ACS Publications. Most Trusted. Most Cited. Most Read
Green Biosynthesis of Tin Oxide Nanomaterials Mediated by Agro-Waste Cotton Boll Peel Extracts for the Remediation of Environmental Pollutant Dyes
My Activity
  • Open Access
Article

Green Biosynthesis of Tin Oxide Nanomaterials Mediated by Agro-Waste Cotton Boll Peel Extracts for the Remediation of Environmental Pollutant Dyes
Click to copy article linkArticle link copied!

  • Boya Palajonnala Narasaiah
    Boya Palajonnala Narasaiah
    CASEST, School of Physics, University of Hyderabad, Prof. C. R Rao Road, Gachibowli, Hyderabad 500046, Telangana, India
    Laboratorio de Cerámicos y Nanomateriales, Facultad de Ciencias Físicas, Universidad Nacional Mayor de San Marcos, Ap. Postal 14-0149, Lima 14, Peru
  • Pravallika Banoth
    Pravallika Banoth
    CASEST, School of Physics, University of Hyderabad, Prof. C. R Rao Road, Gachibowli, Hyderabad 500046, Telangana, India
  • Arya Sohan
    Arya Sohan
    CASEST, School of Physics, University of Hyderabad, Prof. C. R Rao Road, Gachibowli, Hyderabad 500046, Telangana, India
    More by Arya Sohan
  • Badal Kumar Mandal
    Badal Kumar Mandal
    Department of Chemistry, School of Advanced Sciences, Vellore Institute of Technology, Vellore 632014, Tamil Nadu, India
  • Angel G. Bustamante Dominguez
    Angel G. Bustamante Dominguez
    Laboratorio de Cerámicos y Nanomateriales, Facultad de Ciencias Físicas, Universidad Nacional Mayor de San Marcos, Ap. Postal 14-0149, Lima 14, Peru
  • Luis De Los Santos Valladares*
    Luis De Los Santos Valladares
    Laboratorio de Cerámicos y Nanomateriales, Facultad de Ciencias Físicas, Universidad Nacional Mayor de San Marcos, Ap. Postal 14-0149, Lima 14, Peru
    Cavendish Laboratory, Department of Physics, University of Cambridge, J.J. Thomson Avenue, Cambridge CB3 OHE, U.K.
    School of Materials Science and Engineering, Northeastern University, No 11, Lane 3, Wenhua Road, Heping District, Shenyang 110819, Liaoning, People’s Republic of China
    *Email: [email protected]
  • Pratap Kollu*
    Pratap Kollu
    CASEST, School of Physics, University of Hyderabad, Prof. C. R Rao Road, Gachibowli, Hyderabad 500046, Telangana, India
    *Email: [email protected]
    More by Pratap Kollu
Open PDF

ACS Omega

Cite this: ACS Omega 2022, 7, 18, 15423–15438
Click to copy citationCitation copied!
https://doi.org/10.1021/acsomega.1c07099
Published April 26, 2022

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Abstract

Click to copy section linkSection link copied!

The sustainable synthesis of metal oxide materials provides an ecofriendly and more exciting approach in the domain of a clean environment. Besides, plant extracts to synthesize nanoparticles have been considered one of the more superior ecofriendly methods. This paper describes the biosynthetic preparation route of three different sizes of tetragonal structure SnO2 nanoparticles (SNPs) from the agro-waste cotton boll peel aqueous extract at 200, 500, and 800 °C for 3 h and represents a low-cost and alternative preparation method. The samples were characterized by X-ray diffraction, Fourier transform infrared spectrophotometry, ultraviolet–visible absorption spectroscopy, high-resolution transmission electron microscopy (HR-TEM), and energy-dispersive X-ray spectroscopy. Surface area and porosity size distribution were identified by nitrogen adsorption–desorption isotherms and Brunauer–Emmett–Teller analysis. The photocatalytic properties of the SNP samples were studied against methylene blue (MB) and methyl orange (MO), and the degradation was evaluated with three different size nanomaterials of 3.97, 8.48, and 13.43 nm. Photocatalytic activities were carried out under a multilamp (125 W Hg lamps) photoreactor. The smallest size sample exhibited the highest MB degradation efficiency within 30 min than the most significant size sample, which lasted 80 min. Similarly, in the case of MO, the smallest sample showed a more superior degradation efficiency with a shorter period (40 min) than the large-size samples (100 min). Therefore, our studies suggested that the developed SNP nanomaterials could be potential, promising photocatalysts against the degradation of industrial effluents.

This publication is licensed under

CC-BY-NC-ND 4.0 .
  • cc licence
  • by licence
  • nc licence
  • nd licence
Copyright © 2022 The Authors. Published by American Chemical Society

1. Introduction

Click to copy section linkSection link copied!

Over the last years, the novel synthesis of nanomaterials such as metal, metal oxide, and metal sulfide nanoparticles have gained much attention. Metal, metal oxide, and their composites have become an exciting area in nanoscience and technology due to the many applications in versatile fields and even proven promise in human life aspects. (1) Synthesis of nanoparticles through eco-friendly routes attracted much awareness in recent years because of its environment friendly nature, that is, nontoxic to the environment, reaction process, short time, low temperature, and safe, mild reaction conditions. (2) However, modifying the surface area of the nanoparticles through the natural reducing and capping agents remains challenging. Nevertheless, nanoparticles can be generated by green chemical agents that are plant extracts that possess even higher stability. (3) The green process synthesis of metal oxide nanoparticles, particularly tin oxide nanomaterials using plant extract, is an alternative method to the chemical synthesis that could control the chemical toxicity to the environment and control the size and shape of the nanomaterials. (4) State of the art shows the possibilities of synthesis nanomaterials in an aqueous medium with reducing agents that counts the hydroxyl group of polyphenols that possess the stabilizing/capping agents without using hazardous chemicals. (5) Moreover, green synthesized nanomaterials are gaining interest in the research fields, environmental sciences, synthetic chemistry, nanobiotechnology, and material chemistry, respectively. (6) Among various nanomaterials, SnO2 nanomaterials are considered an alternative source for degradation of water purification/environmental pollutants because they possess outstanding UV light absorption and generate reactive oxygen species (ROS) that lead to the degradation of environment pollutant dye molecules. (7,8)
Tin oxide nanoparticles (SnO2 NPs) are widely used in several fields such as photocatalytic activity, energy storage application, and organic transformation used as catalysts in various fields. (9) The literature reports the synthesis of SnO2 NPs (SNPs) through multiple methods. Nejati-Moghadam et al. (2015) reported the production of SnO2 nanostructures by the precipitation method, (10) Ling Zhang et al. (2018) reported cubic sub-micron SnO2 particles for the reduction of nitrogen to sustainable process for scalable NH3 synthesis with less energy consumption, (11) and Haspulat et al. (2020) have synthesized SnO nanoparticles in the presence of Triton-X 100 (TX-100) surfactant assisted via the hydrothermal method. (12) Maharajan et al. (2020) reported a nano-Rattle SnO2@carbon composite by melt diffusion method/thermal decomposition methods for the material for high-energy Li-ion batteries. (13) Karmaoui et al. (2018) studied the band gap properties and structural study of SNPs and synthesized a one-step process by the sol gel method. (14) Zhu et al. (2019) reported ultrafine SNPs (diameter 2–4 nm) fabricated on the surfaces of g-C3N4 material through the combination of the hydrothermal method and ball milling in order to improve the dispersion of SnO2 NPs as well as to strengthen the bonding between SnO2 and g-C3N4. (15) These methods are reported to be effective for controlling the shape and size of the nanoparticles. However, all these methods require high temperature, high cost, and environmentally hazardous chemical precursors during synthesis processes. Therefore, environmentally friendly, low-cost, and straightforward production is preferred.
The potential nanomaterials synthesizing through physical and chemical methods are expensive and require high pressure, temperature, and energy. They include prolonged reactions and required toxic reactants and solvents, which are potentially hazardous to the environment. Hence, there is a continuous demand for developing new techniques for nanomaterials synthesis without the above drawbacks. Green synthesis has been adopted nowadays as the best alternative way to overcome the methods mentioned above to synthesize nanomaterials because of several advantages such as a fast and straightforward reaction process, inexpensive approach, environmentally ecofriendly to nature, reduction of metal ions to metal nano, and use of less toxic and hazardous chemicals. The synthesis of metal oxide nanomaterials through plant extracts has turned very promising because it allows both reducing and capping agents. (16) In fact, plant extracts have various phytochemical constituents, that is, secondary metabolites such as alkaloids, terpenoids, flavonoids, antioxidants, and amino acids, respectively, which exist in fruits, peels, seeds, and leaves that permit the replacement of the traditional toxic chemicals for the synthesis of nanomaterials. (17) Diallo et al. (2016) reported the preparation of n-type SNPs by a green chemistry process using Aspalathus linear natural extract as an effective chelating agent. The obtained SNPs exhibited photocatalytic degradation of methylene blue (MB), Congo red, and Eosin Y. (18) Matussin et al. (2020) also reported plant-extract-mediated SNPs. (19) Singh et al. (2019) synthesized SnO2 NPs through pomegranate leaves (Punica Granatum), and the obtained SnO2 NPs (20 nm) showed 91.5% photodegradation efficiency of MB dye under direct sunlight within 240 min. (20) Bhattacharjee et al. (2015) developed a green process of SnO2 quantum dots using the amino acids aspartic and glutamic as reducing agents and used for the photocatalyst in the degradation of Rose Bengal and Eosin Y dye under direct sunlight. (21) Sinha et al. (2017) green fabricated sphere-shaped Ag–SnO2 nanocomposites (9 nm) employing stem extracts of saccharin officinarum and used as a photocatalyst and prospective antibacterial and antioxidant agents. (22) Sudhaparimala et al. (2016) reported coupled semiconductor SnO2–ZnO NPs using the assisted gel of aloe vera plants as the medium. They studied the photocatalytic activity of the organic dye methyl orange (MO) degradation in the presence of visible light. They also studied the antibacterial activity of staphylococcus aureus and Escherichia Coli growth at the microgram level. (23)
Currently, photocatalysis applications for addressing environmental issues such as environmental pollution and energy crises have attracted more and more attention and gradually become a research hotspot. The most common pollutants identified in urban and industrial areas are volatile organic compounds such as toluene, benzaldehyde, benzyl alcohol, and chlorinated derivatives. (24) Volatile organic compounds are dangerous for human health, and they are toxic, mutagenic, and carcinogenic to all human beings. Nowadays, research has mostly been focused on the toxicity of volatile organic compounds sources or the degradation conditions rather than the byproducts generated during the treatment procedures. (25) The drawback of traditional methods such as absorption, incineration, and condensation used for removing volatile organic compounds from pollutants is that they are costly, have a short life, and lead to the production of secondary pollutants. (26) Some approaches have been devised recently to solve the problem. Photocatalysis is a promising technology for purifying wastewater and textile dyes under visible/ultraviolet light irradiation. For example, Cu–Fe2O3/Ni–ZnO nanoplate is an excellent photocatalyst for the degradation of environmental pollutants in visible light. It allows the mineralization of CO2 and water molecules. (27) The photocatalyst activity of nanoparticles mainly depends on their high surface area, size, shape, and less band gap energy. (28) Tuan et al. (2019) noticed that the photocatalyst activity of SnO2/rGO nanocomposites depends on their morphology, structure, and size. (29) Jiang et al. (2021) studied the anode materials for the degradation of organic pollutants by anodic oxidation. (30) Yuanyuan et al. (2018) estimated the degradation of methyl blue (MB) and Rhodamine B (RhB) by the SnO2 nanoparticle photocatalyst in more than 90% for under UV light irradiation within 50 min and 270 min. (31) Huang et al. (2018) prepared a tetragonal phase Bi2O2CO3 catalyst and reported its photocatalytic behavior with the degradation of MO dye under visible light conditions. (32) They also prepared Cerium-based hybrid nanorods (CN-600 to CN-800) photocatalyst; however, the CN-700 catalyst displays a degradation efficiency of 100% for MB in aqueous solution in 90 min. (33)
The present work reports the green production of photocatalyst of tetragonal structure SNPs of different sizes (2–13 nm) from eco-friendly nature, cost effective manner using Agro-waste cotton boll peels with tin Chloride di-hydrate as a precursor. The prepared tetragonal structure SnO2 photocatalyst nanoparticles were characterized by X-ray diffraction (XRD) technique, high-resolution transmission electron microscopy (HR-TEM), UV–visible diffuse reflectance spectroscopy (UV–vis DRS), and surface area analyzed by Brunauer–Emmett–Teller (BET). The tetragonal structure SnO2 photocatalyst offered an enhanced catalytic activity to degrade MB and MO dyes under UV light irradiation.

2. Experimental Procedure

Click to copy section linkSection link copied!

2.1. Material and Methods

Tin chloride di-hydrate (SnCl2·2H2O), MB, and MO were purchased from Sigma Aldrich to synthesize the SNP samples. All other chemicals and reagents used analytical grade with no purification, and double-distilled water was used as a solvent. Agro-waste cotton peels were collected near the Tungabhadra River, Mantralayam Road, Kurnool District, Andhra Pradesh, India.

2.2. Preparation of Agro-Waste Cotton Peel Extracts

Freshly collected peels were washed 2–3 times under running tap water, followed by sanitizing with Milli-Q water one time and then dried at room temperature under dust-free conditions for 1 week. The preparation followed a similar way reported by Balaji Reddy et al. (2017) to synthesize metal oxide nanoparticles using Eucalyptus globulus leaf. (34) The dried peels were ground into powder through an electric mixer and sieved by a 100 mesh sieving net. About 3 g of powder was added to 100 mL of de-ionized water and kept for boiling on a water bath at 80 °C for 30 min to prepare the aqueous extract. Eventually, the extract was cooled at room temperature, filtered through a Whatman no.1 filter paper, and stored at 4 °C in a refrigerator.

