm, pore size distribution of AgVO3 nanomaterials, and BET surface area plot of AgVO3 nanoparticles. This figure also shows the type IV Langmuir hysteresis curve, depending on the form of the pore in the manufactured sample, which has a surface area of 25.8 m2 g−1.

FIG. 4.

(a) Nitrogen adsorption/desorption isotherm, (b) pore volume, and (c) BET surface area plot of AgVO3 nanoparticles.

FIG. 4.

(a) Nitrogen adsorption/desorption isotherm, (b) pore volume, and (c) BET surface area plot of AgVO3 nanoparticles.

Close modal

Raman scattering is utilized to analyze the crystalline state of nanoparticles as well as to discover defects in the manufactured AgVO3 nanomaterials. Figure 5 depicts the Raman scattering spectrum of AgVO3 nanoparticles produced with jackfruit extract as fuel. The large peaks at 336 and 442 cm1 correspond to the hexagonal structure of AgVO3. The high signal at 884 cm1 indicates monoclinic AgVO3 nanomaterials with a mixed phase. The Raman bands at 383 and 334 cm1 are ascribed to the asymmetric deformation modes of the VO43-tetrahedron.34–36 

FIG. 5.

Raman plot of AgVO3 nanoparticles.

FIG. 5.

Raman plot of AgVO3 nanoparticles.

Close modal
UV-Visible investigations are one of the most efficient methods for determining the initial presence of metal oxide nanoparticles in an aqueous medium and are used to examine the optical characteristics of the produced AgVO3 nanomaterials. Since nanoparticles have a specific absorption wavelength as a result of their distinctive optical performance, this method is unique, where it determines the bandgap energy of a material. Figure 6 demonstrates that the host lattice is responsible for the significant absorption peak at 487 nm, indicating that UV radiation can stimulate AgVO3 nanoparticles. The bandgap energy of AgVO3 nanoparticles can therefore be calculated from the absorbed wavelength value using this absorption wavelength and the energy equation,37 
where E is the Energy, h is Planck’s constant (6.626 × 10−34 J s−1), c is the velocity of light (2.99 × 108 m/s), and λ is the wavelength of the absorption peak. In Fig. 6, an absorption band at a wavelength of 487 nm was seen. The bandgap value was computed using the above-mentioned equation and was determined to be 2.54 eV. The UV–Vis spectrum also supported this result. The small particle size is presumably due to this broad bandgap of 2.54 eV.
FIG. 6.

UV–Visible spectrum of AgVO3.

FIG. 6.

UV–Visible spectrum of AgVO3.

Close modal

The photocatalytic activity involves four steps: first, electrons and holes are generated by the light irradiation; second, charge carriers cause separation and migration of the semiconductor to the surface; third, the involvement of the generated electrons and holes in the oxidation and reduction reactions to intermediate products; finally, the recombination of charge carriers on the photocatalytic surface. The photocatalytic activity depends on many factors, such as crystallite size, surface area, and mainly the bandgap. One of the promising options for improving photocatalytic ability is silver vanadate, primarily because optical absorption can be extended to the visible light region and effectively reduce the recombination of the photogenerated electrons and holes. This allows for the development of an efficient way to fully utilize solar light at higher photocatalytic activity.38,39 This happens as a result of effective absorption centers in the vanadate groups, VO3, where electronic transitions from oxygen (2p orbital) to V5+ (3d orbital) ions can be easily generated.40 

The synthesized AgVO3 nanoparticles were tested for the degradation of other pollutants such as rhodamine-B (Rh–B), methylene blue (MB), methyl orange (MO), and rose bengal (RB). The below-mentioned graph shows the degradation efficiency of synthesized AgVO3 nanoparticles. The degradation study has been performed for 3 h with an interval of 30 min (Fig. 7). The degradation efficiency of synthesized AgVO3 nanoparticles after 180 min for Rh–B, MB, MO, and RB was found to be 75.89%, 98.15%, 68.14%, and 68.36%, respectively. Therefore, methylene dye has been selected for further photocatalytic study.

FIG. 7.

Photodegradation of synthesized AgVO3 nanoparticles for different dyes.

FIG. 7.

Photodegradation of synthesized AgVO3 nanoparticles for different dyes.

