Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 4091 2022-12-12 11:17:01 |
2 format Meta information modification 4091 2022-12-14 04:50:06 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Mbarek, W.B.;  Escoda, L.;  Saurina, J.;  Pineda, E.;  Alminderej, F.M.;  Khitouni, M.;  Suñol, J. Nanomaterials as a Sustainable Choice for Treating Wastewater. Encyclopedia. Available online: https://encyclopedia.pub/entry/38594 (accessed on 19 May 2024).
Mbarek WB,  Escoda L,  Saurina J,  Pineda E,  Alminderej FM,  Khitouni M, et al. Nanomaterials as a Sustainable Choice for Treating Wastewater. Encyclopedia. Available at: https://encyclopedia.pub/entry/38594. Accessed May 19, 2024.
Mbarek, Wael Ben, Lluisa Escoda, Joan Saurina, Eloi Pineda, Fahad M. Alminderej, Mohamed Khitouni, Joan-Josep Suñol. "Nanomaterials as a Sustainable Choice for Treating Wastewater" Encyclopedia, https://encyclopedia.pub/entry/38594 (accessed May 19, 2024).
Mbarek, W.B.,  Escoda, L.,  Saurina, J.,  Pineda, E.,  Alminderej, F.M.,  Khitouni, M., & Suñol, J. (2022, December 12). Nanomaterials as a Sustainable Choice for Treating Wastewater. In Encyclopedia. https://encyclopedia.pub/entry/38594
Mbarek, Wael Ben, et al. "Nanomaterials as a Sustainable Choice for Treating Wastewater." Encyclopedia. Web. 12 December, 2022.
Nanomaterials as a Sustainable Choice for Treating Wastewater
Edit

The removal of dyes from textile effluents utilizing advanced wastewater treatment methods with high efficiency and low cost has received substantial attention due to the rise in pollutants in water. For the purpose of treating sewage, the special properties of nanoparticles are being carefully researched. The ability of nanomaterials to remove organic matter, fungus, and viruses from wastewater is another benefit. Nanomaterials are employed in advanced oxidation techniques to clean wastewater. Additionally, because of their small dimensions, nanoparticles have a wide effective area of contact. Due to this, nanoparticles’ adsorption and reactivity are powerful. The improvement of nanomaterial technology will be beneficial for the treatment of wastewater. 

nanocatalysts nanomembranes nanoadsorbents wastewater treatment

1. Nanophotocatalysis

1.1. Nanocatalysis

1.1.1. Nanomaterials as Photocatalysts

Nanocatalysts are attracting a lot of interest as wastewater treatment materials (especially those nanomaterials made of inorganic substances such as semiconductors and metal oxides). For the purpose of the analysis of oxidation of organic pollutants [1] and antimicrobial effects [2], a variety of nanocatalysts are used in wastewater treatment, including photocatalysts [3][4], electrocatalysts [3], heterojunction photocatalytic materials [5], and Fenton-based catalysts [6].
Due to their extensive and efficient photocatalytic activity for diverse contaminants, nanoparticle photocatalytic reactions, which are based on the interaction of light energy with metallic nanoparticles, are of enormous importance. Usually made of semiconductor metals, these photocatalysts can decompose a variety of persistent organic contaminants found in wastewater, including dyes, detergents, insecticides, and volatile organic compounds [7]. Additionally, semiconductor nanocatalysts are highly efficient at degrading both halogenated and nonhalogenated organic chemicals, as well as certain medications, personal care items, and heavy metals [8]. Semiconductor nanomaterials work well, even at low concentrations, and require relatively benign operating conditions. In the same cotext, Baaloudj et al. [9] deduced that the photocatalytic performance of catalysts varies, which was mostly caused by the experimental conditions and, more significantly, by the bandgap, which was crucial to the photocatalytic activity. Depending on the catalyst’s color and the components that it is comprised of, the bandgap varies from one catalyst to another. This distinction can be explained by the process in which a semiconductor can be quickly stimulated by light illumination, which absorbs photons with an energy greater than or equal to that of its bandgap energy, and can then be excited to generate electron–hole pairs (e and h+) [10]. After that, O2 on the surface of the semiconductor will react with the e on the conduction band to form O2* radicals through a reduction process. The hydroperoxyl radical H2O* is then produced by protonation [11]. The trapped electrons, subsequently, combine with these radicals to form hydrogen peroxide H2O2 and hydroxide radicals OH*. The photogenerated valence band holes can react with either water H2O or the hydroxyl ions OH adsorbed on the catalyst surface to produce OH* radicals, which are powerful oxidants and the primary active oxygen species in the photocatalytic destruction of organic contaminants [12]. Accordingly, it is possible that h+ makes up a portion of the OH radicals, aiding in the breakdown of organic contaminants [13].