2.3. Synthesis of Tin Oxide (SnO2 NPs) from Agro-Waste Cotton Peel Extract

About 0.329 g of Tin chloride di-hydrate (SnCl2·2H2O) (0.05 M) was dissolved in 30 mL of double-distilled water and subsequently 30 mL of agro-waste cotton peel aqueous extract was added drop-wise while thoroughly stirring at 80 °C for 3 h. The color of the solution changed from light yellow to brown color after 30 min because of the heating process. Then, the mixture was cooled at room temperature and subsequently centrifuged for 30 min at 6000 rpm. The residue was carefully collected and washed three times with absolute ethanol, following washing with double-distilled water and eventually dried on the hot plate at 60 °C. The resulting dried powder followed calcination at 200 °C for 3 h inside a muffle furnace. That sample was labeled as tin oxide (SnO2) nanoparticles (SNPs-2). A similar procedure was used to prepare additional samples, once synthesized and calcined at 500 °C, labeled as SNPs-5, and another at 800 °C (labeled as sample SNPs-8) for 3 h inside a muffle furnace.

2.4. Characterization of the SnO2 Nanoparticles

A X-ray powder diffractometer, Bruker D8, was used to measure the SNPs samples by X-ray diffraction. The data were recorded in the range of 2θ values from 10 to 90° with a scanning rate of 4° min–1 by using the Cu-Kα radiation with a λmax of 1.54 Å. The morphology and size of the synthesized SNPs were inspected by dispersing the SNP samples onto Cu grids through HR-TEM (Model: JEM 2100 USA JEOL) with 0.1 nm resolution and an acceleration voltage of 200 keV. The TEM equipment also provided information about the elemental composition of the samples by energy-dispersive X-ray (EDX) analysis. The optical properties of the synthesized SNP samples were also analyzed by using a Shimadzu (model UV-2450) UV–visible spectrophotometer at room temperature. The bang gap calculations assumed the absorption band. The functional groups in the cotton peel agro-waste aqueous extract SNP samples were analyzed using FT-IR (Fourier transform infrared) spectroscopy. Functional groups determined in purified dried SNP samples using a Shimadzu IR AFFINITY-1 against the agro-waste cotton peel aqueous extract. However, pelted took control under similar instrumental conditions using JASCO FT-IR in the wavenumber range 500–4000 cm–1 at a resolution of 4 cm–1 using potassium bromide pellets in the diffuse reflectance mode. Quanta chrome nova-1000 autosorb-1 instrument was used to determine the surface area and pore diameter through nitrogen adsorption–desorption isotherms.

2.5. Photocatalytic Activity of SNP Samples

The SNP samples were used as a photocatalyst to degrade the MB and MO dyes under UV light irradiation. The present investigation was performed with a Heber photoreactor having multilamp UV lamps (254 nm) with a low pressure of 125 W under continuous stirring. The experiment was carried out by loading 0.02 g of SNP samples with 60 mL of MB dye solution (concentration 10 mg/L) into a quartz tube (100 mL capacity) and then followed by continuous stirring, without disturbance, under the dark condition for 50 min without UV-light exposure to obtain the constant equilibrium. After getting the stable equilibrium, the final solution was exposed with UV light radiations at 125 W, 254 nm, and regular time intervals. The sample was collected in an Eppendorf tube (2 mL capacity) and centrifuged. The degraded de-colorization solution was monitored by UV–vis spectro-photometer analysis to record the de-colorization solution’s absorbance, scanned at 200–800 nm. The catalyst was then recollected after complete de-colorization of MB dye molecules through centrifugation. We also recycled it to check the stability and reusability of the SNP catalyst. The same procedure was repeated for the MO dye degradation and all other samples (SNPs-5 and SNPs-8) to degrade MB and MO dyes. We also studied the effects of varying catalysts and dye concentration for MB and MO degradation of dyes. Thus, the degradation of the samples was recorded by UV–vis spectroscopy.

3. Results and Discussion

Click to copy section linkSection link copied!

3.1. Determination of Crystalline Nature from XRD-Analysis of SNP Samples

XRD patterns of the samples SNPs-2, SNPs-5, and SNPs-8 are shown in Figure 1 and analyzed to estimate the crystal structure and crystallinity of the nanomaterials. The diffraction pattern of sample SNPs-2 represents reflections at 2θ = 26.53, 34.38, 52.12, and 65.27° which correspond to the diffraction peaks hkl values of lattice planes (110), (101), (211), and (220), respectively (see Figure 1A). The confirmed diffraction pattern corresponds to the tetragonal structure of SnO2 NPs at 200 °C, which matches the standard Match3 software (JCPDS card no: 96-210-4744). Because the XRD pattern presents monophasic diffraction, that confirms the purity of the SNP-2 nanomaterials. Figure 1B presents the XRD pattern of the SNP-5 sample with 2θ reflections at 26.96, 34.52, 38.22, 52.16, 65.68, 71.53, and 78.72° which attributes to the Miller indices reflecting planes (110), (101), (200), (211), (301), (112), and (321), respectively, and with the tetragonal structure after annealing at 500 °C (JCPDS card no: 96-210-4744). However, there are no impurity peaks in the RDX pattern, which indicates the purity of SNP-5.

Figure 1

Figure 1. XRD pattern of sample SNPs: SNPs-200 °C (A), SNPs-500 °C (B), and SNPs-800 °C (C).

The XRD pattern for sample SNPs-8, shown in Figure 1C, presents 2θ reflections at 26.82, 34.29, 38.26, 52.27, 54.77, 62.11, 65.78, 71.49, 78.86, and 83.69° which are characteristic diffraction peaks of hkl planes (110), (101), (200), (211), (220), (311), (301), (112), (321), and (222), respectively. Therefore, the noticed XRD patterns confirm the tetragonal structure of the SNPs-8 sample and match the standard Match3 software (JCPDS card no: 96-210-4755). The results concluded that all the SNP samples are tetragonal structures in pure phase so that the activity of SNP samples depends on its crystallinity nature.
Table 1 summarizes the results given by XRD: the full width at half-maximum (fwhm), inter-planar distance (d) values, and crystallite sizes. The crystallite sizes were calculated to be 4.7, 9.7, and 14 nm for the SNPs 2, 5, and 8, respectively, meaning that the annealing process plays a significant role in the crystallization and changes the fwhm values. The increase of temperature influences in the diffraction intensity and reduces the fwhm values, attributed to the systematic arrangement of atoms on the crystal system. Therefore, increasing the annealing temperature gradually increases the particle size because of oxygen vacancies, originated by atomic diffusion, lattice strains, and crystal lattice defects, which causes the crystallite size to increase. The SNP samples crystallite size could be calculated according to the following Scherrer formula equation I
(I)
Here, D is the crystallite size, k is the Scherrer’s coefficient (0.891), λ is the X-ray wavelength, θ is the Bragg angle, and β is the fwhm intensity (in radians). Note that the crystallite size is obtained or defined as the measurement of the diffracting crystallite size of the samples that are not equal to the particle size because they exhibit polycrystalline aggregates. Estimation of the crystallite sizes and interplanar distance (d) values of the SNP samples planes fitted with Scherrer’s equation are listed in Table 1. Therefore, calculated average crystallite sizes of sample SNPs-2 at 200 °C, sample SNPs-5 at 500 °C, and sample SNPs-8 at 800 °C are 4, 9, and 14 nm, respectively. Similar results have been obtained by Osuntokun et al. (2017) using the leaf extract of Brassica oleracea L. var. botrytis which results in SNPs with crystallite sizes 3.6–6.3 nm. (35)
Table 1. Structure and Geometric Parameters of SNP Samples of Diffraction Study
SNPslattice planefwhm valued-spacing (Å)Cos(θ)crystallite size (nm)
SnO2 NPs at 200 °C(110)26.532.0443.460.97334.2
 (101)34.382.5572.630.95533.4
 (211)52.122.5871.840.89833.6
 (220)65.271.2731.450.84217.8
average size4.7
SnO2 NPs at 500 °C(110)26.961.7543.370.97244.9
 (101)34.520.7172.680.954912.1
 (200)38.220.7482.430.944811.8
 (211)52.160.9721.860.89819.5
 (220)65.680.9701.410.840110.2
average size9.7
SnO2 NPs at 800 °C(110)26.820.8573.350.972710
 (101)34.290.6702.670.955513
 (200)38.260.4932.480.944717.8
 (211)52.270.9361.750.89779.9
 (220)54.770.6471.790.887914.5
 (311)62.110.6061.540.856616
 (301)65.780.5261.470.839718.8
 (321)78.860.8471.260.772412.7
average size14

3.2. Size and Morphology Identified by HR-TEM Analysis

The size, morphology, d-spacing, elemental composition, and crystalline/amorphous nature of the SnO2 samples were inspected by HR-TEM as shown in Figures 24. The HR-TEM images of sample SNPs-2 at different magnifications, selected area electron diffraction pattern (SAED), particle size distribution, and EDX spectra are presented in Figure 2A–F. Figure 2A,B displays the micrographs of sample SNPs-2 at different magnifications, and the size identified by the Image J software ranges 2.92–4.59 nm so that the average diameter (AD) is 3.97 nm. Moreover, the interplanar distance representing the d-spacing is 0.262 nm which is characteristic for the reflection plane (101), as shown in Figure 2C. The diffraction rings in the SAED pattern shown in Figure 2D expose the amorphous nature, corresponding to the amorphous SnO2 component of the SNPs-2 noticed from the respective XRD background above. The identified elemental composition of the SNPs-2 sample, determined by the EDX spectrum in Figure 2E, reveals weight % and atomic % of tin and oxygen: 86.59 and 13.41 weight % and 46.53 and 53.47 atomic %, respectively, and no other impurities are found. Therefore, the obtained results confirm that the SNPs-2 sample also contains amorphous SnO2 without impurity. Figure 2F presents the particle size histogram, where the particle size distribution is around 2.92–4.59 nm with an AD of 3.97 nm and the standard deviation (SD) is approximately 0.07 nm.

Figure 2

Figure 2. HR-TEM of sample SNPs-2 at annealing at 200 °C; (A,B) different scale magnifications (20 and 10 nm), (C) inter planar d-spacing, (D) SEAD pattern, (E) EDX micrograph, and (F) particle size distribution of the histogram.

Figure 3

Figure 3. HR-TEM of sample SNPs-5 at annealing at 500 °C; (A) 10 nm scale magnification, (B) 20 nm scale magnification, (C) inter planar d-spacing, (D) SEAD pattern, (E) EDX micrograph, and (F) particle size distribution of the histogram.

Figure 4

Figure 4. HR-TEM of sample SNPs-8 at annealing at 800 °C; (A) 20 nm scale magnification, (B) 50 nm scale magnification, (C) inter planar d-spacing, (D) SEAD pattern, (E) EDX micrograph, and (F) particle size distribution of the histogram.

Figure 3A–F shows the HR-TEM, EDX spectrum, SAED pattern, and particle size distribution of the SNPs-5 samples. Figure 3A,B presents TEM micrographs at different magnifications showing that the size of the particles ranges from 5.2 to 12.8 nm, and the AD is 8.4 nm. The interplane distance is around 0.27 nm, characteristic of the reflection plane (101) as represented in (Figure 3C). The circular rings existing in the SAED pattern (see Figure 3D) correspond to the poly-crystallinity nature of the SNPs-5 sample. The elemental composition of the SNPS-5 sample from the EDX spectrum (see Figure 3E) reveal tin and oxygen in atomic % and weight % are as follows: 48.76, 51.24 and 87.60, 12.40% respectively, which indicates that the polycrystalline SNPs-5 sample has no impurities. The particle size distribution is shown in Figure 3F, where particles range from 5.2 to 12.9 nm. The average particle size diameter is 8.5 nm, and the SD is around 0.4 nm.
Figure 4A–F displays the HR-TEM, SAED pattern, EDX spectrum, and particle size distribution of the SNPs-8 sample. Figure 4A,B presents different magnification TEM micrographs showing particle sizes between 15.6 and 7.7 nm with an AD of 13.4 nm. Figure 4C reveals that the inter plane distance (d) is 0.27 nm, which agrees with the respective XRD results and corresponds to the diffraction reflection plane (101). The circular radiant spots in the SAED pattern (see Figure 4D) confirms the poly-crystallinity nature of the sample. Figure 4E shows the EDX spectrum revealing the elemental composition as tin and oxygen in atomic % and weight %: 53.84, 46.16 and 89.87, 10.13%, respectively, which leads to the purity of the SNPs-8 sample. The histogram of the particle size distribution is shown in Figure 4F, where particles range around 15.6–7.7 nm, giving the average particle size diameter (AD) of 13.4 nm and SD is 0.8 nm. Similarly, Ma et al. (2020) reported SNPs obtained by the precipitation method to be in the range of 20–80 nm and an average particle size around 40 nm. (36)
The TEM results of three samples (SNPs-2, SNPs-5, and SNPs-8) demonstrate the strong influence of the temperature on the size of the SnO2 nanomaterials. Increasing the annealing temperature may lead to defects in the crystal system and subsequently regulate the arrangement of atoms in the unit cell resulting in the rise of the size of the nanoparticles and the crystallinity, as confirmed by XRD above. From the results, we concluded that the annealing process could control the size of the NPs.

3.3. Identification of Functional Groups on the Surface of the SNPs by FTIR Analysis

FTIR analysis is a well-known technique to determine the functional groups in the cotton peel aqueous extract and the capping of biomolecules on the surfaces of the SNPs based on the vibrational stretching of functional groups. Figure 5A describes the FTIR spectrum of the agro-waste cotton peel aqueous extract recorded in the range 4000–500 cm–1. In the defined spectrum, the bands located around 3346.78, 3257.63, 1737.86, 1606.64, 13346.78, 1105.97, and 921.58 cm–1 are associated with the stretching vibration of the hydroxyl (−OH) group (indicating the existence of polyphenols in the plant extract); stretching vibration of the aromatic amine (−N–H) group; stretching vibration of the carbonyl (−C═O) group; bending vibration of the amide (−N–H) group corresponding to nitrogen and hydrogen vibrations; vibration of the nitro (−N–O) group; ether group of (−C–O–C−); and bending vibration of aliphatic carbon and hydrogen (−C–H) groups, respectively. Similarly, Arumugam et al. (2021) reported various biomolecules in Syzygium cumini (Java plum) aqueous extract detected from the FTIR technique. (37)

Figure 5

Figure 5. FT-IR analysis of cotton peel aqueous extract (A), sample SNPs-2 at 200 °C, (B), sample SNPs-5 at 500 °C (C), and sample SNPs-8 at 800 °C (D).