Close modal

AgVO3 nanoparticles were tested in a series of doses (10, 20, 30, and 40 mg), and the results are shown in Fig. 8. Notably, the elimination efficiency increased to an astonishing 98% at a dosage of 20 mg. However, the elimination efficiency showed a modest decline to 40% as the dosage was increased to 30 and 40 mg. This behavior is explained by the extraordinary ability of AgVO3 nanoparticles to absorb light, which promotes the formation of photogenerated electrons and holes. These charge carriers create superoxide and hydroxyl radicals when they interact with water and dissolved oxygen. However, as the dosage increases, light transmittance decreases, decreasing how well light is used and, as a result, lowering photodegradation efficiency. Furthermore, large dosages may cause nanoparticles to aggregate, further reducing their effectiveness.

FIG. 8.

Photodegradation with different photocatalyst dosages of AgVO3 nanoparticles.

FIG. 8.

Photodegradation with different photocatalyst dosages of AgVO3 nanoparticles.

Close modal

Figure 9 shows the effect of dye concentration on the percentage of dye degradation under visible light. Different concentrations of MB dye, such as 5, 10, 15, and 20 ppm for 100 ml of the reaction mixture, were tested by maintaining the AgVO3 nanoparticles constant at 20 mg. Figure 9 shows that photodegradation activity gradually decreases, with efficiency decreasing from 98% at 5 ppm to 50% at 20 ppm. Notably, the most efficient amount of effective degradation was attained at 5 ppm with a 98% efficiency. This finding shows that the effectiveness of dye degradation decreases as dye molecule concentrations increase because they tend to adsorb more thickly on the catalyst’s surface. The increase in methylene blue dye concentration also leads to reduced light penetration, which is a likely contributing factor to the decreased degradation of methylene blue.41 

FIG. 9.

Photodegradation of AgVO3 at different concentrations of methylene blue dye.

FIG. 9.

Photodegradation of AgVO3 at different concentrations of methylene blue dye.

Close modal

The pH of a reaction mixture is responsible for the production of reactive species, which are the reason for the degradation of the organic dye. Hence, we explored by varying the pH of a reaction mixture at 3, 5, 7, 9, and 11, while the catalyst is 20 mg and the concentration of dye is 5 ppm. At a pH of 11, synthesized AgVO3 shows maximum degradation at its peak, as shown in Fig. 10.

FIG. 10.

Photodegradation of AgVO3 nanoparticles at different pH values.

FIG. 10.

Photodegradation of AgVO3 nanoparticles at different pH values.

Close modal
In comparison to monostructure nanoparticle semiconductors, heterostructure nanoparticle semiconductors exhibit greater photocatalytic activity and a greater ability to distinguish photogenerated holes and electrons.16 When visible light is used to irradiate AgVO3 nanoparticles, photogenerated electrons and holes are created. Highly reactive oxygen species such as superoxides, peroxides, and hydroxyl radicals are produced as a result of these charge carriers actively participating in redox processes, which involve oxidizing water and reducing oxygen. The oxidative breakdown of the organic dye methylene blue, which ultimately leads to the dye’s entire mineralization into carbon dioxide and water, depends on these photogenerated reactive oxygen species, as shown in Fig. 11. By absorbing light, the produced AgVO3 nanoparticles considerably improve the photocatalytic degradation study of the organic dye methylene blue. Particularly when exposed to UV light, this absorption mechanism efficiently generates photo-induced electron–hole pairs. Figure 11 clearly illustrates how important a role AgVO3 nanoparticles’ catalytic activity plays in preventing the recombination of these electron–hole pairs. AgVO3 nanoparticles’ surface plasmon resonance action results in a considerable UV light absorption capability. Superoxide ions (O2) are produced when photogenerated electrons are successfully absorbed by electron acceptors such as dissolved oxygen, which also helps the methylene blue dye degrade. The oxidation of methylene blue and a reaction with water result in hydroxyl radicals (OH) and H+ ions, and both are being carried out simultaneously by the photogenerated holes. A series of hydroxyl radicals (.OH) are created as a result of the subsequent reaction between H+ ions and O2. The production of all of these reactive oxygen species accelerates the photodegradation of the organic dye methylene blue significantly, according to Ref. 42. A schematic illustration of the proposed mechanism is provided below:
FIG. 11.

Diagrammatic illustration of the degradation mechanism.

FIG. 11.

Diagrammatic illustration of the degradation mechanism.

