Previous Article in Journal
A Review of the Structure, Performance, Fabrication, and Impacts of Application Conditions on Wearable Textile GNSS Antennas
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress on Green-Derived Tin Oxide (SnO2) for the Degradation of Textile Dyes: A Review

by
L. M. Mahlaule-Glory
and
N. C. Hintsho-Mbita
*
Department of Chemistry, University of Limpopo, Private Bag X 1106, Sovenga, Polokwane 0727, South Africa
*
Author to whom correspondence should be addressed.
Textiles 2025, 5(3), 36; https://doi.org/10.3390/textiles5030036
Submission received: 1 June 2025 / Revised: 24 July 2025 / Accepted: 30 July 2025 / Published: 19 August 2025

Abstract

Water contamination from textile dyes is a major environmental hazard. This is due to the textile industry serving among the biggest manufacturers, thus the extensive usage of these dyes. Several methods for the treatment of these pollutants have been used; however, they have limitations in terms of cost, forming secondary pollution, and effectiveness. Metal oxides such as tin oxide (SnO2) have been identified as potential photocatalysts for the degradation of these dyes. The potential of SnO2-based photocatalysts, especially those made using green techniques, has been at the forefront of current research. The physical and optical properties, green synthesis techniques, and photocatalytic uses of SnO2 NPs are examined. Furthermore, methods to improve photocatalytic effectiveness through the formation of heterostructures are also explored. Lastly, the conclusion and future perspectives of these materials as suitable candidates for water treatment are highlighted.

1. Introduction

Water pollution poses a major health hazard to both the environment and humanity globally, resulting in water scarcity in many regions [1]. This is because the current economy’s rapid growth is a factor that contributes to water scarcity by causing natural resources to be continuously depleted while industrial pollution is generated [2]. It has been discovered that numerous industries, including the textile industry, examples of which are paints, cosmetics, paper, plastic, leather, and rubber, have been reported as responsible contributors to water pollution, in addition to household effluents [3,4,5].
Dating back to ancient times, natural dyes derived from natural materials such as wood, minerals, flowers, vegetables, and insect secretions have been used in the textile industry. However, due to the over-reliance on this industry, a huge strain has been added to global water [6,7]. High water demands coupled with extreme discharge and pollution caused by dying, bleaching, and printing use water for the application of chemicals and dyes, and in clothing and fabric [8]. The textile industry is the 2nd biggest polluter of water, coming after agriculture. On average, 280 × 103 tons of textile dyes (Figure 1) are leached and released annually [9].
The textile dyes in water streams cause bioaccumulation, thus several countries have added restrictions on their usage [10,11]. For example, the US and European governments, through their ecology and toxicology associations, put limitations on the usage of heavy metals as colorants in the textile industry [12,13,14]. In Asian countries, namely Japan, India, and China, azo dyes have been banned since they contain compounds such as 3,3-dimethylbenzidine [15]. When these dyes (Figure 2) enter water streams, such as lakes and oceans, the food pyramid is affected, algal bloom is produced, and mutagenic effects are also noted on aquatic plants and other microorganisms [11].
A variety of conventional techniques, such as chemical, biological, and physical processes, have been used for the removal of these dyes [15,16]. The physical techniques consist of adsorption, membrane filtration, coagulation–flocculation, and reverse osmosis [17,18,19]. The chemical methods include advanced oxidation processes (AOPs), such as Fenton reactions, photochemical reactions, ultraviolet irradiation, oxidation, ozonation, and electrochemical destruction [4,9]. For the biological techniques, enzymatic degradation and biosorption are the most preferred. Also, other green methods, such as Fenton-based Advanced Oxidation Processes (AOPs), have been used in oxidative/reductive degradation against Mordant Blue Dye [1]. None of the methods can fully remove or degrade these synthetic dyes from wastewater, despite their extensive use as remedies. This is because every strategy has a unique set of drawbacks, and in this case, these traditional treatment methods have been reported to show limitations of high cost, being time-consuming, poor efficiency, recycling challenges, and the production of secondary pollution [9,20,21,22]. Furthermore, chemical approaches have shown an additional constraint of accumulating concentrated sludge, which consequently causes a disposal issue, in addition to these general limitations. [10,23,24]. Thus, researchers have moved towards photocatalysis in the last two decades as it has been shown to be one of the most successful methods of treatment of wastewaters containing dyes because of the low cost, quick oxidation process, and non-toxicity [9,24,25].
For the photocatalysis process to be efficient, a photocatalyst is required to have improved optical and surface characteristics, which include a low band gap, higher surface-volume ratio, and higher surface energy [26]. For the degradation of organic dyes from wastewater, numerous researchers have employed a variety of photocatalytic materials, including metal–organic frameworks (MOFs), carbon-based materials, polymers, and metal oxides [27,28]. However, semiconducting photocatalysts were preferred over these materials [29]. Moreover, these semi-conductive materials, which include metals, metal oxides, and their composites, have emerged as a fascinating field in nanoscience and technology because of their crucial optical, physicochemical, and electrical characteristics [30,31]. These metal oxide semiconductor (MOS) nanoparticles (NPs) include ZnO, TiO2, CuO, CdO, NiO, CeO2, ZrO2, and many others [9,15]. When they break down the organic dyes in wastewater, their intrinsic qualities—such as having a heterogeneous photocatalyst with a broad range of light absorption, redox capabilities, charge dissociation lifespan, sensitivity to light, and stability—also impact their photodegradation efficiency [9]. As a result, tin oxide (SnO2) from amongst the MOSs has drawn a lot of interest lately from researchers due to its exceptional ability to absorb UV radiation, which produces reactive oxygen species (ROS) and increases dye degradation efficiency [32]. Furthermore, SnO2 is an n-type semiconductor with additional features of a rutile crystal structure, a tetragonal form, and a bandgap of roughly 3.6 eV [5,16,33,34,35].
Though this material has been used widely in catalysis, the synthesis of it using conventional techniques such as hydrothermal, solvothermal, chemical vapor deposition, etc., for synthesizing SnO2 has also demonstrated drawbacks such as high operating costs, time consumption, and the use of hazardous chemicals [36,37,38,39,40]. Generating safe and efficient low-cost materials has been at the forefront in circumventing some of these limitations [32,41,42]. This involves the usage of natural reductants from plant extracts with phytochemicals that can serve as reducing and capping agents from natural sources such as agricultural waste (peels, pith, corbs, stalks, and roots), microbes (bacteria, fungi, algae, and enzymes), animals (proteins, vitamins, nutrients (iron, iodine, etc.), as well as plants (leaves, roots, fruits, barks, latex, seeds, and stems) [43,44,45]. Furthermore, plant extracts have drawn a lot of interest in green synthesis since they help create desired nanomaterials in terms of size, shape, and functionality [38].
Thus, in this review, the synthesis of tin oxide via the green route and its composites are investigated for the degradation of textile dyes. Synthesis using plant extracts is highlighted, including the modification and application of these materials as photocatalysts.

2. SnO2 Nanoparticles

Tin is an n-type semiconductor with a bandgap of 3.6 eV and vacant oxygen defects. It has a tetragonal crystal structure with two oxidation states (Sn2+, Sn4+). The tetragonal structure of SnO2 results from Sn ions being in a hexacoordinated state, whereby the O atoms are tri-coordinated. Al-Hamdi et al. [1] reported that SnO2 forms a single cubic structure with a rutile phase. This material has a high carrier density, chemical stability, and oxygen vacancies. It has the potential to possess optical transparency in the visible region, recombination resistance, and chemical stability. It has previously been applied in optoelectronic devices such as gas sensing, dye sensing, and as a photocatalyst [3]. This is caused by its faster charge transport and charge separation size distribution, which leads to its higher electron mobility (100–200 cm2V−1s−1) [4].
The optoelectronic properties this material possesses are mostly influenced by its synthesis conditions, its stoichiometry in relation to O2, and the presence and absence of impurities. Several authors have synthesized SnO2 materials from 0D to 3D in an effort to obtain various optical properties [4]. These included nanowires, nanosheets, nanobelts, etc. From photoluminescence studies, authors were able to obtain information about the defects, structure, and impurities of a material. Lee et al. [13] formed 1D materials of SnO2 via thermal evaporation, and a peak at 595 nm was noted. At room temperature, low SnO2 emissions had been noted. The PL of SnO2 was mainly caused by the energy states found in the bandgap because of defects such as oxygen vacancies and interstitial and dangling bonds. Other authors also noted that upon synthesizing nanoribons and nanorods, it was observed at room temperature that a red light luminescence at 605 nm was observed. This was caused by oxygen vacancies. Nanoblades, via the hydrothermal process at low temperature, were also formed, and from their analysis, a blue shift emission at 445 nm was noted [21]. This could be caused by the triple to ground state transition. From these studies, it can be noted that the growth mechanism can influence the photoluminescence efficiency of these materials.