Figure 5B displays the FTIR spectrum of the SNP sample at 200 °C. The absorption bands at 3321.25, 1639.13, and 518.44 cm–1 could be assigned to stretching vibrations of the hydroxyl (−OH) group, alkene (−C═C−) group, and oxygen (O–Sn–O), respectively. Figure 5C presents the FTIR spectrum of the SNPs at 500 °C where the bands located at 3329.98, 1627.29, and 548.87 cm–1 correspond to hydroxyl (−OH) group stretching vibrations of phenols present in the extract; stretching vibration of alkene (−C═C−) group and oxygen/metal/oxygen (O–Sn–O) stretching vibrations, respectively. The FTIR spectrum of the SNPs obtained at 800 °C is shown in Figure 5D. The band at 592.34 cm–1 corresponds to the oxygen–metal–oxygen (O–Sn–O) stretching vibration group. Both samples, SNPs-2 and SNPs-5, contain stretching and bending vibrations of biomolecules on their surface, but the SNPs-8 sample does not exhibit those bonds. This suggests that annealing at higher temperatures decomposes the bio-molecules on the surface of the SNPs. Therefore, the O–Sn–O stretching vibration bands in the SNPs annealed at 200 and 800 °C increases with the increase in the annealing temperature caused by higher atomic diffusion in the crystal system that possesses size enhancement nanomaterials. Our current results resemble recently reported results: stretching vibration of oxygen and metal, oxygen of SNPs in the range of 540–660 cm–1, which indicates O–Sn–O and Sn–O stretching vibration modes reported by Paramarta et al. (2016). (38)

3.4. UV–Visible Analysis

As mentioned in the Experimental section, in the present investigation, crystalline SnO2 nanomaterials are synthesized without using any commercial reducing or surfactant chemicals. Instead, the synthesis of the nanoparticles was performed by reducing agents that possess the cotton peel extracts that act as a capping/size controlling agent. Figure 6 represents the characteristic UV–vis absorption spectrum of the SNP samples in the band range 200–800 nm. Figure 6A–C demonstrates absorption bands at around λmax = 275, 263, and 244 nm for the SNPs-2, SNPs-5, and SNPs-8, respectively. Note that the λmax value shifts toward to the UV region when increasing the size of the particles. However, the maximum absorbance for sample SNPs-2 is in the visible region, and the maximum absorbance for sample SNPs-5 lies in both UV and visible regions.

Figure 6

Figure 6. UV–vis absorption spectra of SNP sample at 200 °C (A), SNP sample at 500 °C (B), SNP sample at 800 °C (C), band gap energy of SNPs-2 (D), band gap energy of SNPs-5 (E), and band gap energy of SNPs-8 (F).

In contrast, sample SNPs-8 has an absorption in the UV region. In this manner, the λmax absorption value depends upon the size of the nanomaterials. With the increase in size, the maximum absorption band shifted toward the UV region. The band gap/energy gap (Eg), one of the optical properties that demonstrate the semiconducting nature of nanomaterials and the absorption coefficients for the tin oxide samples, was determined by the following II.
(II)
where Eg is the energy band gap, hυ is the photon energy, and α is the optical absorption coefficient. Figure 6 also shows the energy band gaps of the SNPs-2, SNPs-5, and SNPs-8 samples, obtained by extrapolating the Tauc plot (αhυ)1/2 of the tin oxide nanomaterials versus the photon energy (hυ). Note that the band gap values depend on the sintering sample. The band gap energy values are 2.68 eV (Figure 6D), 2.92 eV (Figure 6E), and 3.14 eV (Figure 6F) for the samples annealed at 200, 500, and 800 °C, respectively. Therefore, the above observation suggests that increasing the annealing process increases the band gap of the tin oxide nanomaterials. However, it has been indicated that a decrease in the band gap value is attributed to crystal defects in the material, leading to new energy levels that decrease. (39) Thus, not only the Fermi energy level but also the temperature influences the band gap of the material. The present study reveals that the material’s band gap increased with the increase in sintering temperature, which might influence the decrease of crystal defects.

3.5. Surface Area and Porosity Identification by BET Analysis

BET technique was used to identify the pore diameter, specific surface area, pore volume, and size distribution of the SNP samples by possessing adsorption–desorption nitrogen isotherms. Figure 7A represents the adsorption–desorption of nitrogen isotherms of the SNPs-2 samples measured at 77 K. The BET surface area is around 153.20 m2/g, and the pore size distribution Barrett–Joyner–Halenda (BJH) exhibits a narrow curve of calculated pore size diameter ∼ 3.10 nm, as shown in Figure 7B. Similarly, the SNPs-5 sample exhibits a BET surface area ∼ 90.34 m2/g, with pore diameter 4.95 nm denoted in Figure 7C,D. In the case of the sample calcined at 800 °C (SNPs-8), it possesses a specific surface area ∼43.72 m2/g and pore diameter 5.41 nm, which is displayed in Figure 7E,F.

Figure 7

Figure 7. BET analysis of sample SNPs at 200 °C total surface area (A), pore diameter (B), at 500 °C total surface area (C), pore diameter (D), and at 800 °C total surface area (E), pore diameter (F).

These results reveal that the sample SNPs-2 has the highest surface area compared to samples SNPs-5 and SNPs-8. From the point of view of the size and higher surface area, this material (SNPs-2) is more promising than the others to show excellent photocatalytic activity for the remediation of organic pollutants. Our results are analogous to that one reported by Ullah et al. (2017), (40) where the SNP surface area was reported to be 47.85 m2/g for samples calcined at 600 °C and the BJH curve is narrow. However, the pore diameter was noticed to be 10.5 nm.

4. Photocatalytic Activity for the Remediation of Organic Pollutants

Click to copy section linkSection link copied!

The release of excess dyes from the industry may cause environmental effects. From this point of view, nowadays, there is more and more interest in the remediation of MB and MO dyes by semiconducting materials. Thus, SNPs were tested for the photodegradation of MB and MO dyes widely used in textiles and paper. In the present work, the photocatalytic degradation of MB and MO dyes and reaction were performed under UV light exposure.

4.1. Photocatalytic Activity of SNPs for the Degradation of MB Dye

The photocatalytic activity of cotton peel extract-mediated SNPs-3, SNPs-5, and SNPs-8 nanoparticles in remediation of MB dye was carried out in the presence of UV light (λmax = 254 nm) irradiation. Figure 8 shows the MB dye’s absorbance spectra under regular intervals recorded in the range 200–800 nm. Strong absorption bands arise at 663 nm, representing the maximum wavelength of the MB dye, and gradually decrease intensity in the absorbance spectrum concerning time in the presence of samples SNPs-2, SNPs-5, and SNPs-8. These results are similar to those reported by Huang et al. (2018) for the degradation of MB 98.7% within 120 min using the Pd/BiOI/MnOx hollow sphere catalyst. (41) The synthesized SNP samples in the present work show a remarkable decrease of MB dye degradation in the UV–vis spectra.

Figure 8

Figure 8. Photocatalytic activity of SNPs-2 for the degradation of MB dye (A), SNPs-5 for the degradation of MB dye (B), SNPs-8 for the degradation of MB dye (C), SNPs-2 percentage of MB dye degradation (D), SNPs-5 percentage of MB dye degradation (E), and SNPs-8 percentage of MB dye degradation (F).

The degradation was calculated using III.
(III)
where C0 is the MB concentration of absorbance and Ct is the degradation time taken for the MB absorbance. The results indicate that in the case of the SNPs-2 sample, the MB dye degradation represents 98.56% within 30 min, as shown in Figure 8A. In the case of sample SNPs-5, the MB degradation is 97.84% within 50 min, as shown in Figure 8B. However, in the case of sample SNPs-8, the degradation is 97.12% within 80 min in the presence of UV light, as shown in Figure 8C. The obtained degradation study is compared to other results reported in the literature in Table 2. In the beginning, the MB dye degradation was monitored without the catalyst. The results show that the degradation of MB without a catalyst for sample SNPs-2 is 9.35%, as shown in Figure 8D. In the case of sample SNPs-5, the degradation is 7.91%, as represented in Figure 8E, whereas in the case of sample SNPs-8, a 7.17% degradation occurs, as shown in Figure 8F. A previous study by Elango et al. (2016) reported that MB 85% degrades in 70 min of spherical shape SNPs. (42) However, there are no similar studies using SNPs.
Table 2. Comparison of Photocatalytic Activity of SNP Samples to the Degradation of MB and MO Dyes by Different Nanocatalysts
Rmaterial usedNPs dose/dye dosesize/shapetimedegradation efficiency (%)refs
MBSnO2/SnO NPs50.0014–70 nm/spherical shape180 min90.28% MB (46)
 Sr-doped ZnO nanocatalyst33.3325–45 nm/hexagonal120 min78.50% MB (47)
 spindle-like TiO2100.0050–70 nm/spindle-like120 min62.70% MB (48)
 Cu/MMT nanocatalyst31.268 nm/spherical shape120 min95.06% MB (49)
 SNPs-2 at 200 °C33.333.97 nm/spherical shape30 min98.56% MBPresent Work
 SNPs-5 at 500 °C33.338.48 nm/spherical shape50 min97.84% MBPresent Work
 SNPs-8 at 800 °C33.3313.43 nm/spherical shape80 min97.12% MBPresent Work
MOZnO nanocatalyst100.0040 nm/spherical shape120 min83.99% MO (50)
 SnO2 nanocatalyst100.0010–42 nm/spherical shape120 min94.00% MO (51)
 NiFe2O4 nanocatalyst50.0034.74 nm/quasi globular-shaped300 min72.66% MO (52)
 Fe nanocatalyst12.007–14 nm/tetragonal shaped100 min95.00% MO (53)
 SNPs-2 at 200 °C33.333.97 nm/spherical shape40 min98.26% MOPresent Work
 SNPs-5 at 500 °C33.338.48 nm/spherical shape70 min97.39% MOPresent Work
 SNPs-8 at 800 °C33.3313.43 nm/spherical shape100 min96.52% MOPresent Work

4.2. Photocatalytic Activity for the MO Dye Degradation

The photocatalytic activity analysis was performed to evaluate the efficiency of the green synthesized SNPs in the degradation of MO under UV light irradiation. Figure 9 shows the absorbance spectra of MO dye degradation noticed between 200 and 800 nm with time intervals. The maximum absorbance occurs at λmax ∼ 464 nm in the UV–vis spectra. In the case of the catalyst, the degradation of MO dye in sample SNPs-2 drastically reached up to 98.26% in 40 min under UV light exposure in the UV–vis spectrum; no intermediate peaks are detected in Figure 9A. In the case of sample SNPs-5, the MO dye degradation reaches up to 97.39% within 70 min, as shown in Figure 9B, and in the case of sample SNPs-8, the degradation significantly reaches 96.52% within 100 min in the presence of UV light irradiation as represented in Figure 9C.

Figure 9

Figure 9. Photocatalytic activity of SNPs-2 for the degradation of MO dye (A), SNPs-5 for the degradation of MO dye (B), SNPs-8 for the degradation of MO dye (C), SNPs-2 percentage of MO dye degradation (D), SNPs-5 percentage of MO dye degradation (E), and SNPs-8 percentage of MO dye degradation (F).

However, the degradation of MO dye does not change significantly without a catalyst; only 9.56% degrades under UV light irradiation, as represented in Figure 9D. Similarly, the degradation of MO in the case of sample SNPs-5 is around 8.69%, as illustrated in Figure 9E. In contrast, for sample SNPs-8, the 10.43% degradation occurs within 100 min, as shown in Figure 9F. Our results are pretty similar to those reported by Yao et al. (2021), which report MO dye degradation 91.3% after 14 h in the presence of visible light irradiation. (43) Hence, the above III determined well the % of MO dye degradation and the effectiveness of the photocatalytic treatment process. The obtained results of photodegradation of MB and MO were compared with the degradation time and the % of degradation reported literature (see Table 2).

4.3. Effect of Catalyst Dose on MO/Methyl Blue Degradation

MO/MB dye degradation was tested with 10, 20, and 30 mg catalyst sample SNPs-2 at a constant concentration of MB and MO pollutants (10 mg/L). The photodegradation efficiency of MB gradually increased from 96.42 to 99.28%, while under the same condition, the MO dye degradation efficiency increased from 96.61 to 99.15%, as indicated in Table 3. The photocatalytic of MB/MO dyes remediation was estimated from the pseudo-first-order kinetics, as follow (IV).
(IV)
where r is the MB and MO dye degradation rate, C is the concentration of MB and MO dye solution, K is the absorption coefficient of MB and MO, t is the degradation time taken for MB/MO, and k is the constant rate reaction of MB/MO. Because the initial concentration of MB/MO is C0 = 10 mg/L, the above equation can be approximated to a pseudo-first-order model. For MB dye photodegradation, a similar procedure was previously employed to the GO/TiO2 nanocomposite loading by Kurniawan et al. (2020). (44) The degradation of MB/MO dye by photocatalyst increased, leading to more and more active sites generated, particularly on the photocatalyst surface, increasing the number of radicals. Therefore, the amount of catalyst enormously increased the rate constant, and consequently, the degradation time was reduced for completing the reaction as noticed and represented in Table 3. The results concluded that at the high dosage catalyst improves the MO/MB dye degradation efficiency, as is represented in Figure 10A,B. However, from the results, the amount of catalyst (mg) increases the rate constant (k) and decreases the degradation duration time (t), as shown in Table 3. From the result, we concluded that the photocatalytic properties of the SNP samples exhibited against the MB and MO dye degradation. However, the degradation efficiency depends on the smaller size SNP samples (3.97, 8.48, and 13.43 nm, respectively), lower band gap energies (2.68, 2.92, and 3.14 eV, respectively), and surface area (153.20, 90.34, and 43.72 m2/g, respectively), which play a key role in the photocatalytic dye degradation.

Figure 10

Figure 10. Kinetic plots of MB and MO dye degradation, (A) MB degradation kinetics curves for varying catalyst doses (10–30 mg), (B) MO degradation kinetics curves for varying catalyst doses (10–30 mg), (C) effect of dye concentration (5–15 mg/L) on the degradation of MB dye, and (D) effect of dye concentration (5–15 mg/L) on the degradation of MO dye.