Close modal

A radical scavenging activity experiment was carried out to better understand the contributions provided by reactive oxygen species in the photodegradation activity of AgVO3 nanomaterials during the breakdown of methylene blue organic dye. Figure 14 shows the outcomes of this experiment, which used the scavengers EDTA, TBA, and K2Cr2O7.

The degrading efficiency of AgVO3 nanoparticles can be significantly affected by the pH of the solution during the photocatalytic process. It can improve the methylene blue dye’s surface adsorption onto the photocatalyst. In addition, both H+ and OH play a role in the production of reactive oxygen species, which is directly influenced by the pH levels in the solution. Therefore, a thorough examination of the pH effect was conducted over the pH range of 3–11, and the outcomes are shown in Fig. 10.

In order to investigate the reactive species that are accountable for the degradation of methylene blue organic dye by the AgVO3 nanoparticles, to check the role of the generated active species in the degradation of MB, various scavengers were employed. These scavengers serve to capture and reduce the production of reactive species, such as holes (h+), hydroxyl radicals (·OH), and superoxide anion radicals (O2). As shown in Fig. 12, in our experiment, we used the scavengers tert-butyl alcohol (TBA, 1 mM), potassium dichromate (1 mM), and ethylenediaminetetraacetic acid disodium (EDTA-2Na) to target holes (h+), superoxides (O2), and hydroxyl radicals (·OH), respectively. The results conveyed that the percentage of degradation decreases in the presence of TBA. Hence, this indicates that hydroxyl radicals play a major role in the degradation of organic dye. Figure 13 shows the EPR spectra for the detection of oxygen radicals. This technique was used to identify the reactive species during the reaction. There was no O2 found when the reaction mixture was kept under dark conditions. Then, under visible light for 10 min, the intensities were detected, which suggests agreement with the results of the radical trapping test.43 

FIG. 12.

Photodegradation activity of the AgVO3 nanoparticles on methylene blue with different scavengers.

FIG. 12.

Photodegradation activity of the AgVO3 nanoparticles on methylene blue with different scavengers.

Close modal
FIG. 13.

EPR spectra for the detection of reactive oxygen species.

FIG. 13.

EPR spectra for the detection of reactive oxygen species.

Close modal

The photodegradation of organic dye primarily arises from the presence of reactive oxygen species, particularly hydroxyl radicals. Using coumarin as a probe molecule and the PL (photoluminescence) method, it is possible to measure how many of these hydroxyl radicals are produced. This reaction produces 7-hydroxycoumarin, which exhibits a strong fluorescence at 452 nm wavelength, by combining coumarin and hydroxyl radicals. Initially, 50 mg of the photocatalyst was dissolved in a 100-ml solution of 1 mM coumarin to start the reaction. An equilibrium adsorption–desorption process was established by letting the mixture rest for 30 min. After that, the solution was exposed to visible light. 2 ml aliquots of the reaction mixture were taken out and examined with a photoluminescent spectrophotometer at intervals of 10 min. Figure 14 shows how the concentration of hydroxyl radicals goes along with the strength of the peak seen in the spectrum. This experimental process underscores the critical role played by reactive oxygen species, particularly hydroxyl radicals, in the degradation of methylene blue organic dye. Figure 15 shows the EPR spectra for the confirmation of hydroxyl radicals.

FIG. 14.

OH radical trapping PL spectra of AgVO3 nanoparticles.

FIG. 14.

OH radical trapping PL spectra of AgVO3 nanoparticles.

Close modal
FIG. 15.

EPR spectra for the detection of hydroxyl free radicals.

FIG. 15.

EPR spectra for the detection of hydroxyl free radicals.

Close modal

The stability of a photocatalyst primarily hinges on its recycling and reproducibility, maintaining consistent efficiency.44 In our study, we conducted a series of repeated experiments comprising five cycles, employing 50 mg of AgVO3 under UV light. The nano-catalyst was thoroughly washed and dried at 110 °C for 2 h following each cycle. Remarkably, the efficiency of photodegradation remained nearly constant throughout these cycles. This underscores the exceptional stability and reproducibility of AgVO3 nanoparticles in the context of photochemical reactions, as demonstrated in Fig. 16.

FIG. 16.

(a) Degradation of methylene blue dye after irradiation under visible light for five successive photocatalytic cycles. (b) XRD plot of a catalyst after the photodegradation process.

FIG. 16.