1.1.2. Nanomaterials as Electrocatalysts

In recent decades, electrocatalysis has undergone significant progress. Numerous fascinating themes in the field of catalysis have arisen as a result of the distinctive characteristics of numerous nanostructures of electrocatalytic materials and their surfaces. Electrocatalysis, a crucial area of catalysis, is a crucial type of catalytic reaction that can transform and store energy through processes involving the transport of electrons. However, because of the extremely complex reaction network, the wide range of reaction selectivity, and the perplexing reaction processes, studying electrocatalysis is extremely difficult. Rare-earth (Gd3+, Nd3+, and Sm3+)-doped cerium oxide has been successfully employed in textile color removal and the decomposition of azo dye RO 107, demonstrating a significant effect on the electrocatalytic activity, according to Rajkumar et al.’s study. The electro-oxidation and electrocatalytic oxidation procedures enable the degradation of high concentration and highly chromoselective dye solutions. The proposed electrochemical degradation process is a successful method of decolorization, according to UV-Vis and FTIR spectral analyses. According to mineralization experiments on RO 107, electrocatalytic oxidation utilizing ceria oxides doped with Gd3+, Nd3+, and Sm3+ increased TOC removal values from 32 to 35.7% after 20 min. The azo bond of the dye structure was the first potentially damaged when the azo bond was attacked, which resulted in the decolorization of the dye. This intermediate was discovered with the GC–MS technique. The intermediates continued to be decomposed into carbon dioxide and water with the aid of the hydroxyl radical and other radicals, which caused the dye solution to become mineralized. When combined with electro-oxidation, cerium-doped Gd3+, Nd3+, and Sm3+ oxides are excellent at removing contaminants from textile dye effluent rapidly. Further research on this method could be performed to find alternative methods for treating wastewater. The findings showed that the cocatalysts for electrocatalytic oxidation processes are Ce0.8Gd0.2O2, Ce0.8Nd0.2O2, and Ce0.8Sm0.2O2. The swift suppression of the electrocharge carriers by the catalyst was the primary electrochemical process responsible for the increased rate. An electron acceptor’s favorable function is the production of additional radicals, which effectively degrade the contaminants through the radical chain branching mechanism. To test this at the industrial scale and with other types of organic effluent, more research is required [14].
Using an electrocatalytic vanadium-doped TiO2 nanocatalyst, Chang et al. examined the degradation of acid red 27 (AR 27). The results showed that AR 27 may be successfully degraded by nano-V/TiO2 electrodes; the greatest color and total organic carbon (TOC) removal efficiencies reached 99% and 76%, respectively, under 0.10 VT (molar ratio of vanadium to titanium) conditions. A high specific surface area nano-V/TiO2 electrode aided in electrocatalytic degradation. This electrocatalytic device performed best at a current density of 25 mA cm2, and as the current density grew, more oxygen was produced. In this electrocatalytic system, the electrical consumption of the nano-V/TiO2 electrode and pure-TiO2 electrode was approximately 0.11 kWh L−1 and 0.02 kWh L−1, respectively. As a result, it can be concluded that the nano-V/TiO2 electrode had both high degradation and energy-saving qualities. The nano-V/TiO2 electrode also demonstrated its potential for repeated use [15].