Table 3. Degradation (%) and with Ratio of Catalyst Dose (mg) to Dye Dose (mg) and Rate Constant of MB and MO Dye
dye nameNPs dose (mg)/dye dose (mg)time (min)degradation (%)rate constant (k) min–1
MB16.666096.420.0858
 33.333098.560.2053
 50.002099.280.3316
 66.661599.590.3408
 33.333098.560.2053
 22.227097.910.1211
MO16.667096.610.1082
 33.334098.260.2025
 50.003099.150.3174
 66.662098.480.3783
 33.334098.260.2025
 22.228096.960.1355

4.4. Effect of MO/MB Dye Concentration

The effect of MO/MB dye concentration on the degradation of MO and MB at different dye concentrations (5, 10, and 15 mg/L) is analyzed at a particular constant catalyst dose (20 mg) of photocatalyst sample SNPs-2. Figure 10C,D shows that the degradation efficiency of MB and MO dyes gradually decreased, resulting in longer degradation times. The MB and MO pollutant concentrations 5 to 15 mg/L were tested with 20 mg of SNPs-2 sample photocatalyst under UV light exposure. Table 3 shows that the MB and MO dye photodegradation rate decreased while the MB and MO dye concentration increased from 5 to 15 mg/L. However, the MB and MO dye photodegradation efficiency rate did not increase significantly. It is at a constant rate even upon increasing the concentration of the MB and MO dye solution. Similarly, Chen et al. (2017) reported the effect of dye solution concentration at a particular constant catalyst (ZnO NPs) for Congo red, MO, and DB38 dye degradation efficiency in the presence of photoirradiation. (45)

4.5. Degradation Efficiency Comparison with Published Reports

The results for the degradation efficiency of MB and MO dyes providing the corresponding ratio of SNP samples nanocatalyst dose (mg) to dye concentration (mg/mg), size, surface area, shape, band gap energy of the NPs, degradation efficiency, and time (min) are listed in Table 2. Our results are compared to similar reported works, including the photocatalyst SnO/SnO2 hybrids, spindle-like TiO2 nanocatalysts, Sr-doped ZnO nanocatalysts, as well as copper nanoparticles supported with montmorillonite clays nanocatalyzed for the degradation efficiency of MB dye. (46−49) Moreover, ZnO nanocatalysts, SnO2 nanocatalysts, and Fe nanocatalyst degradation of MO dye reported in references (50)() () (53) and degradation efficiencies are compared in the table.
It is noticed that the degradation time depends on the ratio of the SNP photocatalyst dose to dye concentration (mg/mg), shape, surface area, and particle size of the NPs, as represented in Table 2. However, there have been recent studies that report that during the degradation of MB and MO by nano photocatalysts, massive free radicals that represent ROS are released, reacting with the dye molecules and leading to nontoxic products such as mineral acids, water molecules (H2O), and carbon dioxide (CO2), which support our observation. Thus, the synthesized SNP samples exhibit extremely enhanced degradation efficiency summarized to the recently reported literature given in Table 2. However, it strongly depends on the smaller particle size and uniform spherical shape of the nanoparticles. The MB degradation 98.56 percentage degraded within 30 min for a rate constant (k) 0.2053 min–1, similar to MO 98.26 percentage degradation within 40 min concerning rate constant (k) 0.2025 min–1. However, the SNP sample photocatalyst dye degradation reaction experiment was processed at a particular constant (20 mg) photocatalyst and 60 mL of volume of dye solution at a specific dye concentration (10 mg/L) under UV light exposure. Similar experimental studies of the effect of photocatalytic at varied catalyst doses and concentrations of dye solution and that provide results of rate constant (k), degradation efficiency, and degradation time are tabulated in Table 2. Therefore, the synthesized SNP samples exhibited extraordinary photocatalytic activity performance and could be promising materials utilized in degradation water purification and toxic and hazardous organic pollutant dyes.

4.6. Possible Mechanism of MB and MO Dye Degradation

The reasonable possible mechanism of MB and MO dye degradation demonstrated by the obtained semiconductor SnO2 nanomaterials in the present work is represented in Scheme 1. Photocatalyst semiconductors have many characteristics and properties (such as small size, high surface area, low band gap energy, tetragonal structure, and so forth), which can influence the removal of MB and MO dyes. For example, the tetragonal structure of tin oxide nanomaterials has shown high photocatalytic activity and feasibility in dye removal. (54) The photocatalytic activity described in Scheme 1 is carried out in two steps. The first step includes the absorbance of dye molecules on the surface of the SNP sample during the constant stirring process without UV light exposure. The second step consists of forming electron–hole pairs, which generates superoxide radicals and hydroxide radicals in the SNP matrix, which degrade the dye molecules to complete mineralization (H2O and CO2) of the pollutant molecules.

Scheme 1

Scheme 1. Possible Mechanism of MB and MO Dye Degradation by SNPs Obtained in This Work
The feasible mechanism describes the various reaction succession steps occurring during the degradation of MB and MO dyes in the presence of SnO2 nanomaterials under UV light irradiation and assumes the excess of highly ROS in the solution. Vidya et al. (2017) proposed a similar mechanism for the ZnO NP photocatalytic activity of Cong red dye degradation in the presence of UV light exposure. (55) The possible mechanism involves sequence steps described by the following equations. UV light bombarding the SNP nanomaterials generates electrons and transfers electrons from the valence band to the conduction band with corresponding energy higher than the band gap energy of the SNP nanomaterials that should promote holes in the valence band and generate electrons in the conduction band. This process might generate high ROS species that are reactive with MB and MO dye molecules such as successive carbon dioxide and water molecules. At the beginning, under UV light irradiation, electrons moved from the valence band to the conduction band on the surface of the SNPs, which is represented in eq 1. In this way, conduction band electrons (eCB) readily react with oxygen to generate superoxide (O2–•) ions, as shown in eq 2, and subsequently, these highly reactive oxide ions react with hydrogen (H+) ions to release the −OOH, as represented in eq 3, whereas in the valence band, (hvVB+) holes readily contact with water molecules to produce hydroxide (−OH) ions which in turn generate highly active hydroxide radicals (−OH), as represented in eqs 4 and 5. The produced −OOH can readily react with conduction-band electrons and hydrogen ions to produce highly active superoxide (H2O2) molecules. Eventually, the cleavage of hydrogen peroxide (H2O2) molecules occurs to produce the highly active hydroxyl radical (−OH), and also, the photoelectrons diminish the oxygen molecules (O2) adsorbed on the SNP catalyst surface generating superoxide (O2–•) radicals. Eventually, MB and MO dye molecules are decomposed by the formed highly active hydroxyl (OH) radicals and superoxide (O2–•) radicals to produce mineral acids, H2O, and CO2, as shown in eqs 6–11. This mechanism is similar to that proposed by Zangeneh et al., (2015), which assumes the formation of radicals during dyes’ degradation. (56)
Liu et al., (2021) reported charge-carrier trapping on the surface of the materials based on the band gap energy. The higher level of energy than the band gap of the materials leads to electron transfer from the conduction band to the valence band and generating hydroxide and superoxide radical’s degradation of the dye molecules. (57) The photocatalytic reactions take place provided that the semiconductor receives irradiation photons with energy greater than the band gap. An irradiating photon with insufficient energy leads to the failure of electron transition and then degradation is not possible to carbon dioxide and water molecules. Li et al. (2021) reported reactive species scavenging experiments and electron spin resonance which revealed that hydroxyl radicals (OH), superoxide radical anions (O2–•), and photogenerated holes (h+) played critical roles for the photocatalytic degradation. (58)
Recent reports discussed the effect of the pH for the degradation of MB and MO dyes. Alkaykh et al. (2020) reported the effect of pH for the MB dye degradation, where at pH2, it shows the 15% removal of MB dyes; however, at pH 9, it shows 60% MB dye removal. (59) Wu et al. (2022) studied the effect of pH from 4 to 9 for the methylene dye degradation that possesses under acidic condition pH 4 is 63.88% degradation of the MB dye; however, at pH 9 basic condition, it shows 60.73% degradation, but at pH 6 neutral condition, it shows the 95.94% degradation of MB dye removal. (60) Abbasi and Hasanpour, (2017) studied the effect of pH from 4 to 10 that shows at pH 7 its higher efficiency of 99% MO dye removal than at pH 4 and pH 10. (61) Cai et al. (2017) studied the effect of pH at 3–9 that represents that at pH 3 and 5, it shows the higher decolorization efficiency of MO at 99 and 98%; however, at pH 9 alkaline medium, it shows 88% of MO decolorization efficiency. (62) Adeel et al. (2021) studied the effect of pH at 2–10; however, at pH 4 and 2, the highest photodegradation of MO rate constant was reported, that is, 0.0144 and 0.0099 min–1, but in the case of at pH 10, 0.0043 min–1 degradation efficiency was reported. (63)

4.7. Reusability of the SNP Sample

The reusability stability of the synthesized photocatalyst sample SNPs-2 has been studied. The SNPs-2 sample residue was collected after the degradation of MB and MO under UV-light exposure, centrifugated, washed, and then reused five consecutive cycles under identical conditions. The removal of MB and MO using the SNPs-2 sample offer after the firstt cycle the highest photodegradation 98.56% for MB and 98.26% for MO. After that, the second cycle removal for MB and MO was 97.84 and 96.52%, respectively. In the case of the third cycle, the MB and MO degradations were 97.12 and 95.65%, respectively. In the case of the fourth cycle, the MB and MO degradation reached 96.54 and 93.91%, respectively. Eventually, in the fifth cycle, the photocatalytic activity gradually decreased to 95.82% for MB and 92.17% for MO. The degradation versus cycle plot is shown in Figure 11A. Thus, sample SNPs-2 exhibits excellent reusability for the photodegradation of MB and MO pollutant dyes. Moreover, to authenticate the stability of sample SNPs-2 after five cycles, any phase change was analyzed by powder XRD (see Figure 11B). Based on the results, the crystallization of the recovered catalyst remains unchanged even after five cycles of use when compared to the as synthesized SNPs-2 sample. Therefore, we concluded that the synthesized SNPs-2 sample is stable and maintains its outstanding photocatalytic activity for consecutive cycles.

Figure 11

Figure 11. Recyclability check of biosynthesized sample SNPs-3 for the degradation of MB and MO under identical experimental conditions (A) and XRD pattern after five consecutive cycles for sample SNPs-3 (B).

5. Conclusions

Click to copy section linkSection link copied!

The present study carried out a simple, nontoxic, eco-friendly, and cost-effective method to produce SNP nanomaterials through the agro-waste cotton boll peel aqueous extract. SNPs were successfully synthesized at different temperatures of 200, 500, and 800 °C without involving toxic chemicals or solvents. XRD patterns confirmed that the crystal structure is tetragonal, and the average crystallite sizes for the SNP samples obtained after 200, 500, and 800 °C are 4.72, 9.68, and 14.06 nm, respectively. HR-TEM images manifested spherical shaped samples with sizes 3.79, 8.48, and 13.43 nm for SNP-2, SNP-5, and SNP-8, respectively. The samples were further analyzed by FT-IR, UV–vis spectroscopy, and BET analysis, and their photocatalytic activity for the degradation of MB and MO was inspected. From the result, we concluded that the smaller size of the SNPs-2 sample (3.79 nm), the lower band gap energy (2.68 eV) and higher the surface area (153.20 m2/g) exhibited superior performance of photocatalytic efficiency for the degradation of MB/MO. Therefore, MB dye 98.56% degradation within 30 min and MO dye 98.26% degradation within 40 min, respectively, under UV light exposure. Moreover, we have also studied the effect of catalyst and the effect of dye concentration. As the amount of catalyst (10 mg to 30 mg) increases, the rate constant increases and decreases the degradation duration time. However, when the dye concentration (5–15 mg/L) increases, the rate constant decreases, and the degradation duration time increases. Consequently, the synthesized SNP samples are promising materials to be used as a photocatalyst for the remediation of pollutant dyes.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
    • Luis De Los Santos Valladares - Laboratorio de Cerámicos y Nanomateriales, Facultad de Ciencias Físicas, Universidad Nacional Mayor de San Marcos, Ap. Postal 14-0149, Lima 14, PeruCavendish Laboratory, Department of Physics, University of Cambridge, J.J. Thomson Avenue, Cambridge CB3 OHE, U.K.School of Materials Science and Engineering, Northeastern University, No 11, Lane 3, Wenhua Road, Heping District, Shenyang 110819, Liaoning, People’s Republic of ChinaOrcidhttps://orcid.org/0000-0001-5930-9916 Email: [email protected]
    • Pratap Kollu - CASEST, School of Physics, University of Hyderabad, Prof. C. R Rao Road, Gachibowli, Hyderabad 500046, Telangana, IndiaOrcidhttps://orcid.org/0000-0001-6250-9396 Email: [email protected]
  • Authors
    • Boya Palajonnala Narasaiah - CASEST, School of Physics, University of Hyderabad, Prof. C. R Rao Road, Gachibowli, Hyderabad 500046, Telangana, IndiaLaboratorio de Cerámicos y Nanomateriales, Facultad de Ciencias Físicas, Universidad Nacional Mayor de San Marcos, Ap. Postal 14-0149, Lima 14, Peru
    • Pravallika Banoth - CASEST, School of Physics, University of Hyderabad, Prof. C. R Rao Road, Gachibowli, Hyderabad 500046, Telangana, India
    • Arya Sohan - CASEST, School of Physics, University of Hyderabad, Prof. C. R Rao Road, Gachibowli, Hyderabad 500046, Telangana, India
    • Badal Kumar Mandal - Department of Chemistry, School of Advanced Sciences, Vellore Institute of Technology, Vellore 632014, Tamil Nadu, India
    • Angel G. Bustamante Dominguez - Laboratorio de Cerámicos y Nanomateriales, Facultad de Ciencias Físicas, Universidad Nacional Mayor de San Marcos, Ap. Postal 14-0149, Lima 14, Peru
  • Author Contributions

    B.P.N. and P.B. are equally contributing first authors.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

This work was supported by the “Incorporation de Investigadores program” CONCYTEC─FONDECYT. UNMSM (contrat no. 12-2019-FONDECYT-BM-INC.INV.). P.K., P.B., and A.S. thank the Institute of Eminence (IoE), Ministry of Human Resource Development (MHRD), Grant number: UoH-IoE-RC2-21-017 at University of Hyderabad, India.

References

Click to copy section linkSection link copied!

This article references 63 other publications.