(a) Degradation of methylene blue dye after irradiation under visible light for five successive photocatalytic cycles. (b) XRD plot of a catalyst after the photodegradation process.

Close modal

In this research, we examined the potential of AgVO3 nanoparticles as effective antimicrobial agents. Utilizing three different doses of AgVO3 nanoparticles, our antimicrobial screening examined their efficiency against both gram-positive and gram-negative bacterial species. AgVO3 nanoparticles at a concentration of 25 g/well in a solution have demonstrated strong antibacterial activity at all dosages tested. It established distinct 20- and 25-mm-wide inhibitory zones for the gram-positive micro-organisms S. aureus and B. subtilis. Significant zones of inhibition were also discovered, with the K. pneumonia zone measuring 22 mm and the P. aeruginosa zone measuring 26 mm (Table II). These nanoparticles work because of their small size and large surface area, which promote electrostatic interaction between the negatively charged bacterial strains and the positively charged nanoparticles. The electrostatic attraction between the two opposing charges is primarily responsible for the antibacterial activity. In addition, as indicated in Table III, our research found that all of the compounds under consideration were effective at preventing spore germination in the fungus under consideration. In this case, the results of our tests for the minimum bactericidal concentration (MBC), minimum fungicidal concentration (MFC), and minimum inhibitory concentration (MIC) are shown in Table IV.

TABLE II.

In vitro antibacterial activity of AgVO3 nanoparticles.

Diameter of zone of inhibition (mm)
Gram-positive bacteriaGram-negative bacteria
S. aureusB. subtilisP. aeruginosaK. pneumoniae
Compound12.5 μg/well25 μg/well50 μg/well12.5 μg/well25 μg/well50 μg/well12.5 μg/well25 μg/well50 μg/well12.5 μg/well25 μg/well50 μg/well
AgVO3 18 20 19 22 25 21 19 26 25 19 22 20 
Chloramphenicol 20 23 24 23 29 22 25 32 28 22 28 22 
Control (DMSO)  
Diameter of zone of inhibition (mm)
Gram-positive bacteriaGram-negative bacteria
S. aureusB. subtilisP. aeruginosaK. pneumoniae
Compound12.5 μg/well25 μg/well50 μg/well12.5 μg/well25 μg/well50 μg/well12.5 μg/well25 μg/well50 μg/well12.5 μg/well25 μg/well50 μg/well
AgVO3 18 20 19 22 25 21 19 26 25 19 22 20 
Chloramphenicol 20 23 24 23 29 22 25 32 28 22 28 22 
Control (DMSO)  

(-) No activity. (±) Standard deviation.

TABLE III.

In vitro antifungal activity of AgVO3 nanoparticles.

Diameter of zone of inhibition (mm)
A. nigerP. chrysogenum
Compound12.5 μg/well25 μg/well50 μg/well12.5 μg/well25 μg/well50 μg/well
AgVO3 16 15 17 14 20 19 
Ketoconazole 18 20 21 17 21 23 
Control (DMSO) 
Diameter of zone of inhibition (mm)
A. nigerP. chrysogenum
Compound12.5 μg/well25 μg/well50 μg/well12.5 μg/well25 μg/well50 μg/well
AgVO3 16 15 17 14 20 19 
Ketoconazole 18 20 21 17 21 23 
Control (DMSO) 

(-) No activity.

TABLE IV.

MIC, MBC, and MFC of AgVO3 nanoparticles.

CompoundMinimum inhibitory concentration, MIC (MBC/MFC) μg/well
S. aureusB. subtilisP. aeruginosaK. pneumoniaeA. nigerP.chrysogenum
AgVO3 25 (100) 6.25 (200) 12.5 (50) 25 (100) 50 (200) 12.5 (50) 
Chloramphenicol 6.25 6.25 6.25 12.5 
Ketoconazole 6.25 6.25 
CompoundMinimum inhibitory concentration, MIC (MBC/MFC) μg/well
S. aureusB. subtilisP. aeruginosaK. pneumoniaeA. nigerP.chrysogenum
AgVO3 25 (100) 6.25 (200) 12.5 (50) 25 (100) 50 (200) 12.5 (50) 
Chloramphenicol 6.25 6.25 6.25 12.5 
Ketoconazole 6.25 6.25 

(-) No activity.