3. Synthesis of SnO2 Nanoparticles

The bottom-up and top-down methods are the two basic techniques used to produce metallic nanostructures (Figure 3). The objective of top-down techniques, which encompass lithography, ball milling, sputtering, and electrospinning as listed in Table 1, is to generate nanoparticles by progressively segmenting or reducing bulk materials [37]. However, these methods have several drawbacks, as highlighted in Table 1, including high costs, elevated impurity levels, structural defects, and a wide distribution of particle sizes and shapes [38]. In contrast, bottom-up methods are used to synthesize nanomaterials cluster by cluster, molecule by molecule, or atom by atom. Notable techniques in this category include precipitation, atomic layer deposition, chemical vapor deposition, hydrothermal/solvothermal methods, and microwave-assisted heating, as also indicated in Table 1 [24,39]. The bottom-up method is extensively utilized in the synthesis of nanostructures due to its cost-effectiveness, ability to produce a diverse array of morphologies, and narrow particle size distribution (ranging from 1 to 20 nm), which can be modified by varying the starting precursors [40,41,42,43,44].
Table 1. Conventional methods used for the removal of textile pollutants.
Table 1. Conventional methods used for the removal of textile pollutants.
Remediation
Methods
FunctionsAdvantagesDisadvantagesRef.
LithographyTechniques that employed concentrated electron or light beam to synthesize or fabricate structures or patterns at the nanoscale.Commercial fabrication on nano science devices Capacity to create a cluster of the required shape and size from a single nanoparticle. Require high equipment maintenance
Costly equipment
Fluids produce bubbles during synthesis
[37,46,47,48]
Ball millingIt uses a mechanical process to break larger particles into smaller particle sizesProduce ultra-fine nanoparticles
Provides higher yield
Greater densification
Homogeneity
Requires high energy
Irregular morphologies
Potential of contamination.
[49]
SputteringProcess of depositing nanoparticles on a surface using a metal-based material after ions have clashed with the surface.Strong adherence
Homogeneity
Capacity to deposit complex compounds
Low deposition rates
High maintenance
Risk of film stress and contamination
Limited scalability for large-area coatings
[37,49,50]
ElectrospinningIt is an electrostatic method that fabricates fiber materials by pulling a polymer solution under a high electric field, resulting in nanofiber production.Creates fibers with a high surface-to-volume ratio
Adjustable porosity
Uniform structure
High maintenance
Low biocompatibility
Long biodegradation
[51,52]
PrecipitationUsed for producing nanoparticles by combining chemical components in a solution, resulting in the formation of solid precipitates.High purity
Large surface area materials
Simple
Use of toxic reagents and chemicals
High sludge production
Difficulty in separation
[49,53]
HydrothermalA process that creates nanoparticles by carrying out reactions at high temperatures and pressures in an autoclave reactor in a sealed vesselControlled particle size and structure
Surface functionality
Optical properties of the resultant nanomaterials
Uses high temperature
Lengthy response periods
High energy consumption
[40,54]
MicrowaveUses microwave irradiation to heat reaction mixtures that form nanoparticles.
  • Excellent purity
  • Small particle size
  • Produce porous nanoparticles
  • Quick reaction times
  • Scalable
  • Efficient
  • Non-stable particles
  • Particle aggregation

[55]
Chemical vapor depositionUsed to decompose gaseous reactants using light, heat, and plasma to form solid nanoparticlesNo need for capping/reducing agents
High purity nanoparticles
Use of higher temperature and pressure[56]
Spray pyrolysisEntails atomizing metal solutions while maintaining a high temperature and gas flow rate in order to break down the metal and create nanoparticlesRegulated crystallinity, composition, chemical uniformity
Adaptability for large-area films
Difficulty in real-time observation and modelling of nanoparticle formation
High-temperature requirements
Relatively slow deposition rates
[38]
BiologicalA process that uses microbes (bacteria, fungi, plants, and algae) to eliminate pollutantsSimple
Eco-friendly
Cost effective
Slow process
Microbes require extra attention.
Non-selective
Low degradation efficiency
[57,58]
In the hydrothermal method, high temperatures are used, above the boiling point of water, by conducting reactions in a sealed vessel [59]. Due to the significant control they offer over the composition, size, and structure of the resultant nanomaterials, the method is gaining increasing popularity. Although the usage is high, it has certain limitations, such as lengthy response periods that can range from a few minutes to 24 h, which make it difficult to use and lead to considerable energy usage [60]. A notable disadvantage of these reactors is that the temperature is not uniform. Ren et al. [40] synthesized SnO2 NPs using the hydrothermal method (Figure 3). Different materials of nanorods, flowerlike microspheres, and hydrangea microspheres were formed.
In the microwave method, excellent purity, small particle size, and quick reaction times were noted. However, the stability of the nanostructures and particle aggregation were some of the limiting factors reducing the high usage of this synthesis method. Manimaran et al. [31] used the microwave-assisted method to form highly porous tetragonal materials with a particle size of 18 nm. From the PL and zeta potential analysis, improved photoluminescence properties were noted, and these could assist in photodegradation applications. For the chemical vapor deposition method, reactive gas-phase species are used to coat materials in the CVD process. The species in the vapor phase undergo chemical reactions at high temperatures. The process of CVD involves multiple phases [34]. First, the gases are adjusted into various chambers. Thereafter, gaseous species are moved to various locations, which ultimately results in their adsorption. The interaction between the reactants and the substrate leads to the formation of nanostructured materials. Finally, the materials exit the compartment after going through desorption. Unfortunately, the limited growth controllability of this method has restricted its broader application in the production of high-quality nanostructures. Lastly, another popular method, solvothermal synthesis, has been shown to have numerous benefits, including inexpensive precursors, environmental safety, low temperature processing (which improves stoichiometric control), and the capacity to regulate parameters and to control the final product [17,61]. It can be used with other techniques such as hot pressing, microwave, electrochemistry, ultrasound, and optical radiation. Another advantage of solvothermal synthesis is its rapid growth rates, which are made possible by rapid diffusion processes. However, there are several disadvantages to the technique, such as the need for expensive autoclaves, issues about safety during the reaction, and the inability to see the reaction process [23]. Though several methods have been used to produce materials of distinct shapes and properties, they still suffer from certain limitations. For example, the hydrothermal approach tends to have lengthy response periods of synthesis, stability, and particle aggregation, and lastly, elevated growth temperature is still a major limitation [57,62].
Among the bottom-up methods, green synthesis has been on the rise. The are methods which include the use of plant, fungi, and bacterial materials. The phytochemicals contained in these extracts, which include quercetin, polyphenols, and tannins, can be used as reducing and capping agents for these materials; thus, research has mostly moved toward materials formed through this route [28].