  1. 1
    Mourdikoudis, S.; Kostopoulou, A.; LaGrow, A. P. Magnetic Nanoparticle Composites: Synergistic Effects and Applications. Adv. Sci. 2021, 8, 2004951,  DOI: 10.1002/advs.202004951
  2. 2
    Singh, J.; Dutta, T.; Kim, K. H.; Rawat, M.; Samddar, P.; Kumar, P. Green synthesis of metals and their oxide nanoparticles: applications for environmental remediation. J. Nanobiotechnol. 2018, 16, 84,  DOI: 10.1186/s12951-018-0408-4
  3. 3
    Iravani, S. Green synthesis of metal nanoparticles using plants. Green Chem. 2011, 13, 26382650,  DOI: 10.1039/c1gc15386b
  4. 4
    Shamaila, S.; Sajjad, A. K. L.; Ryma, N.-u. -A.; Farooqi, S. A.; Jabeen, N.; Majeed, S.; Farooq, I. Advancements in nanoparticle fabrication by hazard free eco-friendly green routes. Appl. Mater. Today 2016, 5, 150199,  DOI: 10.1016/j.apmt.2016.09.009
  5. 5
    Kumar, K. M.; Mandal, B. K.; Tammina, S. K. Green synthesis of nano platinum using naturally occurring polyphenols. RSC Adv. 2013, 3, 40334039,  DOI: 10.1039/c3ra22959a
  6. 6
    Mazari, S. A.; Ali, E.; Abro, R.; Khan, F. S. A.; Ahmed, I.; Ahmed, M.; Nizamuddin, S.; Siddiqui, T. H.; Hossain, N.; Mubarak, N. M.; Shah, A. Nanomaterials: Applications, waste-handling, environmental toxicities, and future challenges - A review. J. Environ. Chem. Eng. 2021, 9, 105028,  DOI: 10.1016/j.jece.2021.105028
  7. 7
    Palanisamy, G.; Bhuvaneswari, K.; Srinivasan, M.; Vignesh, S.; Elavarasan, N.; Venkatesh, G.; Pazhanivel, T.; Ramasamy, P. Two-dimensional g-C3N4 nanosheets supporting Co3O4-V2O5 nanocomposite for remarkable photodegradation of mixed organic dyes based on a dual Z-scheme photocatalytic system. Diamond Relat. Mater. 2021, 118, 108540,  DOI: 10.1016/j.diamond.2021.108540
  8. 8
    Choi, E.; Lee, D.; Shin, H. J.; Kim, N.; Valladares, L. D. L. S.; Seo, J. Role of oxygen vacancy sites on the temperature-dependent photoluminescence of SnO2 nanowires. J. Phys. Chem. C 2021, 125, 1497414978,  DOI: 10.1021/acs.jpcc.1c02937
  9. 9
    Suthakaran, S.; Dhanapandian, S.; Krishnakumar, N.; Ponpandian, N. Hydrothermal synthesis of SnO2 nanoparticles and its photocatalytic degradation of methyl violet and electrochemical performance. Mater. Res. Express 2019, 6, 0850i3,  DOI: 10.1088/2053-1591/ab29c2
  10. 10
    Nejati-Moghadam, L.; Esmaeili, B. K. A.; Salavati, N. M.; Safardoust, H. Synthesis and characterisation of SnO2 nanostructures prepared by a facile precipitation method. J. Nanostruct. 2015, 5, 4753,  DOI: 10.7508/JNS.2015.01.007
  11. 11
    Zhang, L.; Ren, X.; Luo, Y.; Shi, X.; Asiri, A. M.; Li, T.; Sun, X. Ambient NH3 synthesis via electrochemical reduction of N2 over cubic sub-micron SnO2 particles. Chem. Commun. 2018, 54, 1296612969,  DOI: 10.1039/C8CC06524A
  12. 12
    Haspulat, B.; Sarıbel, M.; Kamış, H. Surfactant assisted hydrothermal synthesis of SnO nanoparticles with enhanced photocatalytic activity. Arab. J. Chem. 2020, 13, 96108,  DOI: 10.1016/j.arabjc.2017.02.004
  13. 13
    Maharajan, S.; Kwon, N. H.; Brodard, P.; Fromm, K. M. A Nano-Rattle SnO2@carbon Composite Anode Material for High-Energy Li-ion Batteries by Melt Diffusion Impregnation. Nanomaterials 2020, 10, 804,  DOI: 10.3390/nano10040804
  14. 14
    Karmaoui, M.; Jorge, A. B.; McMillan, P. F.; Aliev, A. E.; Pullar, R. C.; Labrincha, J. A.; Tobaldi, D. M. One-Step Synthesis, Structure, and Band Gap Properties of SnO2 Nanoparticles Made by a Low Temperature Nonaqueous Sol-Gel Technique. ACS Omega 2018, 3, 1322713238,  DOI: 10.1021/acsomega.8b02122
  15. 15
    Zhu, K.; Lv, Y.; Liu, J.; Wang, W.; Wang, C.; Li, S.; Wang, P.; Zhang, M.; Meng, A.; Li, Z. Facile fabrication of g-C3N4/SnO2 composites and ball milling treatment for enhanced photocatalytic performance. J. Alloys Compd. 2019, 802, 1318,  DOI: 10.1016/j.jallcom.2019.06.193
  16. 16
    Chavali, M. S.; Nikolova, M. P. Metal oxide nanoparticles and their applications in nanotechnology. SN Appl. Sci. 2019, 1, 607,  DOI: 10.1007/s42452-019-0592-3
  17. 17
    Yuliarto, B.; Septiani, N. L. W.; Kaneti, Y. V.; Iqbal, M.; Gumilar, G.; Kim, M.; Na, J.; Wu, K. C.-W.; Yamauchi, Y. Green synthesis of metal oxide nanostructures using naturally occurring compounds for energy, environmental, and bio-related applications. New J. Chem. 2019, 43, 1584615856,  DOI: 10.1039/c9nj03311d
  18. 18
    Diallo, A.; Manikandan, E.; Rajendran, V.; Maaza, M. Physical & enhanced photocatalytic properties of green synthesized SnO2 nanoparticles via Aspalathus linearis. J. Alloys Compd. 2016, 681, 561570,  DOI: 10.1016/j.jallcom.2016.04.200
  19. 19
    Matussin, S.; Harunsani, M. H.; Tan, A. L.; Khan, M. M. Plant-Extract-Mediated SnO2 Nanoparticles: Synthesis and Applications. ACS Sustain. Chem. Eng. 2020, 8, 30403054,  DOI: 10.1021/acssuschemeng.9b06398
  20. 20
    Singh, J.; Kaur, H.; Kukkar, D.; Mukamia, V. K.; Kumar, S.; Rawat, M. Green synthesis of SnO2 NPs for solar light induced photocatalytic applications. Mater. Res. Express 2019, 6, 115007,  DOI: 10.1088/2053-1591/ab4412
  21. 21
    Bhattacharjee, A.; Ahmaruzzaman, M. A novel and green process for the production of tin oxide quantum dots and its application as a photocatalyst for the degradation of dyes from aqueous phase. J. Colloid Interface Sci. 2015, 448, 130139,  DOI: 10.1016/j.jcis.2015.01.083
  22. 22
    Sinha, T.; Ahmaruzzaman, M.; Adhikari, P. P.; Bora, R. Green and Environmentally Sustainable Fabrication of Ag-SnO2 Nanocomposite and Its Multifunctional Efficacy As Photocatalyst and Antibacterial and Antioxidant Agent. ACS Sustain. Chem. Eng. 2017, 5, 46454655,  DOI: 10.1021/acssuschemeng.6b03114
  23. 23
    Sudhaparimala, S.; Vaishnavi, M. Biological synthesis of nano composite SnO2- ZnO - Screening for efficient photocatalytic degradation and antimicrobial activity. Mater. Today Proc. 2016, 3, 23732380,  DOI: 10.1016/j.matpr.2016.04.150
  24. 24
    Li, A. J.; Pal, V. K.; Kannan, K. A review of environmental occurrence, toxicity, biotransformation and biomonitoring of volatile organic compounds. Environ. Chem. Ecotoxicol. 2021, 3, 91116,  DOI: 10.1016/j.enceco.2021.01.001
  25. 25
    Rueda-Marquez, J. J.; Levchuk, I.; Fernández Ibañez, P.; Sillanpää, M. A critical review on application of photocatalysis for toxicity reduction of real wastewaters. J. Clean. Prod. 2020, 258, 120694,  DOI: 10.1016/j.jclepro.2020.120694
  26. 26
    Li, Y.; Chang, H.; Yan, H.; Tian, S.; Jessop, P. G. Reversible Absorption of Volatile Organic Compounds by Switchable-Hydrophilicity Solvents: A Case Study of Toluene with N,N-Dimethylcyclohexylamine. ACS Omega 2020, 6, 253264,  DOI: 10.1021/acsomega.0c04443
  27. 27
    Li, Y.; Liu, K.; Zhang, J.; Yang, J.; Huang, Y.; Tong, Y. Engineering the Band-Edge of Fe2O3/ZnO Nanoplates via Separate Dual Cation Incorporation for Efficient Photocatalytic Performance. Ind. Eng. Chem. Res. 2020, 59, 1886518872,  DOI: 10.1021/acs.iecr.0c03388
  28. 28
    Li, D.; Song, H.; Meng, X.; Shen, T.; Sun, J.; Han, W.; Wang, X. Effects of Particle Size on the Structure and Photocatalytic Performance by Alkali-Treated TiO2. Nanomaterials 2020, 10, 546,  DOI: 10.3390/nano10030546
  29. 29
    Tuan, P. V.; Hieu, L. T.; Tan, V. T.; Phuong, T. T.; Tran Thi Quýnh, H.; Khiem, T. N. The dependence of morphology, structure, and photocatalytic activity of SnO2/rGO nanocomposites on hydrothermal temperature. Mater. Res. Express 2019, 6, 106204,  DOI: 10.1088/2053-1591/ab1e12
  30. 30
    Jiang, Y.; Zhao, H.; Liang, J.; Yue, L.; Li, T.; Luo, Y.; Liu, Q.; Lu, S.; Asiri, A. M.; Gong, Z.; Sun, X. Anodic oxidation for the degradation of organic pollutants: anode materials, operating conditions and mechanisms. A mini review. Electrochem. Commun. 2021, 123, 106912,  DOI: 10.1016/j.elecom.2020.106912
  31. 31
    Li, Y.; Yang, Q.; Wang, Z.; Wang, G.; Zhang, B.; Zhang, Q.; Yang, D. Rapid fabrication of SnO2 nanoparticle photocatalyst: computational understanding and photocatalytic degradation of organic dye. Inorg. Chem. Front. 2018, 5, 30053014,  DOI: 10.1039/C8QI00688A
  32. 32
    Huang, Y.; Li, K.; Lin, Y.; Tong, Y.; Liu, H. Enhanced Efficiency of Electron-Hole Separation in Bi2 O2 CO3 for Photocatalysis via Acid Treatment. ChemCatChem 2018, 10, 19821987,  DOI: 10.1002/cctc.201800101
  33. 33
    Huang, Y.; Lu, Y.; Lin, Y.; Mao, Y.; Ouyang, G.; Liu, H.; Zhang, S.; Tong, Y. Cerium-based hybrid nanorods for synergetic photo-thermocatalytic degradation of organic pollutants. J. Mater. Chem. A 2018, 6, 2474024747,  DOI: 10.1039/c8ta06565a
  34. 34
    Siripireddy, B.; Mandal, B. K. Facile green synthesis of zinc oxide nanoparticles by Eucalyptus globulus and their photocatalytic and antioxidant activity. Adv. Powder Technol. 2017, 28, 785797,  DOI: 10.1016/j.apt.2016.11.026
  35. 35
    Osuntokun, J.; Onwudiwe, D. C.; Ebenso, E. E. Biosynthesis and Photocatalytic Properties of SnO2 Nanoparticles Prepared Using Aqueous Extract of Cauliflower. J. Cluster Sci. 2017, 28, 18831896,  DOI: 10.1007/s10876-017-1188-y
  36. 36
    Ma, C. M.; Hong, G. B.; Lee, S. C. Facile Synthesis of Tin Dioxide Nanoparticles for Photocatalytic Degradation of Congo Red Dye in Aqueous Solution. Catalysts 2020, 10, 792,  DOI: 10.3390/catal10070792
  37. 37
    Arumugam, M.; Manikandan, D. B.; Dhandapani, E.; Sridhar, A.; Balakrishnan, K.; Markandan, M.; Ramasamy, T. Green synthesis of zinc oxide nanoparticles (ZnO NPs) using Syzygium cumini: Potential multifaceted applications on antioxidants, cytotoxic and as nanonutrient for the growth of Sesamum indicum. Environ. Technol. Innovation 2021, 23, 101653,  DOI: 10.1016/j.eti.2021.101653
  38. 38
    Paramarta, V.; Taufik, A.; Saleh, R. Better adsorption capacity of SnO2 nanoparticles with different graphene addition. J. Phys.: Conf. Ser. 2016, 776, 012039,  DOI: 10.1088/1742-6596/776/1/012039
  39. 39
    Haq, S.; Rehman, W.; Waseem, M.; Javed, R.; Mahfooz-ur-Rehman, M.; Shahid, M. Effect of heating on the structural and optical properties of TiO2 nanoparticles: antibacterial activity. Appl. Nanosci. 2018, 8, 1118,  DOI: 10.1007/s13204-018-0647-6
  40. 40
    Ullah, H.; Khan, I.; Yamani, Z. H.; Qurashi, A. Sonochemical-driven ultrafast facile synthesis of SnO2 nanoparticles: Growth mechanism structural electrical and hydrogen gas sensing properties. Ultrason. Sonochem. 2017, 34, 484490,  DOI: 10.1016/j.ultsonch.2016.06.025
  41. 41
    Huang, Y.; Xu, H.; Yang, H.; Lin, Y.; Liu, H.; Tong, Y. Efficient charges separation using advanced BiOI-based hollow spheres decorated with palladium and manganese dioxide nanoparticles. ACS Sustainable Chem. Eng. 2018, 6, 27512757,  DOI: 10.1021/acssuschemeng.7b04435
  42. 42
    Elango, G.; Roopan, S. M. Efficacy of SnO 2 nanoparticles toward photocatalytic degradation of methylene blue dye. J. Photochem. Photobiol., B 2016, 155, 3438,  DOI: 10.1016/j.jphotobiol.2015.12.010
  43. 43
    Yao, X.; Zhang, B.; Cui, S.; Yang, S.; Tang, X. Fabrication of SnSO4-modified TiO2 for enhance degradation performance of methyl orange (MO) and antibacterial activity. Appl. Surf. Sci. 2021, 551, 149419,  DOI: 10.1016/j.apsusc.2021.149419
  44. 44
    Kurniawan, T. A.; Mengting, Z.; Fu, D.; Yeap, S. K.; Othman, M. H. D.; Avtar, R.; Ouyang, T. Functionalizing TiO2 with graphene oxide for enhancing photocatalytic degradation of methylene blue (MB) in contaminated wastewater. J. Environ. Manage. 2020, 270, 110871,  DOI: 10.1016/j.jenvman.2020.110871
  45. 45
    Chen, X.; Wu, Z.; Liu, D.; Gao, Z. Preparation of ZnO photocatalyst for the efficient and rapid photocatalytic degradation of azo dyes. Nanoscale Res. Lett. 2017, 12, 143,  DOI: 10.1186/s11671-017-1904-4
  46. 46
    Santhi, K.; Rani, C.; Karuppuchamy, S. Synthesis and characterisation of a novel SnO/SnO2 hybrid photocatalyst. J. Alloys Compd. 2016, 662, 102107,  DOI: 10.1016/j.jallcom.2015.12.007
  47. 47
    Yousefi, R.; Jamali-Sheini, F.; Cheraghizade, M.; Khosravi-Gandomani, S.; Sáaedi, A.; Huang, N. M.; Basirun, W. J.; Azarang, M. Enhanced visible-light photocatalytic activity of strontium-doped zinc oxide nanoparticles. Mater. Sci. Semicond. Process. 2015, 32, 152159,  DOI: 10.1016/j.mssp.2015.01.013
  48. 48
    Arunkumar, S.; Alagiri, M. Synthesis and Characterization of Spindle-Like TiO2 Nanostructures and Photocatalytic Activity on Methyl Orange and Methyl Blue Dyes Under Sunlight Radiation. J. Cluster Sci. 2017, 28, 26352643,  DOI: 10.1007/s10876-017-1245-6
  49. 49
    Mekewi, M. A.; Darwish, A. S.; Amin, M. S.; Eshaq, G.; Bourazan, H. A. Copper nanoparticles supported onto montmorillonite clays as efficient catalyst for methylene blue dye degradation. Egypt. J. Pet. 2016, 25, 269279,  DOI: 10.1016/j.ejpe.2015.06.011
  50. 50
    Karnan, T.; Selvakumar, S. A. S. Biosynthesis of ZnO nanoparticles using rambutan (Nephelium lappaceumL.) peel extract and their photocatalytic activity on methyl orange dye. J. Mol. Struct. 2016, 1125, 358365,  DOI: 10.1016/j.molstruc.2016.07.029
  51. 51
    Srivastava, N.; Mukhopadhyay, M. Biosynthesis of SnO2 Nanoparticles Using Bacterium Erwinia herbicola and Their Photocatalytic Activity for Degradation of Dyes. Ind. Eng. Chem. Res. 2014, 53, 1397113979,  DOI: 10.1021/ie5020052
  52. 52
    Hirthna; Sendhilnathan, S.; Rajan, P. I.; Adinaveen, T. Synthesis and Characterization of NiFe2O4 Nanoparticles for the Enhancement of Direct Sunlight Photocatalytic Degradation of Methyl Orange. J. Supercond. Novel Magn. 2018, 31, 33153322,  DOI: 10.1007/s10948-018-4601-3
  53. 53
    Radini, I. A.; Hasan, N.; Malik, M. A.; Khan, Z. Biosynthesis of iron nanoparticles using Trigonella foenum-graecum seed extract for photocatalytic methyl orange dye degradation and antibacterial applications. J. Photochem. Photobiol., B 2018, 183, 154163,  DOI: 10.1016/j.jphotobiol.2018.04.014
  54. 54
    Najjar, M.; Hosseini, H. A.; Masoudi, A.; Sabouri, Z.; Mostafapour, A.; Khatami, M.; Darroudi, M. Green chemical approach for the synthesis of SnO2 nanoparticles and its application in photocatalytic degradation of Eriochrome Black T dye. Optik 2021, 242, 167152,  DOI: 10.1016/j.ijleo.2021.167152
  55. 55
    Vidya, C.; Manjunatha, C.; Chandraprabha, M. N.; Rajshekar, M.; Raj M.A.L, A. Hazard free green synthesis of ZnO nano-photo-catalyst using Artocarpus Heterophyllus leaf extract for the degradation of Congo red dye in water treatment applications. J. Environ. Chem. Eng. 2017, 5, 31723180,  DOI: 10.1016/j.jece.2017.05.058
  56. 56
    Zangeneh, H.; Zinatizadeh, A. A. L.; Habibi, M.; Akia, M.; Hasnain Isa, M. Photocatalytic oxidation of organic dyes and pollutants in wastewater using different modified titanium dioxides: A comparative review. J. Ind. Eng. Chem. 2015, 26, 136,  DOI: 10.1016/j.jiec.2014.10.043
  57. 57
    Liu, J.; Zhang, Q.; Tian, X.; Hong, Y.; Nie, Y.; Su, N.; Jin, G.; Zhai, Z.; Fu, C. Highly efficient photocatalytic degradation of oil pollutants by oxygen deficient SnO2 quantum dots for water remediation. Chem. Eng. J. 2021, 404, 127146,  DOI: 10.1016/j.cej.2020.127146
  58. 58
    Li, Li.; Huang, J.; Li, R.; Chen, P.; Chen, D.; Cai, M.; Liu, G.; Feng, Y.; Lv, W.; Liu, G. Synthesis of a carbon dots modified g-C3N4/SnO2 Z-scheme photocatalyst with superior photocatalytic activity for PPCPs degradation under visible light irradiation. J. Hazard. Mater. 2021, 401, 123257,  DOI: 10.1016/j.jhazmat.2020.123257
  59. 59
    Alkaykh, S.; Mbarek, A.; Ali-Shattle, E. E. Photocatalytic degradation of methylene blue dye in aqueous solution by MnTiO3 nanoparticles under sunlight irradiation. Heliyon 2020, 6, e03663  DOI: 10.1016/j.heliyon.2020.e03663
  60. 60
    Wu, K.; Shi, M.; Pan, X.; Zhang, J.; Zhang, X.; Shen, T.; Tian, Y. Decolourization and biodegradation of methylene blue dye by a ligninolytic enzyme-producing Bacillus thuringiensis: degradation products and pathway. Enzyme Microb. Technol. 2022, 156, 109999,  DOI: 10.1016/j.enzmictec.2022.109999
  61. 61
    Abbasi, S.; Hasanpour, M. The effect of pH on the photocatalytic degradation of methyl orange using decorated ZnO nanoparticles with SnO2 nanoparticles. J. Mater. Sci.: Mater. Electron. 2017, 28, 13071314,  DOI: 10.1007/s10854-016-5660-5
  62. 62
    Cai, R.; Zhang, B.; Shi, J.; Li, M.; He, Z. Rapid Photocatalytic Decolorization of Methyl Orange under Visible Light Using VS4/Carbon Powder Nanocomposites. ACS Sustain. Chem. Eng. 2017, 5, 76907699,  DOI: 10.1021/acssuschemeng.7b01137
  63. 63
    Adeel, M.; Saeed, M.; Khan, I.; Muneer, M.; Akram, N. Synthesis and Characterization of Co-ZnO and Evaluation of Its Photocatalytic Activity for Photodegradation of Methyl Orange. ACS Omega 2021, 6, 14261435,  DOI: 10.1021/acsomega.0c05092