In this study, we employed jackfruit extract as a green fuel and as a reducing agent to produce AgVO3 nanoparticles by a simple, inexpensive, easy, and environmentally friendly precipitation procedure. AgVO3 nanoparticles were further evaluated for environmentally friendly uses, such as photocatalytic reactions with methylene blue organic dye, by varying the concentration of the catalyst, pH, dye, scavenging activity, and OH radical detection. Based on the findings, we can say that the synthesized AgVO3 can effectively remove organic waste from water, works well as a photocatalyst to remove pollutants, and degrades 98% of the organic waste in 180 min. Ultimately, the synthesized nanoparticles improved the photocatalytic activity against the organic dye methylene blue, which is verified by their bandgap of 2.54 eV. This work shows how easily and affordably greener extracts may be used as fuel for the synthesis of AgVO3 photocatalyst. Another application of AgVO3 is that it acts as a good antimicrobial agent.

The authors extend their appreciation to BET®, JSS College, Ooty Road, Mysuru, and RIE, Bhubaneswar, for providing the necessary facilities to carry out research work.

The authors would also like to acknowledge the support provided by Researchers Supporting (Project No. RSP2024R297), King Saud University, Riyadh, Saudi Arabia.

The authors have no conflicts to disclose.

R. Prakruthi: Conceptualization (equal); Data curation (equal). H. N. Deepakumari: Conceptualization (equal); Data curation (equal). H. D. Revanasiddappa: Conceptualization (equal); Formal analysis (equal). Faisal M. Alfaisal: Conceptualization (equal); Formal analysis (equal). Shamshad Alam: Conceptualization (equal); Data curation (equal). Hasan Sh. Majdi: Data curation (equal); Formal analysis (equal). Mohammad Amir khan: Conceptualization (equal); Data curation (equal). Shareefraza J. Ukkund: Conceptualization (equal); Data curation (equal).

All the data studied during this work are included in this article.