4. Green Synthesis of SnO2 Nanoparticles

The synthesis of SnO2 NPs using green materials has been on the rise in the last decade (Table 2). In particular, the usage of plant extracts, as these are readily available and do not require tedious preservation steps and optimization to suit specific conditions, unlike fungi and bacteria [46]. Microorganisms such as fungi generally experience delayed growth rates, resulting in a prolonged synthesis process, and lastly, detaching the formed NPs from the fungi can be difficult. In the usage of plant extracts, phytochemicals consisting of enol groups, for example, phenols and flavonoids, are utilized as reducing agents [47]. In the presence of a methyl group, the extracts can also serve as a capping agent; thus, plants often play a dual role of reducing/capping depending on the phytochemicals present. Moreover, phytochemicals consisting of an ortho-substituted (OH) group can influence the particle size, optical properties, and morphology of the materials [48].
Jeyaraj et al. [13] using Brassica Juncea seed extract, managed to form spherically shaped SnO2 materials with a zeta potential of −40.1 mV and a particle size of 6.6 nm [13]. Suresh et al. [14], in the presence of Delonix elata leaf extract, used three synthesis routes, sonication, wet chemical, and the microwave-assisted method for the green synthesis of SnO2 NPs. From the plant extract, polyhydroxy compounds and quercetin were the phytochemicals of choice, as these consisted of OH groups and could be used as ligation agents. From the analysis, all the materials exhibited O-Sn-O stretches in the 500–650 cm−1 range. Highly crystalline materials were also shown via XRD. From Figure 4a–f, it can also be noted that these materials had a cluster-like foam morphology with limited agglomeration in all the samples, and moreover, EDS also confirmed the formation of these materials with the major elements of Sn and O recorded. From the three materials, the MW-SnO2 recorded the highest surface area of 196 m2g, likely influenced by the lowest particle size of 13 nm. Lastly, the MW-SnO2 was the most thermally stable of the formed materials.
Gonzalez et al. [20] studied the effect of plant concentration (1, 2, 4%) on the synthesis of SnO2 in understanding the optical and structural properties. From FTIR, a characteristic band in the fingerprint region of Sn-O at 601 cm−1 was observed. XRD, through phase identification, noted a tetragonal rutile crystalline phase. Moreover, as the plant concentration increased, the crystallite size also increased to 15, 16, and 21 nm, respectively. This suggests that the Tillia cordata plant assisted in SnO2 crystal growth. From Figure 5, cluster-like quasi-spherically shaped materials were observed. A high agglomeration on the particles was noted. This formed morphology may be caused by low nucleation due to the higher organic content of the matter.
Ahmad et al. [21] used Sambacus Canadensis to form agglomerated clusters of plate-like NPs. A maximum wavelength of 285 nm was obtained, and at 586 cm−1, an asymmetrical vibration of the Sn-O-Sn bond was obtained.
Gomathi et al. [22], using a one-step method, used a kiwi peel extract of Actinidia deliciosa to form spherically shaped small NPs with a particle size of 10 nm or less. Meanwhile, Haritha et al., in the presence of root barks of C. sniposa, synthesized spherically shaped NPs with a particle size of 47 nm. Through GC-MS, metabolites such as 7-hydroxy-6-methoxy-2H-benzopyran were identified to be in abundance, in the range of 61% [23]. Honarmand et al. managed to form tetragonal rutile SnO2 NPs. Through GC-MS, quercetic, a flavonoid derivative, was able to reduce metal ions into metal NPs [24]. Kumar et al. [25], in the presence of guava leaf extract, confirmed the formation of SnO2 through an absorbance peak at 314 nm. The small particle sizes in the quantum confinement range, less than 10 nm, could greatly assist the materials as photocatalysts.

5. Green-Derived SnO2 as Photocatalysts for Dye Degradation

Studies on the photocatalytic degradation of SnO2 (Table 3) or modified SnO2 as a nano or composite catalyst for its effectiveness against different dyes are included in Table 4. The bandgap energies of these SnO2 materials have been reported to range from 2.4 to 4.17 eV, making it an n-type (oxygen-deficient) semiconductor [35,39]. Given that it is an amphoteric white substance with optical transparency and UV photoactivation within the UV range [37,44], it is anticipated to function effectively during the photocalysis process. As a result, it has been utilized against organic dyes in a number of studies gathered in Table 3.
Begum et al. [11] used a Parkia speciosa Hassk plant extract to synthesise SnO2 nanoparticles for the degradation of Acid yellow 23 under sunlight. The material was completely degraded after 24 min at optimum conditions of 20 mg, 5 ppm, and pH = 3. In another study, amino acids were used as a reductant for the synthesis of SnO2, which was used to remove 96% of the MB dye [4].
Honarmand et al. [24] synthesized SnO2 using Jujube fruit extract to degrade MB, and 88% degradation efficiency was recorded after 5 h. For the photocatalytic process, the light used in these reports was either sunlight or lamp light, and others have used the photocatalytic reactor to degrade various organic pollutants. Most of these results indicate that a higher degradation and removal of dyes in longer irradiation times was achieved.
Luque et al. [26] synthesized a green-derived catalyst with three separate bandgaps of 4.02 eV, 3.95 eV, and 3.79 eV. Degradation of 81%, 100% and 100% for MO, MB, and RhB, respectively, was noted [26]. These findings demonstrate that the larger bandgap of 4.02 eV provided a lower percentage removal of MO as compared to MB and RhB, which both had 100% removal. This suggests that the best removal is produced by narrower bandgaps, while wider bandgaps result in poorer removal percentages. In another study, Borah et al. [33] synthesized a green catalyst with a bandgap of 3.80 eV and obtained removal percentages of 94%, 87%, and 93% against MB, MO, and RhB [33]. The results of the two studies were similar because the SnO2 nano catalyst for the photodegradation against three dyes (MB, MO, and RhB) was produced using a natural reductant/stabilizer. The bandgaps in both of these studies were higher than the bulk bandgap of SnO2, which is 3.60 eV, which could be due to the addition of plant extracts during the synthesis process. Therefore, these results demonstrate that MB and RhB are more easily removed by the nano catalyst than the MO dye. In another study, Selvam et al. [45] synthesized SnO2 (CP) using a carica papaya extract and produced a narrower bandgap of 3.09 eV, and a degradation of 86% of RhB using sunlight within 90 min was recorded. Tammina et al. [43] also had the same bandgap of 3.60 eV, and in their study, they removed 96% of the V4BSN dye within 100 min while using a UV lamp (125 Watts) [61]. Kaur et al. [30] degraded 93% of MB and 47% of RhB in 200 min using a 125 watt UV lamp and a 3.45 eV catalyst [30]. Kumar et al. [25], in the presence of sunlight, with a bandgap of 3.64 eV, had a degradation efficiency of 90% against Reactive Yellow dyes in 180 min at a rate of 0.00476 min−1. High reusability of the catalyst was also noted. Singh et al., using a catalyst derived from Punica granatum with a particle size of 20 nm, managed to degrade 92% of MB dye in 240 min in the presence of sunlight. A reaction rate of 6.86 × 10−3 in the degradation of this textile pollutant was recorded.
Ahmad et al. [21] also generated green SnO2 at 350 °C, producing plant-like flower morphological structures and used their catalyst to degrade various dyes. A maximum degradation removal of 2.87% was recorded after 90 min. This suggests that several factors, such as bandgap size, morphology of the catalyst, and synthesis conditions leading to the formation of these materials, may have influenced degradation by preventing the pollutant from passing through the active areas on the surface. Even though in this particular study the activity was limited, it is very rare to obtain degradation rates of less than 5% even for bare green SnO2 materials; hence, several factors may have influenced this activity.
In many studies, spherical morphology materials, irrespective of the particle size, light source, and bandgap, were noted to have high degradation rates [49,50,63]. For example, three experiments that synthesized spherical morphologies at low temperatures were able to deteriorate textile dyes by more than 99%. This was demonstrated by Luque et al. [26], who removed 99% of MB and RhB, Haq et al. [6], who degraded 99% of R6G, and Narasaiah et al. [27], who degraded 98% and 99% of MO and MB, respectively. This is consistent with the findings that morphological structure influences degradation efficiency. In these circumstances, the spherical shape allows light to penetrate the surface equally because there are no structural constraints or corners. This facilitated light absorption on the surface, causing radicals to attack the trapped dye molecules.
Apart from morphology, the majority of the investigations in Table 3 demonstrated greater degradation rates of over 90% with lower bandgap energies (<3 eV) at shorter time frames (<45 min), and higher band energies (above 3) had efficient rates in extended time frames (2–4 h, etc.). Narasaiah et al. [27] managed to degrade 99% of MB in 30 min at 2.68 eV and 98% of MO in 40 min at 2.98 eV under a UV lamp (125 Watts)
Though there are several patterns that have been noted in terms of relating morphology, bandgap, and degradation rates, with these green materials, there are still many factors contributing to the high efficiency or lack thereof. This could be owing to a variety of variables, including the use of optimum degradation conditions and the contribution of a light source. Thus, a demonstration of other contributing elements influencing the efficacy of the degradation of organic dyes with a modified catalyst is noted.