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 27 publications.

  1. Yobsan Likasa Tiki, Leta Deressa Tolesa, Ardila Hayu Tiwikrama, Tolesa Fita Chala. Ginger (Zingiber officinale)-Mediated Green Synthesis of Silver-Doped Tin Oxide Nanoparticles and Evaluation of Its Antimicrobial Activity. ACS Omega 2024, 9 (10) , 11443-11452. https://doi.org/10.1021/acsomega.3c07855
  2. Nalinee Kanth Kadiyala, Badal Kumar Mandal, L. Vinod Kumar Reddy, Crispin H. W. Barnes, Luis De Los Santos Valladares, Sireesh Babu Maddinedi, Dwaipayan Sen. Biofabricated Palladium Nanoparticle-Decorated Reduced Graphene Oxide Nanocomposite Using the Punica granatum (Pomegranate) Peel Extract: Investigation of Potent In Vivo Hepatoprotective Activity against Acetaminophen-Induced Liver Injury in Wistar Albino Rats. ACS Omega 2023, 8 (27) , 24524-24543. https://doi.org/10.1021/acsomega.3c02643
  3. Nisha Dhiman, Vivek Sharma, Somnath Ghosh. Perspective on Biomass-Based Cotton-Derived Nanocarbon for Multifunctional Energy Storage and Harvesting Applications. ACS Applied Electronic Materials 2023, 5 (4) , 1970-1991. https://doi.org/10.1021/acsaelm.3c00022
  4. Ramya Raman, Sudhahar Sakkarapani, Bhaskaran Arumugam. Eco-friendly synthesis trend of Sn-V bimetallic nanoparticles and its potential biological applications. Materials Letters 2025, 387 , 138277. https://doi.org/10.1016/j.matlet.2025.138277
  5. Tsung-Wei Zeng, Yi-Fang Liang. Effects of water/ethanol solvent ratios on the morphology of SnO2 nanorods. Inorganic Chemistry Communications 2025, 175 , 114148. https://doi.org/10.1016/j.inoche.2025.114148
  6. Lethula E. Mofokeng, Edwin Makhado, Patrick Ndungu. A comprehensive review on biogenic metallic, metal oxide and metal sulfide nanoparticles for water treatment and sensing applications. Journal of Industrial and Engineering Chemistry 2025, 144 , 48-76. https://doi.org/10.1016/j.jiec.2024.09.039
  7. Priyanka Pareek, Himanshi Saini, Lalita Ledwani. Green Synthesis of Tin Oxide (SnO2) Nanoparticles Using Pomelo Fruit Peel for Photocatalytic Degradation of Methylene Blue Dye and Industrial Dye. Topics in Catalysis 2025, 68 (3-4) , 252-272. https://doi.org/10.1007/s11244-024-02031-6
  8. Kimihiro Tani, Suguru Sakamoto, Yukie Tatsumoto, Masanao Imai, Kazumitsu Naoe. Green and Sustainable Synthesis of Silver Nanoparticles Using Wastes of Crude Drugs for Traditional Medicinal Use in Nara, Japan. 2025, 275-294. https://doi.org/10.1007/978-3-031-79062-1_11
  9. Niteen Jawale, Govind Umarji, Shubhangi Damkale, Sudhir Arbuj. Enhanced photocatalytic performance of hydrothermally synthesized TiO2 nanowires for H2 production via water splitting. Next Energy 2025, 6 , 100205. https://doi.org/10.1016/j.nxener.2024.100205
  10. Yonas Etafa Tasisa, Tridib Kumar Sarma, Tarun Kumar Sahu, Ramaswamy Krishnaraj. Phytosynthesis and characterization of tin-oxide nanoparticles (SnO2-NPs) from Croton macrostachyus leaf extract and its application under visible light photocatalytic activities. Scientific Reports 2024, 14 (1) https://doi.org/10.1038/s41598-024-60633-2
  11. Sayyed Hamed Adyani, Esmaiel Soleimani. Green synthesis of magnetic silver nanocomposite: the photocatalytic performance of nanocomposite to decolorize organic dyes. Environmental Technology 2024, 45 (25) , 5244-5258. https://doi.org/10.1080/09593330.2023.2286453
  12. R. Shalini, K. Ravichandran, P. Kavitha, P. K. Praseetha, R. Mohan, P. Ravikumar. Biocatalyst coupling with Mo-doped SnO 2 nanoparticles for efficient photocatalytic dye degradation: An eco-friendly approach for environmental remediation. Biocatalysis and Biotransformation 2024, 42 (5) , 605-619. https://doi.org/10.1080/10242422.2023.2289337
  13. Samuel Demarema, Mahmoud Nasr, Shinichi Ookawara, Amal Abdelhaleem. New insights into green synthesis of metal oxide based photocatalysts for photodegradation of organic pollutants: A bibliometric analysis and techno-economic evaluation. Journal of Cleaner Production 2024, 463 , 142679. https://doi.org/10.1016/j.jclepro.2024.142679
  14. Suganya Govindasamy, Balu Mahendran Gunasekaran, Nandhakumar Vaiyapuri, Balasubramanian Natarajan, Noel Nesakumar, Bargavi Varatharajan, Princess Gracia John Britto, Ariharan Arjunan. Synthesis of tin oxide nanoparticles by chemical and biological methods and their applications in high performance supercapacitor electrode, antibacterial and antifungal activity. Physica Scripta 2024, 99 (7) , 075042. https://doi.org/10.1088/1402-4896/ad5770
  15. Debasish Borah, Puja Saikia, Jayashree Rout, Debika Gogoi, Narendra Nath Ghosh, Chira R. Bhattacharjee. Sustainable green synthesis of SnO2 quantum dots: A stable, phase-pure and highly efficient photocatalyst for degradation of toxic dyes. Materials Today Sustainability 2024, 26 , 100770. https://doi.org/10.1016/j.mtsust.2024.100770
  16. V. Lavanya, K. Santhakumar. In situ biosynthesis, characterization, and degradation of methylene blue dyes from Solanum Crispum flower extracts. Inorganic Chemistry Communications 2024, 163 , 112353. https://doi.org/10.1016/j.inoche.2024.112353
  17. Shafaq Arif, Durr-e-Kashaf, Kiran Shahzadi, Aneeqa Sabah, M. S. Anwar. Antibacterial and solar-driven photocatalytic activities of CoxSn1−xO2−δ nanoparticles for wastewater treatment. Applied Physics A 2024, 130 (3) https://doi.org/10.1007/s00339-024-07370-5
  18. Elavarasan Appadurai, Anthuvan Babu Stantley, Balu Mahendran Gunasekaran, Senthilkumar Muthiah, Jothi Ramalingam Sivanesan, Noel Nesakumar. Electrochemical sensing of aniline based on bio-inspired cauliflower shaped tin-oxide nanoparticle using Solanum trilobatum leaf extract. Inorganic Chemistry Communications 2024, 161 , 112068. https://doi.org/10.1016/j.inoche.2024.112068
  19. Neetika Kimta, Rajni Dhalaria, Kamil Kuča, Richard Cimler, Vandana Guleria, Shivani Guleria, Harsh Kumar. Production of Metallic Nanoparticles From Agriculture Waste and Their Applications. 2024, 131-156. https://doi.org/10.1007/978-3-031-61133-9_6
  20. R. Sudha Periathai, R. Pon Vengatesh, S. Abarna, N. Prithivikumaran. Treatment of water pollution system using SnO2 nanoparticles synthesized by sol–gel process. Applied Nanoscience 2024, 14 (1) , 135-147. https://doi.org/10.1007/s13204-023-02965-5
  21. Avinash Pratap Gupta, Joystu Dutta. Nanotechnological applications in agro-waste management with special reference to Indian agricultural sector. 2024, 295-310. https://doi.org/10.1016/B978-0-443-18486-4.00005-1
  22. Navpreet Kaur. An innovative outlook on utilization of agro waste in fabrication of functional nanoparticles for industrial and biological applications: A review. Talanta 2024, 267 , 125114. https://doi.org/10.1016/j.talanta.2023.125114
  23. Shivani Chaudhary, Vijay Prakash Jain, Deepa Sharma, Gautam Jaiswar. Implementation of agriculture waste for the synthesis of metal oxide nanoparticles: its management, future opportunities and challenges. Journal of Material Cycles and Waste Management 2023, 25 (6) , 3144-3160. https://doi.org/10.1007/s10163-023-01770-0
  24. Ayyapayya S. Mathad, Nagappa L. Teradal, J. Seetharamappa. Electrochemical Sensor Based on N–Doped Carbon Nanodots Derived from the Agro-Waste Cotton Boll Peel Extracts for Nano Molar Determination of an Anti-Cancer Drug, Pemetrexed. Journal of The Electrochemical Society 2023, 170 (10) , 107506. https://doi.org/10.1149/1945-7111/ad048f
  25. Stephen Sunday Emmanuel, Ademidun Adeola Adesibikan, Oluwaseyi Damilare Saliu. Phytogenically bioengineered metal nanoarchitecture for degradation of refractory dye water pollutants: A pragmatic minireview. Applied Organometallic Chemistry 2023, 37 (2) https://doi.org/10.1002/aoc.6946
  26. Ayyapayya Mathad, Karuna Korgaonkar, Seetharamappa Jaldappagari, Shankara Kalanur. Ultrasensitive Electrochemical Sensor Based on SnO2 Anchored 3D Porous Reduced Graphene Oxide Nanostructure Produced via Sustainable Green Protocol for Subnanomolar Determination of Anti-Diabetic Drug, Repaglinide. Chemosensors 2023, 11 (1) , 50. https://doi.org/10.3390/chemosensors11010050
  27. Supin K K, Anson George, Y. Ranjith Kumar, Thejas K. K., Guruprasad Mandal, Anupama Chanda, M. Vasundhara. Structural, optical and magnetic properties of pure and 3d metal dopant-incorporated SnO 2 nanoparticles. RSC Advances 2022, 12 (41) , 26712-26726. https://doi.org/10.1039/D2RA03691F