1.1.3. Heterojunction Photocatalytic Material

The heterojunction photocatalytic material, among the different proposed technologies, has the most potential, since it directly uses solar energy for the creation of valuable chemical fuels (hydrogen, hydrocarbon fuels, etc.) as well as the degradation of hazardous pollutants [16][17][18][19][20][21]. Numerous semiconductors have been researched and produced since Honda and Fujishima’s research on photocatalysis in 1972 [22][23][24][25]. The spatial separation of photogenerated electron–hole pairs was shown to enable the appropriately constructed heterojunction photocatalysts to possess enhanced photocatalytic activity. BiVO4/CeO2 type-II heterojunction photocatalysts for the photocatalytic degradation of methylene blue (MB), methyl orange (MO), and a combination of MB and MO were synthesized hydrothermally, according to Wetchakun et al. [26]. It was discovered that the pH values of 4.56 for BiVO4 and 7.33 for CeO2 corresponded to differing isoelectric points. It has been demonstrated that the difference in isoelectric points between these two semiconductors is advantageous for simultaneously adsorbing cationic and anionic dyes. Particularly, during degrading events, the BiVO4 and CeO2 can each preferentially adsorb cationic MB and anionic MO, respectively. Due to the electrostatic repulsion between the surface charges of the photocatalysts and the charges of the dye molecules, the BiVO4/CeO2 composite had stronger photocatalytic-degradation activity toward the mixture of MB and MO than the individual BiVO4 or CeO2 photocatalysts. The improved electron–hole separation efficiency and potent electrostatic interaction between the composite and the dye molecules were credited for the exceptional activity of the composite photocatalyst. The appropriate coupling of two different semiconductors could both increase the effectiveness of electron–hole separation and provide photocatalysts with good adsorption toward both anionic and cationic dyes [5]. Hu et al.’s study of the linked semiconductor Cu2O/CeO2 photocatalyst’s catalytic activity in the presence of visible light revealed that this heterojunction semiconductor photocatalyst had 20% more photocatalytic degradation of acid orange 7 (AO7) than pure CeO2 [27]. The development of p–n junctions was responsible for the highest photoactivity. Li and Yan [28] examined the photocatalytic degradation of Rhodamine B over a Bi2O3/CeO2 catalyst when it was exposed to visible light. The rhodamine B substrate was totally destroyed within 8 h of irradiation, according to their findings, which showed that Bi2O3/CeO2 in a 9:1 molar ratio gave the greatest photodegradation activity. The improved charge carrier lifetime that was attained by using a composite photocatalyst was connected to the improvement in photocatalytic efficiency. The photocatalytic degradation of rhodamine B over ZnO/CeO2 composite nanofibers was studied by Li et al. [29]. They found that the composite photocatalyst was able to completely degrade the dye substrate within 3 h, while only 17.4% and 82.3% degradation was acquired in the case of pure CeO2 and pure ZnO, respectively. Li et al. studied the photocatalytic degradation of methylene blue (MB) using Bi3TaO7/Ti3C2 heterojunctions, and they deduced a removal efficiency of approximately 99% after 2 h [30]. They reported that synergistic effects between Bi3TaO7 and Ti3C2 improved the photocatalytic performance by enhancing electron–hole pair separation, electronic transmission efficiency, and interfacial charge transfer. They concluded that Ti3C2 serves as an “electronic highway”, isolating the photoelectron–hole pairs and enhancing photocatalytic activity. According to their hypothesized photocatalytic mechanism, photogenerated electrons swiftly migrate on Bi3TaO7 and congregate on Ti3C2 nanosheets. The hydroxyl groups are oxidized by holes gathered on Bi3TaO7 to produce OH*, which is an essential oxidant for dye removal. MB can be oxidized immediately by h+ to produce the tiny non-toxic molecules H2O and CO2 at the same time.
Huang et al. [31] studied the photocatalytic degradation of a Bi2S3/Bi2O3/Bi2O2CO3 nanocomposite. They demonstrated that this manufactured nanocomposite had superior photocatalytic activity in the breakdown of organic contaminants when exposed to visible light. However, under visible light, these Bi2S3/Bi2O3/BOC catalysts could remove 99% of HCHO (500 ppm) in 100 min and were able to remove more than 99% of MO in just 60 min. They claimed that the greater light absorption and effective charge separation were responsible for the composite’s much better performance, and the mechanism behind this high photocatalytic activity showed that superoxide and holes, as opposed to hydroxyl radicals, dominate the photocatalytic process.
According to Low et al., there are at least five basic steps in the photocatalytic process of a semiconductor: (i) the semiconductor’s ability to absorb light, (ii) the generation of photogenerated electron–hole pairs, (iii) their transport and recombination, (iv) the adsorption of reactants and the desorption of products, and (v) the activation of redox reactions on their surface [5].