6. Limitations of Green SnO2 and Their Modification for Dye Degradation

Although SnO2 has exhibited a lot of potential, unfortunately, single-component materials have been discovered to possess weaker photocatalytic activity due to their large bandgap, high electron–hole pair recombination rate, and requirement for UV light for activation [51,52].
Furthermore, SnO2 has little to no light absorption in the visible portion of the solar spectrum due to its wide bandgap semiconductor nature [17,36]. Moreover, its photoexcitation is only triggered by UV radiation [9,29]. It has also been reported that SnO2 becomes agglomerated in suspension at high loading and after a few photodegradation cycles, making it rather challenging to extract and recycle from the solution [1]. Therefore, in an attempt to overcome it, researchers often change the surface material or add a lower bandgap MOS to minimize the bandgap [25,27] (Table 4). As noted in Figure 5, outside of the normal limitations faced by these materials, morphology, structural changes, surface area, and particle size play a role. Thus, the literature has indicated that modifying SnO2 characteristics with various doping species or in combination with other semiconductors is a promising approach to overcome this constraint of a wide bandgap and can increase its photocatalytic degradation efficiency [53,54]. The materials that are considered for doping purposes can be from a variety of alkaline earth metals, rare earth elements, and transition metals, in addition to the lower bandgap of other MOSs [26,55].
In a study by Kumar et al. [34], using SnO2 modified with graphene oxide, they managed an 89% degradation of MB and 94% of MO under sunlight for 150 min [38]. Snigdha et al. [35] formed a multicomponent sheet of SnO2/Zn@g-C3N4 using Murraya Paniculata leaves. A 94% degradation of MG dye was achieved with a rate constant of 0.299 min−1. The authors noted that the photoelectrons layered in gC3N4 assisted in the reduction of the recombination rate. Further, h+ and O2 were the main cause of degradation of the MG dye. Heyradi showed that upon loading Ag on SnO2 supported on bentonite in the presence of perovskia abrotanoides plant extract, at optimum conditions of pH 4, enhanced degradation rates of 97% of 100 ppm CV dye were recorded in under 30 min (Figure 6). In this study, on top of depositing the low bandgap metal, the effect of variables such as pH, dosage, concentration, and time had to be investigated to obtain these high degradation rates. This high efficiency may have also been caused by the quick charge separation in the presence of Ag, while plasmonic decay caused by the transfer of electrons and holes also assisted [56,58]. Moreover, the •OH and O2− radicals were noted to be the species responsible for the degradation of this textile dye [37].
Kanwal et al. [42] also used 100 mg of PANI-SnO2 catalyst with a bandgap of 2.68 to degrade 20 ppm of various textile dyes. A 60% or above degradation of all the pollutants was noted. In another study, Minh et al. [59] modified SnO2 with graphitic carbon nitride (gCN) to form a nanocomposite with a reduced bandgap of 2.44 eV, and managed to remove 99% of MB within 120 min using a 50 watt UV lamp [59]. Kaur et al. [60] reported on strontium-doped SnO2 with a bandgap of 2.43 eV for the degradation of MB and CV dyes. In the presence of solar light, a degradation of 83% and 68% was recorded in 150 min.
These studies (Table 4), though they did not report a significant improvement from green bare SnO2 efficiency, due to the extended time, it is known that the presence of carbon in the formation of composites assists in the improvement of the surface area, thus increasing the surface for active sites.
Despite the efficiency of the nanocomposite catalyst against dye degradation and bandgap reduction, the modification of SnO2 with other metals has been reported to show limitations of agglomeration, so other researchers have opted to use carbon-based or polymers for stability and fast recombination rate, as in the previous study [19].
In addition to the intrinsic photocatalytic limitations of SnO2, practical considerations related to scalability, cost, and recyclability must also be acknowledged. Although green synthesis offers a sustainable and environmentally friendly approach for the synthesis of SnO2 nanoparticles, there is still the challenge of scaling up that is encountered in industrial applications. The seasonal availability of raw materials, the varying compositions of phytochemical concentrations, and variability in plant extract composition may result in inconsistent nanoparticle properties, leading to difficulties in large-scale production. Furthermore, while green synthesis reduces the need for hazardous chemicals, the cost of extraction, purification, and potential low yields may offset its economic advantages if not optimized. Another important factor is the recyclability and reusability of green-derived SnO2 photocatalysts.
In a real wastewater setting, agglomeration is an issue since it often leads to a reduced surface area after multiple photocatalytic cycles, resulting in reusability and recovery challenges [64,65]. This poses limitations for continuous flow systems commonly used in industry. Thus, future work should focus on improving the stability of the photocatalyst through forming heterostructures, developing standardized and scalable synthesis processes, and conducting life cycle and cost–benefit assessments to determine the true feasibility of green SnO2 for industrial wastewater remediation.

7. Conclusions and Future Perspectives

This review has outlined the progress made in the application of green-synthesized SnO2 nanoparticles for the photocatalytic degradation of textile dyes. While SnO2 might be promising due to its stability, non-toxicity, and availability, its large bandgap and rapid electron–hole recombination remain significant limitations, thereby restricting its efficiency under visible light. Green synthesis offers an eco-friendly alternative to conventional chemical methods, yet challenges such as agglomeration, inconsistent extract compositions, and limited catalyst recyclability persist. Researchers have attempted to address these issues through doping with low-bandgap metals, forming heterojunctions with other semiconductors, and incorporating carbon-based materials or polymers to enhance stability, surface area, and light absorption. However, further advancements are required to optimize these modifications for consistent and large-scale use.
In the future, more attention should be given to better understanding and controlling processes that are plant-mediated by identifying exactly the specific phytochemicals that participate in the nanoparticle synthesis. This will assist in the scalability as well as the reproducibility in industrial applications. Moreover, efforts should be based on the development of SnO2-based photocatalysts that can utilize the entire solar spectrum, unlike just UV light, and that can be performed through advanced strategies, such as plasmonic enhancement, hybrid composite formation, and defect engineering. Amongst other issues, reusability and stability remain a concern; thus, there is also a need for the development of a recyclable catalyst that can undergo multiple cycles without losing its efficiency. Most importantly, future studies should conduct environmental assessments, ensuring that there are no unknown or unintended potential ecological risks associated with green-derived SnO2.
Furthermore, whilst recent research mainly uses single dye solutions that are controlled in laboratory conditions, real textile wastewater has major challenges. Thus, it is imperative that real effluent testing and pilot-scale studies are conducted to evaluate practical applications. It is also crucial to consider whether green-synthesized SnO2 can surpass or match conventionally synthesized materials in terms of cost, efficiency, and durability. Comparative performance studies will be key to determining their industrial viability. Although green-derived SnO2 poses a significant potential in sustainable wastewater technologies, an extensive multidisciplinary research approach would be beneficial to overcome the outstanding limitation and realize its full application potential.

Author Contributions

Conceptualization, N.C.H.-M.; formal analysis and investigation, L.M.M.-G. and N.C.H.-M.; writing—original draft preparation, L.M.M.-G. and N.C.H.-M.; writing—review and editing, L.M.M.-G. and N.C.H.-M.; supervision and funding acquisition, N.C.H.-M. All authors have read and agreed to the published version of the manuscript.

Funding

The research received no external funding.

Data Availability Statement

No research data was used in this study, a literature analysis was conducted.