ACS Omega

Cite this: ACS Omega 2022, 7, 18, 15423–15438
Click to copy citationCitation copied!
https://doi.org/10.1021/acsomega.1c07099
Published April 26, 2022

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Article Views

3315

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. XRD pattern of sample SNPs: SNPs-200 °C (A), SNPs-500 °C (B), and SNPs-800 °C (C).

    Figure 2

    Figure 2. HR-TEM of sample SNPs-2 at annealing at 200 °C; (A,B) different scale magnifications (20 and 10 nm), (C) inter planar d-spacing, (D) SEAD pattern, (E) EDX micrograph, and (F) particle size distribution of the histogram.

    Figure 3

    Figure 3. HR-TEM of sample SNPs-5 at annealing at 500 °C; (A) 10 nm scale magnification, (B) 20 nm scale magnification, (C) inter planar d-spacing, (D) SEAD pattern, (E) EDX micrograph, and (F) particle size distribution of the histogram.

    Figure 4

    Figure 4. HR-TEM of sample SNPs-8 at annealing at 800 °C; (A) 20 nm scale magnification, (B) 50 nm scale magnification, (C) inter planar d-spacing, (D) SEAD pattern, (E) EDX micrograph, and (F) particle size distribution of the histogram.

    Figure 5

    Figure 5. FT-IR analysis of cotton peel aqueous extract (A), sample SNPs-2 at 200 °C, (B), sample SNPs-5 at 500 °C (C), and sample SNPs-8 at 800 °C (D).

    Figure 6

    Figure 6. UV–vis absorption spectra of SNP sample at 200 °C (A), SNP sample at 500 °C (B), SNP sample at 800 °C (C), band gap energy of SNPs-2 (D), band gap energy of SNPs-5 (E), and band gap energy of SNPs-8 (F).

    Figure 7

    Figure 7. BET analysis of sample SNPs at 200 °C total surface area (A), pore diameter (B), at 500 °C total surface area (C), pore diameter (D), and at 800 °C total surface area (E), pore diameter (F).

    Figure 8

    Figure 8. Photocatalytic activity of SNPs-2 for the degradation of MB dye (A), SNPs-5 for the degradation of MB dye (B), SNPs-8 for the degradation of MB dye (C), SNPs-2 percentage of MB dye degradation (D), SNPs-5 percentage of MB dye degradation (E), and SNPs-8 percentage of MB dye degradation (F).

    Figure 9

    Figure 9. Photocatalytic activity of SNPs-2 for the degradation of MO dye (A), SNPs-5 for the degradation of MO dye (B), SNPs-8 for the degradation of MO dye (C), SNPs-2 percentage of MO dye degradation (D), SNPs-5 percentage of MO dye degradation (E), and SNPs-8 percentage of MO dye degradation (F).

    Figure 10

    Figure 10. Kinetic plots of MB and MO dye degradation, (A) MB degradation kinetics curves for varying catalyst doses (10–30 mg), (B) MO degradation kinetics curves for varying catalyst doses (10–30 mg), (C) effect of dye concentration (5–15 mg/L) on the degradation of MB dye, and (D) effect of dye concentration (5–15 mg/L) on the degradation of MO dye.

    Scheme 1

    Scheme 1. Possible Mechanism of MB and MO Dye Degradation by SNPs Obtained in This Work

    Figure 11

    Figure 11. Recyclability check of biosynthesized sample SNPs-3 for the degradation of MB and MO under identical experimental conditions (A) and XRD pattern after five consecutive cycles for sample SNPs-3 (B).

  • References


    This article references 63 other publications.