1.2. Nanomaterial-Based Fenton Catalysis

Using the Fenton reaction to improve the effluent’s biocompatibility or transform the majority of the organic contaminants into low-molecular-mass carboxylic acids and even CO2 is one of the most economically advantageous ways to treat wastewater with low-to-medium levels of total organic carbon. By reducing H2O2 with Fe II, hydroxyl radicals that are extremely aggressive are produced in the Fenton reaction. As an alternative to Fe II, transition metal ions, such as Cu+ and Mn2+, can also aid in the process’s progress [32] There have been numerous reports of homogeneous catalysts being effective for the Fenton reaction [33]. The benefit of heterogeneous catalysts is that they make it simpler to separate material from effluent and eventually reuse it. As a result of this, heterogeneous catalysis are viewed as a natural development of homogeneous catalysis [34]. The creation of heterogeneous catalysts for the Fenton reaction has seen a rise in activity in recent years [32][35]. Recently, a review of the development of clays, silicas, and zeolites for heterogeneous Fenton catalysts was published [32]. Activated carbon has been used by numerous research groups either as a heterogeneous catalyst for the Fenton reaction or in other processes. Additionally, it is capable of supporting metals and metal oxides that have catalytic properties for the Fenton reaction [32]. Reviewing the usage of metal nanoparticles as heterogeneous catalysts for the production of OH radicals from H2O2 is also interesting, given the constant increase in the application of the Fenton reaction for the treatment of several industrial effluents. A new generation of nanoparticle-based heterogeneous catalysts has recently been created for the Fenton reaction. Due to the fact that they can have unique characteristics from the macroscopic or bulk forms of the same material, nanoparticles are important [32]. It is well known that some nanoparticles in catalysis have catalytic characteristics that are missing from bulk materials [36]. Additionally, the size, shape, surface structure, and bulk composition of nanocatalysts all have a significant impact on their activity and selectivity.

2. Nanoadsorbents

Several adsorption methods are used to successfully adsorb dye from polluted waters onto the surface of an adsorbent. It should be mentioned that electrostatic attraction, π–π interactions, van der Waals forces, hydrogen bonds, acid–base reactions, and hydrophobic interactions are the key mechanisms controlling the adsorption of water contaminants on adsorbents. The features of nanoparticles that make them appropriate as nanoadsorbents for sequestration of any cleanup procedure are a high splitting coefficient, chemical and thermal stability in the solvent, chemical inertness, high porosity, being easy to remove from a solution after adsorption, sensitivite and selective towards the target pollutant, being easily regenerable and reusable numerous times, and easy and inexpensive to manafacture.
The primary intrinsic characteristics of nanoadsorbents, such as their fundamental functional groups and surface modification, are being studied in order to improve their capacity to remove hazardous pollutants throughout the process of wastewater treatment. By creating nanocomposites, such as silver/carbon, carbon/titanium oxide, etc., substantial efforts were also made to reduce toxicity. The use of nano-adsorbents in wastewater treatment is the most encouraging method due to their cost-effectiveness, biocompatibility, ease of commercialization, toxic-free method, biodegradability, use of less trained workers, selective separation, ease of recovery, and, most importantly, their high efficacy in removing pollutants.
The following characteristics, including size, shape, surface chemistry, aggregation ability, crystallinity, and chemical reactivity, among others, are crucial in determining how effectively a procedure removes contaminants from an aquatic environment [37].