Acknowledgments

The authors greatly acknowledge the contribution made by the Department of Chemistry, University of Limpopo.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Al-Hamdi, A.M.; Rinner, U.; Sillanpää, M. Tin dioxide as a photocatalyst for water treatment: A review. Process Saf. Environ. Prot. 2017, 107, 190–205. [Google Scholar] [CrossRef]
  2. Barbosa, M.O.; Moreira, N.F.F.; Ribeiro, A.R.; Pereira, M.F.R.; Silva, A.M.T. Occurrence and removal of organic micropollutants: An overview of the watch list of EU Decision 2015/495. Water Res. 2016, 94, 257–279. [Google Scholar] [CrossRef]
  3. Bhattacharjee, A.; Ahmaruzzaman, M. A green and novel approach for the synthesis of SnO2 nanoparticles and its exploitation as a catalyst in the degradation of methylene blue under solar radiation. Mater. Lett. 2015, 145, 74–78. [Google Scholar] [CrossRef]
  4. Elango, G.; Kumaran, S.M.; Kumar, S.S.; Muthuraja, S.; Roopan, S.M. Green synthesis of SnO2 nanoparticles and its photocatalytic activity of phenolsulfonphthalein dye. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 145, 176–180. [Google Scholar] [CrossRef]
  5. Haq, S.; Rehman, W.; Waseem, M.; Shah, A.; Khan, A.R.; Rehman, M.U.; Ahmad, P.; Khan, B.; Ali, G. Green synthesis and characterization of tin dioxide nanoparticles for photocatalytic and antimicrobial studies. Mater. Res. Express 2020, 7, 025012. [Google Scholar] [CrossRef]
  6. Ligaray, M.; Kim, N.; Park, S.; Park, J.-S.; Park, J.; Kim, Y.; Cho, K.H. Energy projection of the seawater battery desalination system using the reverse osmosis system analysis model. Chem. Eng. J. 2020, 395, 125082. [Google Scholar] [CrossRef]
  7. Paździor, K.; Klepacz-Smółka, A.; Ledakowicz, S.; Sójka-Ledakowicz, J.; Mrozińska, Z.; Żyłła, R. Integration of nanofiltration and biological degradation of textile wastewater containing azo dye. Chemosphere 2009, 75, 250–255. [Google Scholar] [CrossRef]
  8. Pradinaud, C.; Northey, S.; Amor, B.; Bare, J.; Benini, L.; Berger, M.; Boulay, A.-M.; Junqua, G.; Lathuillière, M.J.; Margni, M.; et al. Defining freshwater as a natural resource: A framework linking water use to the area of protection natural resources. Int. J. Life Cycle Assess. 2019, 24, 960–974. [Google Scholar] [CrossRef]
  9. Rashad, M.M.; Ismail, A.A.; Osama, I.; Ibrahim, I.; Kandil, A.-H.T. Photocatalytic decomposition of dyes using ZnO doped SnO2 nanoparticles prepared by sol-vothermal method. Arab. J. Chem. 2014, 7, 71–77. [Google Scholar] [CrossRef]
  10. Tohamy, R.; Ali, S.S.; Li, F.; Okasha, K.M.; Mahmoud, Y.A.-G.; Elsamahy, T.; Jiao, H.; Fu, Y.; Sun, J. A critical review on the dye containing wastewater. Ecotoxilogical and health concerns of textile dyes and possible remediation approaches for environmental safety. Ecotoxicol. Environ. Saf. 2022, 231, 113160. [Google Scholar] [CrossRef] [PubMed]
  11. Begum, S.; Devi, T.B.; Ahmaruzzaman, M. l-lysine monohydrate mediated facile and environment friendly syn-thesis of SnO2 nanoparticles and their prospective applications as a catalyst for the reduction and photodegradation of aro-matic compounds. J. Environ. Chem. Eng. 2016, 4, 2976–2989. [Google Scholar] [CrossRef]
  12. Lee, J.S.; Kwon, O.S.; Jang, J. Facile synthesis of SnO2 nanofibers decorated with N-doped ZnO nanonodules for visible light photocatalysts using single-nozzle co-electrospinning. J. Mater. Chem. 2012, 22, 14565. [Google Scholar] [CrossRef]
  13. Jeyaraj, S.; Saral, A.M. Synthesis of SnO2 Nanoparticles using Brassica juncea (L.) Czern Seed Extract: Characteriza-tion and Biological Applications. Chem. Sel. 2025, 10, e202404490. [Google Scholar] [CrossRef]
  14. Suresh, K.C.; Surendhiran, S.; Manoj Kumar, P.; Kumar, E.R.; Khadar, Y.A.S.; Balamurugan, A. Green synthesis of SnO2 nanoparticles using Delonix elata leaf extract: Evaluation of its structural, optical, morphological and photocatalytic properties. SN Appl. Sci. 2020, 2, 1735. [Google Scholar] [CrossRef]
  15. Singh, J.; Kaur, H.; Kukkar, D.; Mukamia, V.K.; Kumar, S.; Rawat, M. Green synthesis of SnO2 NPs for solar light induced photocalytic applications. Mater. Res. Express 2019, 6, 115007. [Google Scholar] [CrossRef]
  16. Tasisa, Y.E.; Sarma, T.; Sahu, T.K.; Krishnara, R. Phytosynthesis and characterization of tin-oxide nanoparticles (SnO2-NPs) from Croton macrostachyus leaf extract and its application under visible light photocatalytic activities. Sci. Rep. 2024, 14, 10780. [Google Scholar] [CrossRef]
  17. Najjar, M.; Hosseini, H.A.; Masoudi, A.; Hashemzadeh, A.; Darroudi, M. Preparation of tin oxide (IV) nanoparticles by a green chemistry method and investigation of its role in the removal of organic dyes in water purification. Res. Chem. Intermed. 2020, 46, 2155–2168. [Google Scholar] [CrossRef]
  18. Lanjwani, M.F.; Tuzen, M.; Khuhawar, M.Y.; Saleh, T.A. Trends in photocatalytic degradation of organic dye pollutants using nanoparticles: A review. Inorg. Chem. Commun. 2024, 159, 111613. [Google Scholar] [CrossRef]
  19. Buniyamin, I.; Khusaimi, Z.; Rusop, M. Utilization of tin oxide nanoparticles synthesized through plant-mediated methods and their application in photocatalysis: A brief review. Int. J. Chem. 2023, 24, 116–123. [Google Scholar]
  20. González, E.Q.; Medina, E.L.; Robles, R.V.Q.; Gálvez, H.E.G.; Lopez, Y.A.B.; Viveros, E.V.; Molina, F.A.; Nestor, A.R.V.; Morales, P.A.L. A Study of the Optical and Structural Properties of SnO2 Nanoparticles Synthesized with Tilia cordata Applied in Methylene Blue Degradation. Symmetry 2022, 14, 2231. [Google Scholar] [CrossRef]
  21. Ahmad, W.; Pandey, A.; Ahmed, S.; Nur-e-Alam, M. Sambucus Canadensis leaf extract mediated synthesis and analysis of antibacterial and photocatalytic degradation property of stannous oxide nanoparticles. Nanotechnol. Environ. Eng. 2023, 8, 707–716. [Google Scholar] [CrossRef]
  22. Gomathi, E.; Jayapriya, M.; Arulmozhi, M. Environmental benign synthesis of tin oxide (SnO2) nanoparticles using Actinidia deliciosa (Kiwi) peel extract with enhanced catalytic properties. Inorg. Chem. Commun. 2021, 130, 108670. [Google Scholar] [CrossRef]
  23. Haritha, E.; Roopan, S.M.; Madhavi, G.; Elango, G.; Al-Dhabi, N.A.; Arasu, M.V. Green chemical approach towards the synthesis of SnO2 NPs in argument with photocatalytic degradation of diazo dye and its kinetic studies. J. Photochem. Photobiol. B Biol. 2016, 162, 441–447. [Google Scholar] [CrossRef]
  24. Honarmand, M.; Golmohammadi, M.; Naeimi, A. Biosynthesis of tin oxide (SnO2) nanoparticles using jujube fruit for photocatalytic degradation of organic dyes. Adv. Powder Technol. 2019, 30, 1551–1557. [Google Scholar] [CrossRef]
  25. Kumar, M.; Mehta, A.; Mishra, A.; Singh, J.; Rawat, M.; Basu, S. Biosynthesis of tin oxide nanoparticles using Psidium Guajava leave extract for photocatalytic dye degradation under sunlight. Mater. Lett. 2018, 215, 121–124. [Google Scholar] [CrossRef]
  26. Luque, P.A.; Chinchillas-Chinchillas, M.J.; Nava, O.; Lugo-Medina, E.; Martínez-Rosas, M.E.; Carrillo-Castillo, A.; Vilchis-Nestor, A.R.; Madrigal-Muñoz, L.E.; Garrafa-Gálvez, H.E. Green synthesis of tin dioxide nanoparticles using Camellia sinensis and its application in photocatalytic degradation of textile dyes. Optik 2021, 229, 166259. [Google Scholar] [CrossRef]
  27. Narasaiah, B.P.; Banoth, P.; Sohan, A.; Mandal, B.K.; Bustamante Dominguez, A.G.; De Los Santos Valladares, L.; Kollu, P. Green Biosynthesis of Tin Oxide Nanomaterials Mediated by Agro-Waste Cotton Boll Peel Extracts for the Remediation of Environmental Pollutant Dyes. ACS Omega 2022, 7, 15423–15438. [Google Scholar] [CrossRef] [PubMed]
  28. 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, 561–570. [Google Scholar] [CrossRef]