1.
M.
Ismail
,
K.
Akhtar
,
M. I.
Khan
,
T.
Kamal
,
M. A.
Khan
,
A. M.
Asiri
,
J.
Seo
, and
S. B.
Khan
, “
Pollution, toxicity and carcinogenicity of organic dyes and their catalytic bio-remediation
,”
Curr. Pharm. Des.
25
,
3645
3663
(
2019
).
2.
I. S.
Yunus
,
Harwin
,
A.
Kurniawan
,
D.
Adityawarman
, and
A.
Indarto
, “
Nanotechnologies in water and air pollution treatment
,”
Environ. Technol. Rev.
1
,
136
148
(
2012
).
3.
M. B.
Lande
,
V. R.
Phatake
,
P. B.
Chavha
,
M. P.
Choudhari
,
R.
Chauha
, and
H. J.
Solunke
, “
Role of nanotechnology to control water pollution
,”
Int. J. Res. Biosci., Agric., Technol.
15
,
113
116
(
2020
).
4.
S. K.
Srikar
,
D. D.
Giri
,
D. B.
Pal
,
P. K.
Mishra
, and
S. N.
Upadhyay
, “
Green synthesis of silver nanoparticles: A review
,”
Green Sustainable Chem.
6
,
34
56
(
2016
).
5.
S. P.
Goutam
,
G.
Saxena
,
D.
Roy
,
A.
Yadav
, and
R. N.
Bharagava
, “
Green synthesis of nanoparticles and their applications in water and wastewater treatment
,”
Bioremed. Ind. Waste Environ. Saf.
27
,
349
379
(
2019
).
6.
C.
Mallikarjunaswamy
,
P.
Parameswara
,
S.
Pramila
,
G.
Nagaraju
,
H. N.
Deepakumari
, and
V.
Lakshmi Ranganatha
,
J. Mater. Sci.: Mater. Electron.
33
,
20361
20372
(
2022
).
7.
H. N.
Deepakumari
,
V.
Lakshmi Ranganatha
,
G.
Nagaraju
,
R.
Prakruthi
, and
C.
Mallikarjunaswamy
, “
Facile green synthesis of zirconium phosphate nanoparticles using Aegle marmelos: Antimicrobial and photodegradation studies
,”
Mater. Today Proc.
62
,
5169
5173
(
2022
).
8.
T. I.
Shaheen
,
A.
Fouda
, and
S. S.
Salem
, “
Integration of cotton fabrics with biosynthesized CuO nanoparticles for bactericidal activity in the terms of their cytotoxicity assessment
,”
Ind. Eng. Chem. Res.
60
,
1553
1563
(
2021
).
9.
M.
Lu
,
Y.
Cui
,
S.
Zhao
, and
A.
Fakhri
, “
Cr2O3 cellulose hybrid nanocomposites with unique properties: Facile synthesis, photocatalytic, bactericidal and antioxidant application
,”
J. Photochem. Photobiol., B
205
,
111842
(
2020
).
10.
A. M.
Allahverdiyev
,
E. S.
Abamor
,
M.
Bagirova
, and
M.
Rafailovich
, “
Antimicrobial effects of TiO2 and Ag2O nanoparticles against drug-resistant bacteria and leishmania parasites
,”
Future Microbiol.
6
,
933
940
(
2011
).
11.
A. A.
Fayyadh
and
M. H.
Jaduaa Alzubaidy
, “
Green-synthesis of Ag2O nanoparticles for antimicrobial assays
,”
J. Mech. Behav. Mater.
30
,
228
236
(
2021
).
12.
C.
Mallikarjunaswamy
,
S.
Pramila
,
G. S.
Shivaganga
,
H. N.
Deepakumari
,
R.
Prakruthi
,
G.
Nagaraju
,
P.
Parameswara
, and
V.
Lakshmi Ranganatha
, “
Facile synthesis of multifunctional bismuth oxychloride nanoparticles for photocatalysis and antimicrobial test
,”
Mater. Sci. Eng. B
290
,
116323
(
2023
).
13.
V. K. M.
Katta
and
R.
Dubey
, “
Green synthesis of silver nanoparticles using Tagetes erecta plant and investigation of their structural, optical, chemical and morphological properties
,”
Mater. Today Proc.
45
,
794
798
(
2021
).
14.
A.
Biswas
,
L.
Chawngthu
,
C.
Vanlalveni
,
R.
Hnamte
,
R.
Lalfakzuala
, and
L.
Rokhum
, “
Biosynthesis of silver nanoparticles using Selaginella bryopteris plant extracts and studies of their antimicrobial and photocatalytic activities
,”
J. Bionanosci.
12
,
227
232
(
2018
).
15.
P.
Attri
,
S.
Garg
,
J. K.
Ratan
, and
A. S.
Giri
, “
Silver nanoparticles from Tabernoemontana divaricate leaf extract: Mechanism of action and bio-application for photodegradation of 4-aminopyridine
,”
Environ. Sci. Pollut. Res.
30
(
10
),
24856
24875
(
2023
).
16.
S. E.
Laouini
,
A.
Bouafia
,
A. V.
Soldatov
,
H.
Algarni
,