2.1. Metal Oxide-Based Nanoadsorbents

The inorganic compounds known as metal oxide adsorbents have the distinctive qualities of an increased surface area, high solubility, and reduced production of secondary contaminants. These metal oxide nanoadsorbents can mediate electrostatic interactions due to their charged surfaces, which helps the solute transfer process.
In order to overcome the impacts of fragility, aggregation, and a pressure drop, among other things, a large number of research works on the synthesis of stable nanomaterials and engineering them with the specified functional molecules have been reported. The synthesis of stable nanoceria with amine functionalization was previously reported, and the adsorption procedures used against anionic azo dyes, such as acid yellow 36 and acid yellow 17, were shown to be successful [38][39].
The use of metal oxides to sequester water contaminants has been the subject of numerous studies in recent years [38][40][41]. Their contribution to the environment comes in the form of nanoscale CeO2 or nanoceria, which functions as a photocatalyst for the decomposition of dyes. Their suitability for the absorption of heavy metal ions is determined by the defined surface features of nanoceria and acceptable electrical charge values [42]. Rajarathinam et al. focused on the synthesis and surface functionalization of nanoceria, and explored their adsorption capability for the removal of azo dye Fenalan Yellow G (FYG) taking these properties into consideration. With an adsorbent dosage of 0.1 g for a dye concentration of 10 mg/L of FYG, the maximum removal of 93.62% was seen after 210 min at a pH of 2.0. According to these results, surface-functionalized nanoceria (sf-gNC) can be used as a substitute material for traditional adsorbents in dye-removal procedures [39].
Studies in this topic have been conducted on how heavy metal ions or colors that are pollutants in wastewater degrade. In order to remove malachite green oxalate (MGO) and hexavalent chromium (Cr) from an aqueous solution, Kumar et al. effectively generated metal oxide nanoparticles, such as ZnO and SnO2, using a precipitation technique, with specific surface areas of 15.75 and 24.48 m2/g, respectively. ZnO and SnO2 had 95% and 92% efficiency in decolorizing MGO, respectively. Similarly, Cr adsorbs to ZnO and SnO2, and they removed 95% and 87% of Cr, respectively [43].