  29. Jafeesh, J.; Aswathy Aromal, S.; Krishnan, R.R. Bio synthesis and characterization of tin oxide for photocatalytic, anti-microbial and anti-oxidant applications. Mater. Chem. Phys. 2023, 307, 128027. [Google Scholar] [CrossRef]
  30. Kaur, M.; Prasher, D.; Sharma, A.; Ghosh, D.; Sharma, R. Natural sunlight driven photocatalytic dye degradation by biogenically synthesized tin oxide (SnO2) nanostructures using Tinospora crispa stem extract and its anticancer and antibacterial applications. Environ. Sci. Pollut. Res. 2023, 30, 38869–38885. [Google Scholar] [CrossRef]
  31. Manimaran, M.; Muthuvel, A.; Said, N.M. Microwave-assisted green synthesis of SnO2 nanoparticles and their photocatalytic degradation and antibacterial activities. Nanotechnol. Environ. Eng. 2023, 8, 413–423. [Google Scholar] [CrossRef]
  32. Al-Enazi, N.M.; Ameen, F.; Alsamhary, K.; Dawoud, T.; Al-Khattaf, F.; AlNadhari, S. Tin oxide nanoparticles (SnO2-NPs) synthesis using Galaxaura elongata and its anti-microbial and cytotoxicity study: A greenery approach. Appl. Nanosci. 2023, 13, 519–527. [Google Scholar] [CrossRef]
  33. Borah, D.; Saikia, P.; Rout, J.; Gogoi, D.; Ghosh, N.N.; Bhattacharjee, C.R. Sustainable green synthesis of SnO2 quantum dots: A stable, phase-pure and highly efficient photocatalyst for degradation of toxic dyes. Mater. Today Sustain. 2024, 26, 100770. [Google Scholar] [CrossRef]
  34. Kumar, J.V.; Karthika, D.; Rosaiah, P.; Devanesan, S.; Mythili, R.; Dhananjaya, M.; Joo, S.W. Fabrication of SnO2/NGO hybrid nanocomposite as an effective photocatalyst for binary dye degradation under sunlight illumination. Process Saf. Environ. Prot. 2024, 188, 398–405. [Google Scholar] [CrossRef]
  35. Snigdha; Gautam, A.; Gautam, N.; Singh, K.B.; Upadhyay, D.D.; Pandey, G. Facile green synthesis of SnO2:Zn@g-C3N4 heterojunction nanocomposite using Murraya paniculata leaves extract for effective photocatalytic degradation of organic dye malachite green by sunlight illumination. J. Mol. Struct. 2024, 1308, 138013. [Google Scholar] [CrossRef]
  36. Mahlaule-Glory, L.M.; Hintsho-Mbita, N.C. Green Derived Zinc Oxide (ZnO) for the Degradation of Dyes from Wastewater and Their Antimicrobial Activity: A Review. Catalysts 2022, 12, 833. [Google Scholar] [CrossRef]
  37. Heyradi, S. Biosynthesis of the Ag Doped SnO2 Nanorods Supported Bentonite Clay with Enhanced Photocatalytic Activity Against Crystal Violet Dye. J. Inorg. Organomet. Polym. Mater. 2025, 35, 2162–2173. [Google Scholar] [CrossRef]
  38. Gebreslassie, Y.T.; Gebretnsae, H.G. Green and Cost-Effective Synthesis of Tin Oxide Nanoparticles: A Review on the Synthesis Methodologies, Mechanism of Formation, and Their Potential Applications. Nanoscale Res. Lett. 2021, 16, 97. [Google Scholar] [CrossRef]
  39. Nipa, S.T.; Akter, R.; Raihan, A.; Rasul, S.b.; Som, U.; Ahmed, S.; Alam, J.; Khan, M.R.; Enzo, S.; Rahman, W. State-of-the-art biosynthesis of tin oxide nanoparticles by chemical precipitation method towards photocatalytic application. Environ. Sci. Pollut. Res. 2022, 29, 10871–10893. [Google Scholar] [CrossRef]
  40. Ren, P.; Qi, L.; You, K.; Shi, Q. Hydrothermal Synthesis of Hierarchical SnO2 Nanostructures for Improved For-maldehyde Gas Sensing. Nanomaterials 2022, 12, 228. [Google Scholar] [CrossRef]
  41. Buniyamin, I.; Asli, N.A.; Akhir, R.M.; Jafar, S.M.; Eswar, K.A.; Mahmood, M.K.A.; Idorus, M.Y.; Shamsudin, M.S.; Rahman, A.F.M.M.; Mahmood, M.R.; et al. Biofabricated SnO2 Nanoparticles Derived from Leaves Extract of Morinda citrifolia and Pandanus amaryllifolius for Photocatalytic Degradation. J. Clust. Sci. 2025, 36, 3. [Google Scholar] [CrossRef]
  42. Kanwal, A.; ur Rehman, M.; Sattar, R.; Thebo, K.H.; Kazi, M. Highly efficient tin oxide and polyaniline-tin-oxide nanostructured materials for photocatalytic degradation of organic dyes. J. Mol. Struct. 2024, 1312, 138454. [Google Scholar] [CrossRef]
  43. Tammina, S.K.; Mandal, B.K. Tyrosine mediated synthesis of SnO2 nanoparticles and their photocatalytic activity towards Violet 4 BSN dye. J. Mol. Liq. 2016, 221, 415–421. [Google Scholar] [CrossRef]
  44. Rathinabala, R.; Thamizselvi, R.; Padmanabhan, D.; Alshahateet, S.F.; Fatimah, I.; Kassegn Sibhatu, A.; Kassegn Weldegebrieal, G.; Izwan Abd Razak, S.; Sagadevan, S. Sun light-assisted enhanced photocatalytic activity and cytotoxicity of green synthesized SnO2 nanoparticles. Inorg. Chem. Commun. 2022, 143, 109783. [Google Scholar] [CrossRef]
  45. Sundara Selvam, P.S.; Ganesan, D.; Rajangam, V.; Raji, A.; Kandan, V. Green Synthesis of SnO2 Nanoparticles for Catalytic Degradation of Rhodamine B. Iran. J. Sci. Technol. Trans. A Sci. 2020, 44, 661–676. [Google Scholar] [CrossRef]
  46. Zalevsky, Z.; Livshit, P.; Gur, E. New Approaches to Image Processing Based Failure Analysis of Nano-Scale ULSI Devices; William Andrew: Norwich, NY, USA, 2014; pp. 1–5. [Google Scholar] [CrossRef]
  47. Pimpin, A.; Srituravanich, W. Reviews on micro- and nanolithography techniques and their applications. Eng. J. 2012, 16, 37–56. [Google Scholar] [CrossRef]
  48. Sharma, E.; Rathi, R.; Misharwal, J.; Sinhmar, B.; Kumari, S.; Dalal, J.; Kumar, A. Evolution in Lithography Techniques: Microlithography to Nanolithography. Nanomaterials 2022, 12, 2754. [Google Scholar] [CrossRef]
  49. Abdullah, H.I.; Al-Amiery, A.A.; Al-Baghdadi, S.B. The using of nanomaterials as catalysts for photodegradations. J. Phys. Conf. Ser. 2021, 185, 012052. [Google Scholar] [CrossRef]
  50. Kour, D.; Kaur, T.; Kumari, S.; Bassi, A.; Khan, S.S.; Gusain, M.; Shoket, H.; Sindhu, M.; Sharma, B.; Singh, S. Advances and innovations in green nanotechnology for agro-environmental sustainability. Nanotechnol. Environ. Eng. 2025, 10, 49. [Google Scholar] [CrossRef]
  51. Nalbandian, M.J. Development and Optimization of Chemically-Active Electro Spun Nanofibers For treatment of Impaired Water Sources. Ph.D. Thesis, University of California Riverside, Riverside, CA, USA, 2014. [Google Scholar]
  52. Zulkifli, M.Z.A.; Nordin, D.; Shaari, N.; Kamarudin, S.K. Overview of Electrospinning for Tissue Engineering Applications. Polymers 2023, 15, 2418. [Google Scholar] [CrossRef]
  53. Mekuye, B.; Abera, B. Nanomaterials: An overview of synthesis, classification, characterization, and applications. Nano Sel. 2023, 4, 486–501. [Google Scholar] [CrossRef]
  54. Robkhob, P.; Tang, I.M.; Thongmee, S. Magnetic properties of the dilute magnetic semiconductor Zn1-xCoxO nanoparticles. J. Supercond. Nov. Magn. 2019, 32, 3637–3645. [Google Scholar] [CrossRef]
  55. De Silva, G. Polarized Light and Optical Activity. Bachelor’s Thesis, University of Colombo, Colombo, Sri Lanka, 2020. [Google Scholar] [CrossRef]
  56. Sabzi, M.; Mousavi Anijdan, S.H.; Shamsodin, M.; Farzam, M.; Hojjati-Najafabadi, A.; Feng, P.; Park, N.; Lee, U. A Review on Sustainable Manufacturing of Ceramic-Based Thin Films by Chemical Vapor Deposition (CVD): Reactions Kinetics and the Deposition Mechanisms. Coatings 2023, 13, 188. [Google Scholar] [CrossRef]
  57. Mahlaule-Glory, L.M.; Mathobela, S.; Hintsho-Mbita, N.C. Biosynthesized Bimetallic (ZnOSnO2) Nanoparticles for Photocatalytic Degradation of Organic Dyes and Pharmaceutical Pollutants. Catalysts 2022, 12, 334. [Google Scholar] [CrossRef]
  58. Álvarez-Chimal, R.; Arenas-Alatorre, J.A. Green Synthesis of Nanoparticles: A Biological Approach. In Advances in Green Chemistry—Prevention-Assurance-Sustainability; Kinjal, J.-S., Ed.; IntechOpen: London, UK, 2023. [Google Scholar] [CrossRef]