M.
Tedjani
,
G. A. M.
Ali
, and
A.
Barhoum
, “
A. Barhou green synthesized of Ag/Ag2O nanoparticles using aqueous leaves extracts of Phoenix dactylifera L. and their azo-dye photodegradation
,”
Membranes
11
,
468
(
2021
).
17.
A. C.
Ekennia
,
D.
Uduagwu
,
N. N.
Nwaji
,
C. C.
Nwosa
,
O. J.
Olowu
,
O. L.
Nwanji
,
D. A.
Udu
,
S. U.
Christopher
,
T. A.
Andrew
,
A. N.
Nkwor
, and
J. E.
Inya
, “
Facile green synthesis and biological evaluation of biogenic silver nanoparticles using aqueous extract of Alchornea laxiflora leaf
,”
Inorg. Nano-Met. Chem.
52
,
979
990
(
2022
).
18.
M. A.
Awad
,
A. A.
Hendi
,
K. M.
Ortashi
,
B.
Alzahrani
,
D.
Soliman
et al, “
Biogenic synthesis of silver nanoparticles using Trigonella foenum-graecum seed extract: Characterization, photocatalytic and antibacterial activities
,”
Sens. Actuators, A
323
,
112670
(
2021
).
19.
D. A.
Lomeli-Rosales
,
A.
Zamudio-Ojeda
,
O. K.
Reyes-Maldonado
,
M. E.
López-Reyes
,
G. C.
Basulto-Padilla
,
E. J.
Lopez-Naranjo
,
V. M.
Zuñiga-Mayo
, and
G.
Velázquez-Juárez
, “
Green synthesis of gold and silver nanoparticles using leaf extract of Capsicum chinense plant
,”
Molecules
27
,
1692
(
2022
).
20.
M.
Oves
,
M.
Ahmar Rauf
,
M.
Aslam
,
H. A.
Qari
,
H.
Sonbol
,
I.
Ahmad
,
G.
Sarwar Zaman
, and
M.
Saeed
, “
Green synthesis of silver nanoparticles by Conocarpus Lancifolius plant extract and their antimicrobial and anticancer activities
,”
Saudi J. Biol. Sci.
29
,
460
471
(
2022
).
21.
B.
Ajitha
,
Y.
Ashok Kumar Reddy
, and
P.
Sreedhara Reddy
, “
Green synthesis and characterization of silver nanoparticles using Lantana camara leaf extract
,”
Mater. Sci. Eng. C
49
,
373
381
(
2015
).
22.
R.
Jain
,
S.
Mendiratta
,
L.
Kumar
, and
A.
Srivastava
, “
Green synthesis of iron nanoparticles using Artocarpus heterophyllus peel extract and their application as a heterogeneous Fenton-like catalyst for the degradation of Fuchsin Basic dye
,”
Curr. Res. Green Sustainable Chem.
4
,
100086
(
2021
).
23.
S.
Krishnan Sundarrajan
and
L.
Pottail
, “
Green synthesis of bimetallic Ag@Au nanoparticles with aqueous fruit latex extract of Artocarpus heterophyllus and their synergistic medicinal efficacies
,”
Appl. Nanosci.
11
,
971
981
(
2021
).
24.
U. B.
Jagtap
and
V. A.
Bapat
, “
Green synthesis of silver nanoparticles using Artocarpus heterophyllus Lam. Seed extract and its antibacterial activity
,”
Ind. Crops Prod.
46
,
132
137
(
2013
).
25.
C.
Vidya
,
M. C.
Prabha
, and
M. A.
Raj
, “
Green mediated synthesis of zinc oxide nanoparticles for the photocatalytic degradation of Rose Bengal dye
,”
Environ. Nanotechnol., Monit. Manage.
6
,
134
138
(
2016
).
26.
N.
Motakef-Kazemi
and
M.
Yaqoubi
, “
Green synthesis and characterization of bismuth oxide nanoparticle using Mentha Pulegium extract
,”
Iran. J. Pharm. Res.
19
,
70
79
(
2020
).
27.
C.
Belver
,
C.
Adan
,
S.
García-Rodríguez
, and
M.
Fernández-García
, “
Photocatalytic behavior of silver vanadates: Microemulsion synthesis and post-reaction characterization
,”
Chem. Eng. J.
224
,
24
31
(
2013
).
28.
Z.
Yan
,
S.
Zhang
,
M.
Gao
,
Z.
Zhu
,
J.
Li
,
Z.
Huang
,
T.
Pan
,
T.
Hu
,
Q.
Li
,
X.
Wang
, and
Y.
Lin
, “
Ultrasensitive and wide-range-detectable flexible breath sensor based on silver vanadate nanowires
,”
ACS Appl. Electron. Mater.
5
,
520
525
(
2023
).
29.
K.
Hua
,
X.
Li
,
R.
Bao
,
D.
Fang
,
M.
Jiang
,
J.
Yi
,
Z.
Luo
,
Y.
Shu
, and
B.
Sun
, “
Electrochemical performance of silver vanadate/silver nanowire composite for lithium-ion batteries
,”
Solid State Ionics
325
,
133
140
(
2018
).