2.2. Carbon-Based Nanoadsorbents

Due to the distinctive atomic structure of the carbon atom, carbon materials exhibit a variety of structures and special qualities [44][45]. Carbon materials are classified as zero-dimensional nanomaterials (Buckminster fullerenes and carbon dots), one-dimensional nanomaterials (carbon nanotubes and carbon nanofibers), two-dimensional nanomaterials (graphene), and three-dimensional nanomaterials (carbon sponges) based on their shape, size, and dimensionality [46].
Carbon nanotubes (CNTs) can be found in two-dimensional graphene sheets or three-dimensional nanotubes. They can be divided into two categories based on how many layers or sheets are folded: single-walled CNTs (SWCNTs) and multi-walled CNTs (MWCNTs). SWCNTs play a significant role in clean-up strategies due to their increased surface area, numerous adsorption sites, and other characteristics. The adsorption of a solute on the surface of the adsorbent can be mediated by conventional hydrophobic interactions. As a result, these CNTs are among the adsorbents that have received the most attention in recent years for their capacity to remove a variety of heavy metal ions and organic dyes from wastewater [47].
The maximum adsorption capacity of a given organic pollutant on CNTs was determined by the CNT surface area, surface functional groups on CNTs, the pores in CNT aggregates, surface curvature, and defects of CNT monomers, according to Yang and Xing’s review of the aqueous adsorption of organic pollutants using various carbon nanomaterials [48]. The ability of organic pollutants to adhere to CNTs depends in part on the structures and characteristics of the contaminants. The diverse surfaces of CNTs interact in a variety of ways with distinct molecule configurations. To thoroughly analyze the adsorptive interactions between CNTs and organic pollutants, Chen et al. studied the adsorption of contaminants with different physical-chemical properties to three different types of CNTs. They discovered that while the adsorption affinity did not significantly correlate with hydrophobicity, it did rise in the following order: nonpolar aliphatic, nonpolar aromatic, nitroaromatic, and with the number of nitrofunctional groups within the nitroaromatic group [49]. As a result, the type, quantity, and position of functional groups in organic molecules determine how organic pollutant functional groups affect the adsorption of organic pollutants. The order of the substituted groups at a given position following aniline and phenol’s affinity for adsorption on CNTs is as follows: nitro group > chloride group > methyl group [50].
The adsorption capacities of graphene were much higher than those of other adsorbents under similar conditions, making it a promising adsorbent for the removal of heavy metal ions such as Au(III), Pt(IV) [51], Pb(II) [52], Cu(II) [53], Zn(II) [54], Cd(II) [55], and Co(II) [56]. Graphene is one of the most surprising modern carbon allotropes with distinct properties. The most well-known technique of chemical synthesis is the Hummers method [57], which involves oxidizing graphite to create the two-dimensional oxide form of the compound. With a focus on catalytic and degrading activities, the effectiveness of testing graphene oxide (GO) for the elimination of different dyes was thoroughly examined [58]. TiO2–graphene composites, among a variety of other graphene composites, are frequently employed for the photodegradation of dyes. At the same time, the photocatalytic degradation of dyes, and properties of several metals and metal oxide composites were investigated. The most often studied dye among them for degradation was methylene blue, followed by rhodamine B [58]. Different graphene materials have been used to study dye absorption as well as photocatalytic degradation. Rhodamine B (RB), malachite green (MG), and acriflavine (AF), which are carcinogenic dyes, were applied to highly porous, lightweight graphene oxide foams and high removal capacities were observed of 446, 321 and 228 mg/g, respectively [59]. A simple and affordable lyophilization procedure was used to create the specific 3D GO. Due to the 3D architecture, which is one advantage in terms of its practical use, the foam might be used directly without any preparation, such as in ultrasonication. The excellent antibacterial activity these foams have simultaneously shown against Escherichia coli bacteria in aqueous and nutrient growth conditions further suggests their potential for use in water treatment [59].

2.3. Silica-Based Nanoadsorbents

Sand-like silica is one of the Earth’s crust’s most abundant components. Owing to its special qualities as a lightweight material, silica is a necessity in the production of electrical and communication devices. Silica is utilized in traditional chromatographic methods to separate the desired solute from a complex mixture. One of the most popular uses for silica is the adsorption of pollutants; it has been widely used to remove colors, heavy metals, and other contaminants from drinking water [60]. As a result, numerous research studies on the synthesis of silica have been documented. The ability of a synthesized silica nanoparticle (SSN) to remove dye from single and multicomponent (ternary) systems was reported by Mahmoodi et al. Their findings indicated that the SSN, an environmentally benign adsorbent with a high capacity for cationic dye adsorption, would be a good substitute to remove dyes from multicomponent systems [61].