  59. Minh, D.T.C.; Do Dat, T.; Quan, T.T.; Nam, N.T.H.; Thi Thanh Huong, Q.; Dat, N.M.; Hieu, N.H. Phyto-synthesis of tin oxide nanoparticles using Diospyros mollis leaves extract doped graphitic carbon nitride for photocatalytic methylene blue degradation and hydrogen peroxide production. Colloids Surf. A Physicochem. Eng. Asp. 2024, 687, 133454. [Google Scholar] [CrossRef]
  60. Kaur, M.; Prasher, D.; Dhiman, V.; Murugesan, P.; Ghosh, D.; Sharma, R. Green energy induced photocatalytic decomposition of methylene blue and crystal violet dyes by strontium doped tin oxide nanoparticles and its antibacterial activity. Phys. B Condens. Matter 2023, 661, 414924. [Google Scholar] [CrossRef]
  61. Sun, C.; Yang, J.; Xu, M.; Cui, Y.; Ren, W.; Zhang, J.; Zhao, H.; Liang, B. Recent intensification strategies of SnO2-based pho-tocatalysts: A review. Chem. Eng. J. 2022, 427, 131564. [Google Scholar] [CrossRef]
  62. Abduh, N.; Al-Adoyni, A.-B. Green Synthesis of ZnO/SnO hybrid nanocomposites for degradation of cationic and an-ionic dyes under sunlight radiation. Materials 2023, 16, 7398. [Google Scholar] [CrossRef]
  63. Joy, J.; Krishnamoorthy, A.; Tanna, A.; Kamathe, V.; Nagar, R.; Srinivasan, S. Recent Developments on the Synthesis of Nanocomposite Materials via Ball Milling Approach for Energy Storage Applications. Appl. Sci. 2022, 12, 9312. [Google Scholar] [CrossRef]
  64. Kamenická, B.; Švec, P.; Weidlich, T. Separation of Anionic Chlorinated Dyes from Polluted Aqueous Streams Using Ionic Liquids and Their Subsequent Recycling. Int. J. Mol. Sci. 2023, 24, 12235. [Google Scholar] [CrossRef] [PubMed]
  65. Kamenická, B.; Weidlich, T. A Comparison of Different Reagents Applicable for Destroying Halogenated Anionic Textile Dye Mordant Blue 9 in Polluted Aqueous Streams. Catalysts 2023, 13, 460. [Google Scholar] [CrossRef]
Figure 1. Different types of textile dyes found in wastewater streams (OA).
Figure 1. Different types of textile dyes found in wastewater streams (OA).
Textiles 05 00036 g001
Figure 2. Ecotoxological impact of dye-containing textile wastewater [11] (OA).
Figure 2. Ecotoxological impact of dye-containing textile wastewater [11] (OA).
Textiles 05 00036 g002
Figure 3. (A) XRD, (B) SEM, and (C) synthesis method of the various SnO2-formed microspheres [40] (OA). In (B), inserts (d) (nanorods with urchin like spheres), (e) (spherical flower like one dimensional nanorods) and ((f) (one dimensional nanorods) represent the magnified view the SEM images of inserts a,b,c respectively).
Figure 3. (A) XRD, (B) SEM, and (C) synthesis method of the various SnO2-formed microspheres [40] (OA). In (B), inserts (d) (nanorods with urchin like spheres), (e) (spherical flower like one dimensional nanorods) and ((f) (one dimensional nanorods) represent the magnified view the SEM images of inserts a,b,c respectively).
Textiles 05 00036 g003
Figure 4. (af) SEM of SnO2 NPs using different synthesis methods and EDS analysis [14] (OA).
Figure 4. (af) SEM of SnO2 NPs using different synthesis methods and EDS analysis [14] (OA).
Textiles 05 00036 g004
Figure 5. SnO2-based photocatalysts and their possible modifications [61].
Figure 5. SnO2-based photocatalysts and their possible modifications [61].
Textiles 05 00036 g005
Figure 6. (a,b) UV-Vis spectra and (c,d) degradation plot of Ag/SnO2 and Ag-SnO2-bent with trapping studies [37] (OA).
Figure 6. (a,b) UV-Vis spectra and (c,d) degradation plot of Ag/SnO2 and Ag-SnO2-bent with trapping studies [37] (OA).
Textiles 05 00036 g006
Table 2. Green synthesis of SnO2 NPs and their characteristics.
Table 2. Green synthesis of SnO2 NPs and their characteristics.
ReductantProduct Calcination Temperature and DurationExtract DosageExtract Temperature and DurationMorphology and Phase IdentificationParticle Size
(nm)
Refs.
Sambucus Canadensis leaf extract350 °C for 30 min20 g in 250 mL60 °C for 12 hPlant like-[21]
Actinidia deliosa (Kiwi) peel extract300 °C for 3 h1:10 ratio60 °C for 2 hSphere5–10 [22]
Catunaregam spinosa (barks roots)450 °C for 2 h300 g in 500 mL60 °C for 2 hSphere47 ± 2[23]
Jujube fruit extract500 °C for 1 h200 g in 200 mL80 °C for 30 min Sphere18 [24]
Psidium Guajava leaf extract400 °C for 4 h-60 °C for 4 hSphere
Tetragonal rutile
8[25]
Camelia sinesis400 °C for 1 h0.5 g in 50 mL
1 g in 50 mL
2 g in 50 mL
60 °C for 1 hQuazi-spherical
Hexagonal rutile
-[26]
Cotton bool waste
Peel extract
200 °C for 3 h
500 °C for 3 h
700 °C for 3 h
3 g in 100 mL80 °C for 30 minSphere
Tetragonal
4
8
13
[27]
Aspalathus linearis200 °C for 3 h
700 °C for 3 h
900 °C for 3 h
--Quasi-spherical1–5[28]
Persia Americana300 °C for 3 h--Sphere
Tetragonal rutile
4[5]
Daphne
mucronata (D. mucronata) leaf extract
Oven dried
100 °C for 6 h
20 g in 500 mL100 °C for 1 hDifferent shapes64[6]
Garcinia Cambogia500 °C for 2 h10 g in 100 mL100 °C for 5 min.Sphere
Tetragonal
14
11
10
[29]
Tnospora crispa (TCSE) stem extract550 °C for 4 h4 g in 150 mL100 °CRods and
spheres
-[30]
Solanum nigrum leaf extract600 °C for
6 h
10 g in 250 mL85 °C in 20 minSphere5 & 18[31]
Galaxaura elongata (red algae).500 °C for 3 h2 g in 200 mL70 °C for 48 hSphere
Tetragonal
35[32]
Croton macrostachyus
leaf extract
500 °C for 3 h10 g in 200 mL60 °C for 4 hSphere32[17]
marine alga, Ulva sp. 400 °C for 3 h2 g in 100 mL70 °C for 20 minSphere and cuboidal5[33]
Aquilaria malaccensis leaf extract800 °C for 2 h80 g in 400 mL80 °C for 20 minSphere
tetragonal
27[20]
Indigofera tinctoria leaf extractNo calcination, oven dried for 8 h at 60 °C5 g in 100 mL60 °C for 30 minSphere and nanosheets
Tetragonal
-[34]
Murraya paniculate (orange jasmine) leave extractOven dried at 60 °C10 g in 100 mL80 °C for 180 minSphere and sheets3–5[35]
Table 3. Photocatalytic degradation of textile dyes using green SnO2.
Table 3. Photocatalytic degradation of textile dyes using green SnO2.
PhotocatalystDyeBandgapLight Source (Watts)Parameters
(Dosage, Conc and pH)
Contact TimeDegradation RemovalReferences
SnO2CBB
MR
CBB
MR
-Solar
Solar
UV lamp
UV lamp
30 mg90 min2.87%
1.94%
2.87%
0.27%
[21]
SnO2MB
MO
RhB
- 5 mg
10−5 M
8 min
8 min
6 min
89%
86%
97%
[22]
SnO2CR-UV Reactor2.5 mg
10−4 M
45 min90%[23]
SnO2MB
EBT
-Solar8 mg
100 ppm
300 min90%
83%
[24]
SnO2RY186-Solar (637 W)10 mg
40 ppm
180 min90%[25]
SnO2MO
MB
RhB
4.02 eV
3.95 eV
3.79 eV
Solar and UV lamp50 mg
15 ppm
180 min81%
100%
100%
[26]
SnO2MB
MO
2.68 eV
2.98 eV
3.14 eV
UV lamp (125 Watts)20 mg
10 ppm
30 min
40 min
99%
98%
[27]
SnO2CR
MB
EY
-UV Lamp2.5 × 10−5
2.7 × 10−6 M
20 min2.2 nm
2.2 nm
19 nm
[28]
SnO2R6G-Solar20 mg
15 ppm
390 min99%[6]
SnO2MB-Solar20 mg
15 ppm
240 min90%[29]
SnO2MB
RhB
3.45 eVSolar50 mg
10 ppm
200 min93%
47%
[30]
SnO2EB3.3 eVSolar2 mg
1.5 × 10−4 M
90 min96%[31]
SnO2RhB
MB
3.03 eV,
2.71 eV,
2.61 eV, and 2.41 eV
UV lamp50 mg
0.1 g/L
pH = 6
80 min91%
96%
[17]
SnO2 QDsMB
MO
RhB
3.80 eVSolar15 mg
1 × 10−4 M
16 min
26 min
44 min
94%
87%
93%
[32]
SnO2MB4.17 eV
3.78 eV
4.0 eV
Solar10 mg
10−4 M
240 min96%[37]
Psidium GuajavaRY3.64 eVSunlight 180 min90 [40]
SnO2MB3.12 eV 70 min94%[41]
SnO2V4BSN3.6 eVUV lamp (125 Watts)10 mg
25 ppm
40 min100%[43]
SnO2RhB-Solar
UV lamp (125 Watts)
50 mg
20 ppm
30 min97%
46%
[44]
SnO2RhB-Solar100 mg
1 × 10−5 M
pH
90 min86%[45]
SnO2EBT-UV lamp (500 W)(2.87 × 10−5 M,
0.211 g L−1
pH = 9
120 min36%[18]
MB = Methylene Blue, MO = Methyl Orange, EBT = Eriochrome Black T, MR = Methylene Red, RhB = Rhodamine Blue, RY186 = Reactive yellow 186, CBB = Coomasie brilliant Blue, CR = Congo Red, EY = Eosin Y, R6G = Rhodamine 6 G, V4BSN = Acid Violet 3.
Table 4. Photocatalytic degradation of SnO2 composites for dye degradation.
Table 4. Photocatalytic degradation of SnO2 composites for dye degradation.
PhotocatalystDyeBandgapLight Source (Watts)Parameters
(Dosage, Conc and pH)
Contact TimeDegradation RemovalReferences
SnO2/NGOMB
MO
-Solar20 mg
20 ppm
150 min89%
94%
[34]
SnO2:Zn@g-C3N4MG
 