    1. 1
      Mourdikoudis, S.; Kostopoulou, A.; LaGrow, A. P. Magnetic Nanoparticle Composites: Synergistic Effects and Applications. Adv. Sci. 2021, 8, 2004951,  DOI: 10.1002/advs.202004951
    2. 2
      Singh, J.; Dutta, T.; Kim, K. H.; Rawat, M.; Samddar, P.; Kumar, P. Green synthesis of metals and their oxide nanoparticles: applications for environmental remediation. J. Nanobiotechnol. 2018, 16, 84,  DOI: 10.1186/s12951-018-0408-4
    3. 3
      Iravani, S. Green synthesis of metal nanoparticles using plants. Green Chem. 2011, 13, 26382650,  DOI: 10.1039/c1gc15386b
    4. 4
      Shamaila, S.; Sajjad, A. K. L.; Ryma, N.-u. -A.; Farooqi, S. A.; Jabeen, N.; Majeed, S.; Farooq, I. Advancements in nanoparticle fabrication by hazard free eco-friendly green routes. Appl. Mater. Today 2016, 5, 150199,  DOI: 10.1016/j.apmt.2016.09.009
    5. 5
      Kumar, K. M.; Mandal, B. K.; Tammina, S. K. Green synthesis of nano platinum using naturally occurring polyphenols. RSC Adv. 2013, 3, 40334039,  DOI: 10.1039/c3ra22959a
    6. 6
      Mazari, S. A.; Ali, E.; Abro, R.; Khan, F. S. A.; Ahmed, I.; Ahmed, M.; Nizamuddin, S.; Siddiqui, T. H.; Hossain, N.; Mubarak, N. M.; Shah, A. Nanomaterials: Applications, waste-handling, environmental toxicities, and future challenges - A review. J. Environ. Chem. Eng. 2021, 9, 105028,  DOI: 10.1016/j.jece.2021.105028
    7. 7
      Palanisamy, G.; Bhuvaneswari, K.; Srinivasan, M.; Vignesh, S.; Elavarasan, N.; Venkatesh, G.; Pazhanivel, T.; Ramasamy, P. Two-dimensional g-C3N4 nanosheets supporting Co3O4-V2O5 nanocomposite for remarkable photodegradation of mixed organic dyes based on a dual Z-scheme photocatalytic system. Diamond Relat. Mater. 2021, 118, 108540,  DOI: 10.1016/j.diamond.2021.108540
    8. 8
      Choi, E.; Lee, D.; Shin, H. J.; Kim, N.; Valladares, L. D. L. S.; Seo, J. Role of oxygen vacancy sites on the temperature-dependent photoluminescence of SnO2 nanowires. J. Phys. Chem. C 2021, 125, 1497414978,  DOI: 10.1021/acs.jpcc.1c02937
    9. 9
      Suthakaran, S.; Dhanapandian, S.; Krishnakumar, N.; Ponpandian, N. Hydrothermal synthesis of SnO2 nanoparticles and its photocatalytic degradation of methyl violet and electrochemical performance. Mater. Res. Express 2019, 6, 0850i3,  DOI: 10.1088/2053-1591/ab29c2
    10. 10
      Nejati-Moghadam, L.; Esmaeili, B. K. A.; Salavati, N. M.; Safardoust, H. Synthesis and characterisation of SnO2 nanostructures prepared by a facile precipitation method. J. Nanostruct. 2015, 5, 4753,  DOI: 10.7508/JNS.2015.01.007
    11. 11
      Zhang, L.; Ren, X.; Luo, Y.; Shi, X.; Asiri, A. M.; Li, T.; Sun, X. Ambient NH3 synthesis via electrochemical reduction of N2 over cubic sub-micron SnO2 particles. Chem. Commun. 2018, 54, 1296612969,  DOI: 10.1039/C8CC06524A
    12. 12
      Haspulat, B.; Sarıbel, M.; Kamış, H. Surfactant assisted hydrothermal synthesis of SnO nanoparticles with enhanced photocatalytic activity. Arab. J. Chem. 2020, 13, 96108,  DOI: 10.1016/j.arabjc.2017.02.004
    13. 13
      Maharajan, S.; Kwon, N. H.; Brodard, P.; Fromm, K. M. A Nano-Rattle SnO2@carbon Composite Anode Material for High-Energy Li-ion Batteries by Melt Diffusion Impregnation. Nanomaterials 2020, 10, 804,  DOI: 10.3390/nano10040804
    14. 14
      Karmaoui, M.; Jorge, A. B.; McMillan, P. F.; Aliev, A. E.; Pullar, R. C.; Labrincha, J. A.; Tobaldi, D. M. One-Step Synthesis, Structure, and Band Gap Properties of SnO2 Nanoparticles Made by a Low Temperature Nonaqueous Sol-Gel Technique. ACS Omega 2018, 3, 1322713238,  DOI: 10.1021/acsomega.8b02122
    15. 15
      Zhu, K.; Lv, Y.; Liu, J.; Wang, W.; Wang, C.; Li, S.; Wang, P.; Zhang, M.; Meng, A.; Li, Z. Facile fabrication of g-C3N4/SnO2 composites and ball milling treatment for enhanced photocatalytic performance. J. Alloys Compd. 2019, 802, 1318,  DOI: 10.1016/j.jallcom.2019.06.193
    16. 16
      Chavali, M. S.; Nikolova, M. P. Metal oxide nanoparticles and their applications in nanotechnology. SN Appl. Sci. 2019, 1, 607,  DOI: 10.1007/s42452-019-0592-3
    17. 17
      Yuliarto, B.; Septiani, N. L. W.; Kaneti, Y. V.; Iqbal, M.; Gumilar, G.; Kim, M.; Na, J.; Wu, K. C.-W.; Yamauchi, Y. Green synthesis of metal oxide nanostructures using naturally occurring compounds for energy, environmental, and bio-related applications. New J. Chem. 2019, 43, 1584615856,  DOI: 10.1039/c9nj03311d
    18. 18
      Diallo, A.; Manikandan, E.; Rajendran, V.; Maaza, M. Physical & enhanced photocatalytic properties of green synthesized SnO2 nanoparticles via Aspalathus linearis. J. Alloys Compd. 2016, 681, 561570,  DOI: 10.1016/j.jallcom.2016.04.200
    19. 19
      Matussin, S.; Harunsani, M. H.; Tan, A. L.; Khan, M. M. Plant-Extract-Mediated SnO2 Nanoparticles: Synthesis and Applications. ACS Sustain. Chem. Eng. 2020, 8, 30403054,  DOI: 10.1021/acssuschemeng.9b06398
    20. 20
      Singh, J.; Kaur, H.; Kukkar, D.; Mukamia, V. K.; Kumar, S.; Rawat, M. Green synthesis of SnO2 NPs for solar light induced photocatalytic applications. Mater. Res. Express 2019, 6, 115007,  DOI: 10.1088/2053-1591/ab4412
    21. 21
      Bhattacharjee, A.; Ahmaruzzaman, M. A novel and green process for the production of tin oxide quantum dots and its application as a photocatalyst for the degradation of dyes from aqueous phase. J. Colloid Interface Sci. 2015, 448, 130139,  DOI: 10.1016/j.jcis.2015.01.083
    22. 22
      Sinha, T.; Ahmaruzzaman, M.; Adhikari, P. P.; Bora, R. Green and Environmentally Sustainable Fabrication of Ag-SnO2 Nanocomposite and Its Multifunctional Efficacy As Photocatalyst and Antibacterial and Antioxidant Agent. ACS Sustain. Chem. Eng. 2017, 5, 46454655,  DOI: 10.1021/acssuschemeng.6b03114
    23. 23
      Sudhaparimala, S.; Vaishnavi, M. Biological synthesis of nano composite SnO2- ZnO - Screening for efficient photocatalytic degradation and antimicrobial activity. Mater. Today Proc. 2016, 3, 23732380,  DOI: 10.1016/j.matpr.2016.04.150
    24. 24
      Li, A. J.; Pal, V. K.; Kannan, K. A review of environmental occurrence, toxicity, biotransformation and biomonitoring of volatile organic compounds. Environ. Chem. Ecotoxicol. 2021, 3, 91116,  DOI: 10.1016/j.enceco.2021.01.001
    25. 25
      Rueda-Marquez, J. J.; Levchuk, I.; Fernández Ibañez, P.; Sillanpää, M. A critical review on application of photocatalysis for toxicity reduction of real wastewaters. J. Clean. Prod. 2020, 258, 120694,  DOI: 10.1016/j.jclepro.2020.120694
    26. 26
      Li, Y.; Chang, H.; Yan, H.; Tian, S.; Jessop, P. G. Reversible Absorption of Volatile Organic Compounds by Switchable-Hydrophilicity Solvents: A Case Study of Toluene with N,N-Dimethylcyclohexylamine. ACS Omega 2020, 6, 253264,  DOI: 10.1021/acsomega.0c04443
    27. 27
      Li, Y.; Liu, K.; Zhang, J.; Yang, J.; Huang, Y.; Tong, Y. Engineering the Band-Edge of Fe2O3/ZnO Nanoplates via Separate Dual Cation Incorporation for Efficient Photocatalytic Performance. Ind. Eng. Chem. Res. 2020, 59, 1886518872,  DOI: 10.1021/acs.iecr.0c03388
    28. 28
      Li, D.; Song, H.; Meng, X.; Shen, T.; Sun, J.; Han, W.; Wang, X. Effects of Particle Size on the Structure and Photocatalytic Performance by Alkali-Treated TiO2. Nanomaterials 2020, 10, 546,  DOI: 10.3390/nano10030546
    29. 29
      Tuan, P. V.; Hieu, L. T.; Tan, V. T.; Phuong, T. T.; Tran Thi Quýnh, H.; Khiem, T. N. The dependence of morphology, structure, and photocatalytic activity of SnO2/rGO nanocomposites on hydrothermal temperature. Mater. Res. Express 2019, 6, 106204,  DOI: 10.1088/2053-1591/ab1e12
    30. 30
      Jiang, Y.; Zhao, H.; Liang, J.; Yue, L.; Li, T.; Luo, Y.; Liu, Q.; Lu, S.; Asiri, A. M.; Gong, Z.; Sun, X. Anodic oxidation for the degradation of organic pollutants: anode materials, operating conditions and mechanisms. A mini review. Electrochem. Commun. 2021, 123, 106912,  DOI: 10.1016/j.elecom.2020.106912
    31. 31
      Li, Y.; Yang, Q.; Wang, Z.; Wang, G.; Zhang, B.; Zhang, Q.; Yang, D. Rapid fabrication of SnO2 nanoparticle photocatalyst: computational understanding and photocatalytic degradation of organic dye. Inorg. Chem. Front. 2018, 5, 30053014,  DOI: 10.1039/C8QI00688A
    32. 32
      Huang, Y.; Li, K.; Lin, Y.; Tong, Y.; Liu, H. Enhanced Efficiency of Electron-Hole Separation in Bi2 O2 CO3 for Photocatalysis via Acid Treatment. ChemCatChem 2018, 10, 19821987,  DOI: 10.1002/cctc.201800101
    33. 33
      Huang, Y.; Lu, Y.; Lin, Y.; Mao, Y.; Ouyang, G.; Liu, H.; Zhang, S.; Tong, Y. Cerium-based hybrid nanorods for synergetic photo-thermocatalytic degradation of organic pollutants. J. Mater. Chem. A 2018, 6, 2474024747,  DOI: 10.1039/c8ta06565a
    34. 34
      Siripireddy, B.; Mandal, B. K. Facile green synthesis of zinc oxide nanoparticles by Eucalyptus globulus and their photocatalytic and antioxidant activity. Adv. Powder Technol. 2017, 28, 785797,  DOI: 10.1016/j.apt.2016.11.026
    35. 35
      Osuntokun, J.; Onwudiwe, D. C.; Ebenso, E. E. Biosynthesis and Photocatalytic Properties of SnO2 Nanoparticles Prepared Using Aqueous Extract of Cauliflower. J. Cluster Sci. 2017, 28, 18831896,  DOI: 10.1007/s10876-017-1188-y
    36. 36
      Ma, C. M.; Hong, G. B.; Lee, S. C. Facile Synthesis of Tin Dioxide Nanoparticles for Photocatalytic Degradation of Congo Red Dye in Aqueous Solution. Catalysts 2020, 10, 792,  DOI: 10.3390/catal10070792
    37. 37
      Arumugam, M.; Manikandan, D. B.; Dhandapani, E.; Sridhar, A.; Balakrishnan, K.; Markandan, M.; Ramasamy, T. Green synthesis of zinc oxide nanoparticles (ZnO NPs) using Syzygium cumini: Potential multifaceted applications on antioxidants, cytotoxic and as nanonutrient for the growth of Sesamum indicum. Environ. Technol. Innovation 2021, 23, 101653,  DOI: 10.1016/j.eti.2021.101653
    38. 38
      Paramarta, V.; Taufik, A.; Saleh, R. Better adsorption capacity of SnO2 nanoparticles with different graphene addition. J. Phys.: Conf. Ser. 2016, 776, 012039,  DOI: 10.1088/1742-6596/776/1/012039
    39. 39
      Haq, S.; Rehman, W.; Waseem, M.; Javed, R.; Mahfooz-ur-Rehman, M.; Shahid, M. Effect of heating on the structural and optical properties of TiO2 nanoparticles: antibacterial activity. Appl. Nanosci. 2018, 8, 1118,  DOI: 10.1007/s13204-018-0647-6
    40. 40
      Ullah, H.; Khan, I.; Yamani, Z. H.; Qurashi, A. Sonochemical-driven ultrafast facile synthesis of SnO2 nanoparticles: Growth mechanism structural electrical and hydrogen gas sensing properties. Ultrason. Sonochem. 2017, 34, 484490,  DOI: 10.1016/j.ultsonch.2016.06.025
    41. 41
      Huang, Y.; Xu, H.; Yang, H.; Lin, Y.; Liu, H.; Tong, Y. Efficient charges separation using advanced BiOI-based hollow spheres decorated with palladium and manganese dioxide nanoparticles. ACS Sustainable Chem. Eng. 2018, 6, 27512757,  DOI: 10.1021/acssuschemeng.7b04435
    42. 42
      Elango, G.; Roopan, S. M. Efficacy of SnO 2 nanoparticles toward photocatalytic degradation of methylene blue dye. J. Photochem. Photobiol., B 2016, 155, 3438,  DOI: 10.1016/j.jphotobiol.2015.12.010
    43. 43
      Yao, X.; Zhang, B.; Cui, S.; Yang, S.; Tang, X. Fabrication of SnSO4-modified TiO2 for enhance degradation performance of methyl orange (MO) and antibacterial activity. Appl. Surf. Sci. 2021, 551, 149419,  DOI: 10.1016/j.apsusc.2021.149419
    44. 44
      Kurniawan, T. A.; Mengting, Z.; Fu, D.; Yeap, S. K.; Othman, M. H. D.; Avtar, R.; Ouyang, T. Functionalizing TiO2 with graphene oxide for enhancing photocatalytic degradation of methylene blue (MB) in contaminated wastewater. J. Environ. Manage. 2020, 270, 110871,  DOI: 10.1016/j.jenvman.2020.110871
    45. 45
      Chen, X.; Wu, Z.; Liu, D.; Gao, Z. Preparation of ZnO photocatalyst for the efficient and rapid photocatalytic degradation of azo dyes. Nanoscale Res. Lett. 2017, 12, 143,  DOI: 10.1186/s11671-017-1904-4
    46. 46
      Santhi, K.; Rani, C.; Karuppuchamy, S. Synthesis and characterisation of a novel SnO/SnO2 hybrid photocatalyst. J. Alloys Compd. 2016, 662, 102107,  DOI: 10.1016/j.jallcom.2015.12.007
    47. 47
      Yousefi, R.; Jamali-Sheini, F.; Cheraghizade, M.; Khosravi-Gandomani, S.; Sáaedi, A.; Huang, N. M.; Basirun, W. J.; Azarang, M. Enhanced visible-light photocatalytic activity of strontium-doped zinc oxide nanoparticles. Mater. Sci. Semicond. Process. 2015, 32, 152159,  DOI: 10.1016/j.mssp.2015.01.013
    48. 48
      Arunkumar, S.; Alagiri, M. Synthesis and Characterization of Spindle-Like TiO2 Nanostructures and Photocatalytic Activity on Methyl Orange and Methyl Blue Dyes Under Sunlight Radiation. J. Cluster Sci. 2017, 28, 26352643,  DOI: 10.1007/s10876-017-1245-6
    49. 49
      Mekewi, M. A.; Darwish, A. S.; Amin, M. S.; Eshaq, G.; Bourazan, H. A. Copper nanoparticles supported onto montmorillonite clays as efficient catalyst for methylene blue dye degradation. Egypt. J. Pet. 2016, 25, 269279,  DOI: 10.1016/j.ejpe.2015.06.011
    50. 50
      Karnan, T.; Selvakumar, S. A. S. Biosynthesis of ZnO nanoparticles using rambutan (Nephelium lappaceumL.) peel extract and their photocatalytic activity on methyl orange dye. J. Mol. Struct. 2016, 1125, 358365,  DOI: 10.1016/j.molstruc.2016.07.029
    51. 51
      Srivastava, N.; Mukhopadhyay, M. Biosynthesis of SnO2 Nanoparticles Using Bacterium Erwinia herbicola and Their Photocatalytic Activity for Degradation of Dyes. Ind. Eng. Chem. Res. 2014, 53, 1397113979,  DOI: 10.1021/ie5020052
    52. 52
      Hirthna; Sendhilnathan, S.; Rajan, P. I.; Adinaveen, T. Synthesis and Characterization of NiFe2O4 Nanoparticles for the Enhancement of Direct Sunlight Photocatalytic Degradation of Methyl Orange. J. Supercond. Novel Magn. 2018, 31, 33153322,  DOI: 10.1007/s10948-018-4601-3
    53. 53
      Radini, I. A.; Hasan, N.; Malik, M. A.; Khan, Z. Biosynthesis of iron nanoparticles using Trigonella foenum-graecum seed extract for photocatalytic methyl orange dye degradation and antibacterial applications. J. Photochem. Photobiol., B 2018, 183, 154163,  DOI: 10.1016/j.jphotobiol.2018.04.014
    54. 54
      Najjar, M.; Hosseini, H. A.; Masoudi, A.; Sabouri, Z.; Mostafapour, A.; Khatami, M.; Darroudi, M. Green chemical approach for the synthesis of SnO2 nanoparticles and its application in photocatalytic degradation of Eriochrome Black T dye. Optik 2021, 242, 167152,  DOI: 10.1016/j.ijleo.2021.167152
    55. 55
      Vidya, C.; Manjunatha, C.; Chandraprabha, M. N.; Rajshekar, M.; Raj M.A.L, A. Hazard free green synthesis of ZnO nano-photo-catalyst using Artocarpus Heterophyllus leaf extract for the degradation of Congo red dye in water treatment applications. J. Environ. Chem. Eng. 2017, 5, 31723180,  DOI: 10.1016/j.jece.2017.05.058
    56. 56
      Zangeneh, H.; Zinatizadeh, A. A. L.; Habibi, M.; Akia, M.; Hasnain Isa, M. Photocatalytic oxidation of organic dyes and pollutants in wastewater using different modified titanium dioxides: A comparative review. J. Ind. Eng. Chem. 2015, 26, 136,  DOI: 10.1016/j.jiec.2014.10.043
    57. 57
      Liu, J.; Zhang, Q.; Tian, X.; Hong, Y.; Nie, Y.; Su, N.; Jin, G.; Zhai, Z.; Fu, C. Highly efficient photocatalytic degradation of oil pollutants by oxygen deficient SnO2 quantum dots for water remediation. Chem. Eng. J. 2021, 404, 127146,  DOI: 10.1016/j.cej.2020.127146
    58. 58
      Li, Li.; Huang, J.; Li, R.; Chen, P.; Chen, D.; Cai, M.; Liu, G.; Feng, Y.; Lv, W.; Liu, G. Synthesis of a carbon dots modified g-C3N4/SnO2 Z-scheme photocatalyst with superior photocatalytic activity for PPCPs degradation under visible light irradiation. J. Hazard. Mater. 2021, 401, 123257,  DOI: 10.1016/j.jhazmat.2020.123257
    59. 59
      Alkaykh, S.; Mbarek, A.; Ali-Shattle, E. E. Photocatalytic degradation of methylene blue dye in aqueous solution by MnTiO3 nanoparticles under sunlight irradiation. Heliyon 2020, 6, e03663  DOI: 10.1016/j.heliyon.2020.e03663
    60. 60
      Wu, K.; Shi, M.; Pan, X.; Zhang, J.; Zhang, X.; Shen, T.; Tian, Y. Decolourization and biodegradation of methylene blue dye by a ligninolytic enzyme-producing Bacillus thuringiensis: degradation products and pathway. Enzyme Microb. Technol. 2022, 156, 109999,  DOI: 10.1016/j.enzmictec.2022.109999
    61. 61
      Abbasi, S.; Hasanpour, M. The effect of pH on the photocatalytic degradation of methyl orange using decorated ZnO nanoparticles with SnO2 nanoparticles. J. Mater. Sci.: Mater. Electron. 2017, 28, 13071314,  DOI: 10.1007/s10854-016-5660-5
    62. 62
      Cai, R.; Zhang, B.; Shi, J.; Li, M.; He, Z. Rapid Photocatalytic Decolorization of Methyl Orange under Visible Light Using VS4/Carbon Powder Nanocomposites. ACS Sustain. Chem. Eng. 2017, 5, 76907699,  DOI: 10.1021/acssuschemeng.7b01137
    63. 63
      Adeel, M.; Saeed, M.; Khan, I.; Muneer, M.; Akram, N. Synthesis and Characterization of Co-ZnO and Evaluation of Its Photocatalytic Activity for Photodegradation of Methyl Orange. ACS Omega 2021, 6, 14261435,  DOI: 10.1021/acsomega.0c05092