3. Nanomembranes

Using size exclusion and solution diffusion, nanomembranes, a special type of membrane made of various nanofibers, have been used to remove pollutants based on viruses, inorganic ions, and organic and inorganic nanoparticles from water resources. This method facilitates extremely high elimination rates with condensed fouling propensities, and it also serves as a pretreatment step for reverse osmosis [62]. Numerous studies on membrane nanotechnology have been published in an effort to create multifunctional membranes employing various nanomaterials in various polymer-based membranes [63]. Reverse osmosis, nanofiltration, and other water-treatment methods have used water-porous membranes. A porous support with a composite layer is present in the membrane. The significant composite layer is often composed of a carbon-based material (graphene oxide/CNT) spread in a polymer matrix.
The membranes are commercially available and appropriate for a wide range of uses. However, the effort to create new water resources from sewage calls for membranes with higher productivity and a lower cost related to fouling resistance. Organic substances in water interact with hydrophobic membranes to create membrane fouling. The accumulation of particles on the membrane’s surface or inside its pores is the cause of fouling. Almost all membrane processes experience membrane fouling, which is often brought on by precipitation and particle or molecule deposition on the membrane’s surface or in its pores [64]. Increased membrane separation resistances, decreased productivity, and/or altered membrane selectivity are the effects of membrane fouling [65]. Due to fluctuating product quality and poor recovery, this has an impact on the separation factor for the targeted species in the feed. Pore blockage and solute aggregation, which results in a cake development or a gel layer on the membrane surface, in addition to adsorption, which is exacerbated by concentration polarization and convective forces to and through the membrane, are all typical components of the fouling process. Membrane qualities, such as the material from which the membrane is formed, and feed solution properties, such as composition, concentration, pH, and ionic conditions, are the two main groups of factors that affect membrane fouling. In order to remove suspended solids and condition the feed and membrane surface to reduce the tendency for membrane fouling, pretreatments of feed that have an impact on the feed’s properties in membrane systems are crucial.
A high specific surface area and high porosity with small pores are two distinctive characteristics of electrospun nanofiber membranes. The development of an efficient technology for processing wastewater dyes has been the subject of numerous studies. A variety of techniques have been developed to remove synthetic colors from water and wastewater to reduce their environmental impact. Adsorption on inorganic or organic matrices, color removal via photocatalysis, oxidation, microbiological or enzymatic breakdown, etc., are some of the technologies used [66]. One of the most efficient and financially viable methods for removing textile colours from wastewater is adsorption [67].
The surface area and structure of an adsorbent are its most crucial characteristics. Additionally, the adsorbent surface’s polarity and chemical make-up may affect the attractive forces that bind the adsorbent to the adsorbate [68]. Due to these characteristics, electrospun nanofiber membranes have been utilized to filter out heavy metal ions [69][70] and dye molecules from textile effluent [71]. The sorption potential of electrospun TPU and PVA nanofiber membranes was assessed by Akduman et al. These nanofiber membranes’ large surface areas per unit volume make them perfectly suited for the physical adsorption-based removal of certain materials. The hydrophobic structure of TPU nanofiber membranes, however, resulted in relatively low adsorptions. However, PVA nanofiber membranes, in particular those that were BTCA cross-linked, performed well when it came to the sorption of the dye Reactive Red 141. The highest possible sorption capacity was 88.31 mg/g. The sorption capacity was reduced, nevertheless, as the heat setting temperature rose from 110 °C to 130 °C. The membranes’ volume was also extremely low following the drying process and adsorption process.