 
3.22 eV
 
 
Solar30 mg
pH = 9
10 ppm
100 min94%
 
 
[35]
Ag/SnO2-bentoCV visible1 g/L−130 min97[37]
PANI-SnO2RhB
MO
MB
MR
2.68 eV 100 mg
20 ppm
84%
92%
60%
90%
[42]
SnO2/gCNMB2.44 eVUV lamp
(50 W)
50 mg
10 ppm
120 min99%[59]
Sr:SnO2MB
CV
2.43 eVSolar25 mg
10 ppm
150 min83%
68%
[60]
Fe3O4/SnO2/Pr+MB
RhB
50 W LED 150 min89%
83%
SnO2/ZnOMB UV 150 min88%[57]
ZnO/SnO2MB
MO
RhB
Solar sunlight 55 min
55 min
65 min
100[62]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mahlaule-Glory, L.M.; Hintsho-Mbita, N.C. Recent Progress on Green-Derived Tin Oxide (SnO2) for the Degradation of Textile Dyes: A Review. Textiles 2025, 5, 36. https://doi.org/10.3390/textiles5030036

AMA Style

Mahlaule-Glory LM, Hintsho-Mbita NC. Recent Progress on Green-Derived Tin Oxide (SnO2) for the Degradation of Textile Dyes: A Review. Textiles. 2025; 5(3):36. https://doi.org/10.3390/textiles5030036

Chicago/Turabian Style

Mahlaule-Glory, L. M., and N. C. Hintsho-Mbita. 2025. "Recent Progress on Green-Derived Tin Oxide (SnO2) for the Degradation of Textile Dyes: A Review" Textiles 5, no. 3: 36. https://doi.org/10.3390/textiles5030036

APA Style

Mahlaule-Glory, L. M., & Hintsho-Mbita, N. C. (2025). Recent Progress on Green-Derived Tin Oxide (SnO2) for the Degradation of Textile Dyes: A Review. Textiles, 5(3), 36. https://doi.org/10.3390/textiles5030036

Article Metrics

Back to TopTop