30.
Y.
Pooja
,
P.
Chaitanya
,
C.
Vinay
,
K. S.
Uloopi
,
P.
Ahalya
, and
R. K.
Alla
, “
Antimicrobial and mechanical properties of GIC incorporated with silver vanadate nanoparticles: An in-vitro study
,”
J. Clin. Diagn. Res.
17
,
29
(
2023
).
31.
A. M.
Abdelghany
,
M. Y.
Abdelaal
,
D.
Hanna
, and
A.
Aboelkheir
, “
Investigation of chitosan embedded with silver vanadate nanorods and their optical properties
,”
J. Text. Color. Polym.
15
,
37
45
(
2018
).
32.
J.
Zhang
,
J.
Wang
,
H.
Xu
,
X.
Lv
,
Y. X.
Zeng
,
J.
Duan
, and
B.
Hou
, “
The effective photocatalysis and antibacterial properties of AgBr/AgVO3 composites under visible-light
,”
RSC Adv.
9
,
37109
37118
(
2019
).
33.
S. B.
Somvanshi
,
S. A.
Jadhav
,
M. V.
Khedkar
,
P. B.
Kharat
,
S. D.
More
, and
K. M.
Jadhav
, “
Structural, thermal, spectral, optical and surface analysis of rare earth metal ion (Gd3+) doped mixed Zn–Mg nano-spinel ferrites
,”
Ceram. Int.
46
,
13170
13179
(
2020
).
34.
S.
Liang
,
J.
Zhou
,
A.
Pan
,
Y.
Li
,
T.
Chen
,
Z.
Tian
, and
H.
Ding
, “
Facile, synthesis of beta-AgVO3 nanorods as cathode for primary lithium batteries
,”
Mater. Lett.
74
,
176
179
(
2012
).
35.
R. D.
Holtz
,
B. A.
Lima
,
A. G.
Souza Filho
,
M.
Brocchi
, and
O. L.
Alves
, “
Nanostructured silver vanadate as a promising antibacterial additive to water-based paints
,”
Nanomed. Nanotechnol. Biol. Med.
8
,
935
940
(
2012
).
36.
S. J.
Bao
,
Q. L.
Bao
,
C. M.
Li
,
T. P.
Chen
,
C. Q.
Sun
,
Z. L.
Dong
,
Y.
Gan
, and
J.
Zhang
, “
Synthesis and electrical transport of novel channel-structured beta-AgVO3
,”
Small
3
,
1174
1177
(
2007
).
37.
A.
Zyoud
,
A.
Zu
,
M. H. S.
Helal
,
D.
Park
,
G.
Campet
, and
H. S.
Hilal
, “
Optimizing photo-mineralization of aqueous methyl orange by nano-ZnO catalyst under simulated natural conditions
,”
J. Environ. Health Sci. Eng.
13
,
1
10
(
2015
).
38.
Y.
Lu
,
Y.
Wan
,
S.
Bi
,
H.
Weng
,
Y.
Huang
,
L.
Qin
, and
H. J.
Seo
, “
Synthesis, surface and optical properties of Ag2CaV4O12 nanoparticles for dye removal under visible irradiation
,”
Mater. Chem. Phys.
180
,
263
271
(
2016
).
39.
A. P.
de Melo Monteiro
,
R.
Dias Holtz
,
L.
Carneiro Fonseca
,
C. H.
Zanini Martins
,
M.
de Sousa
,
L. A. V.
de Luna
,
D. L.
de Sousa Maia
, and
O. L.
Alves
, “
Nano silver vanadate AgVO3: Synthesis, new functionalities and applications
,”
Chem. Rec.
18
,
973
985
(
2018
).
40.
R.
Konta
,
H.
Kato
,
H.
Kobayashi
, and
A.
Kudo
, “
Photophysical properties and photocatalytic activities under visible light irradiation of silver vanadates
,”
Phys. Chem. Chem. Phys.
5
,
3061
3065
(
2003
).
41.
S.
Sharma
,
K.
Kumar
,
N.
Thakur
,
S.
Chauhan
, and
M. S.
Chauhan
, “
The effect of shape and size of ZnO nanoparticles on their antimicrobial and photocatalytic activities: A green approach
,”
Bull. Mater. Sci.
43
,
20
(
2020
).
42.
M.
Ganesh babu
,
N.
Kannan
,
S.
Sundar Venkatesh
,
G.
Paulraj
,
K.
Jeganathanand
, and
D. M.
Mubarkali
, “
Synthesis and characterization of BiVO4 nanoparticles for environmental applications
,”
RSC Adv.
10
,
18317
18322
(
2020
).
43.
J.
Liu
,
J.
Yuan
,
X.
Hou
,
X.
Luo
,
H.
Hu
,
Z.
Yang
,
Q.
Sun
,
S.
Chen
, and
D.
Xu
, “
Enhanced photodegradation of organic contaminants on Ag3VO4/covalent organic frameworks under UV light: Mechanism analysis
,”
J. Environ. Chem. Eng.
11
,
111542
(
2023
).
44.
U.
Sulaeman
,
S. F.
Ramadhanti
,
H.
Diastuti
,
P.
Iswanto
,
I.
Isnaeni
, and
S.
Yin
, “
The enhanced photo-stability of defective Ag3PO4 tetrahedron prepared using tripolyphosphate
,”
Arabian J. Chem.
16
,
104409
(
2023
).