References

  1. Ma, H.; Wang, H.; Na, C. Microwave-Assisted Optimization of Platinum-Nickel Nanoalloys for Catalytic Water Treatment. Appl. Catal. B Environ. 2015, 163, 198–204.
  2. Chaturvedi, S.; Dave, P.N.; Shah, N.K. Applications of Nano-Catalyst in New Era. J. Saudi Chem. Soc. 2012, 16, 307–325.
  3. Dutta, A.K.; Maji, S.K.; Adhikary, B. γ-Fe2O3 Nanoparticles: An Easily Recoverable Effective Photo-Catalyst for the Degradation of Rose Bengal and Methylene Blue Dyes in the Waste-Water Treatment Plant. Mater. Res. Bull. 2014, 49, 28–34.
  4. Luo, Y.; Wang, Y.; Hua, F.; Xue, M.; Xie, X.; Xie, Y.; Yu, Y.; Shouhan, Z.; Longshuai, Y.; Zuozhu, X.; et al. Adsorption and photodegradation of reactive red 120 with nickel-iron-layered double hydroxide/biochar composites. J. Hazard. Mater. 2023, 443, 130300.
  5. Low, J.; Yu, J.; Jaroniec, M.; Wageh, S.; Al-Ghamdi, A.A. Heterojunction Photocatalysts. Adv. Mater. 2017, 29, 1601694.
  6. Kurian, M.; Nair, D.S. Heterogeneous Fenton Behavior of Nano Nickel Zinc Ferrite Catalysts in the Degradation of 4-Chlorophenol from Water under Neutral Conditions. J. Water Process Eng. 2015, 8, e37–e49.
  7. Lin, S.T.; Thirumavalavan, M.; Jiang, T.Y.; Lee, J.F. Synthesis of ZnO/Zn Nano Photocatalyst Using Modified Polysaccharides for Photodegradation of Dyes. Carbohydr. Polym. 2014, 105, 1–9.
  8. Adeleye, A.S.; Conway, J.R.; Garner, K.; Huang, Y.; Su, Y.; Keller, A.A. Engineered Nanomaterials for Water Treatment and Remediation: Costs, Benefits, and Applicability. Chem. Eng. J. 2016, 286, 640–662.
  9. Baaloudj, O.; Assadi, I.; Nasrallah, N.; El Jery, A.; Khezami, L.; Assadi, A.A. Simultaneous Removal of Antibiotics and Inactivation of Antibiotic-Resistant Bacteria by Photocatalysis: A Review. J. Water Process Eng. 2021, 42, 102089.
  10. Reyes, C.; Fernandez, J.; Freer, J.; Mondaca, M.A.; Zaror, C.; Malato, S.; Mansilla, H.D. Degradation and inactivation of tetracycline by TiO 2 photocatalysis. J. Photochem. Photobiol. A Chem. 2006, 184, 141–146.
  11. Dong, S.; Cui, L.; Zhang, W.; Xia, L.; Zhou, S.; Russell, C.K.; Fan, M.; Feng, J.; Sun, J. Double-shelled ZnSnO3 hollow cubes for efficient photocatalytic degradation of antibiotic wastewater. Chem. Eng. J. 2020, 384, 123279.
  12. Liu, M.; Zhang, D.; Han, J.; Liu, C.; Ding, Y.; Wang, Z.; Wang, A. Adsorption enhanced photocatalytic degradation sulfadiazine antibiotic using porous carbon nitride nanosheets with carbon vacancies. Chem. Eng. J. 2020, 382, 123017.
  13. Kumar, R.; Barakat, M.A.; Al-Mur, B.A.; Alseroury, F.A.; Eniola, J.O. Photocatalytic degradation of cefoxitin sodium antibiotic using novel BN/CdAl2O4 composite. J. Clean. Prod. 2020, 246, 119076.
  14. Rajkumar, K.; Muthukumar, M.; Mangalaraja, R.V. Electrochemical Degradation of C.I. Reactive Orange 107 Using Gadolinium (Gd3+), Neodymium (Nd3+) and Samarium (Sm3+) Doped Cerium Oxide Nanoparticles. Int. J. Ind. Chem. 2015, 6, 285–295.
  15. Chang, J.H.; Cheng, S.F. The Remediation Performance of a Specific Electrokinetics Integrated with Zero-Valent Metals for Perchloroethylene Contaminated Soils. J. Hazard. Mater. 2006, 131, 153–162.
  16. Xie, G.; Zhang, K.; Guo, B.; Liu, Q.; Fang, L.; Gong, J.R. Graphene-Based Materials for Hydrogen Generation from Light-Driven Water Splitting. Adv. Mater. 2013, 25, 3820–3839.
  17. Kudo, A.; Miseki, Y. Heterogeneous Photocatalyst Materials for Water Splitting. Chem. Soc. Rev. 2008, 38, 253–278.
  18. Wang, P.; Wang, J.; Huo, J.; Xu, W.; Wang, X.; Wang, G.; Wang, P.P.; Wang, J.Q.; Huo, J.T.; Xu, W.; et al. Fast Degradation of Azo Dye by Nanocrystallized Fe-Based Alloys. Sci. China Phys. Mech. Astron. 2017, 60, 076112.
  19. Tong, H.; Ouyang, S.; Bi, Y.; Umezawa, N.; Oshikiri, M.; Ye, J. Nano-Photocatalytic Materials: Possibilities and Challenges. Adv. Mater. 2012, 24, 229–251.
  20. Gaya, U.I.; Abdullah, A.H. Heterogeneous Photocatalytic Degradation of Organic Contaminants over Titanium Dioxide: A Review of Fundamentals, Progress and Problems. J. Photochem. Photobiol. C Photochem. Rev. 2008, 9, 1–12.
  21. Fujishima, A.; Zhang, X.; Tryk, D.A. TiO2 Photocatalysis and Related Surface Phenomena. Surf. Sci. Rep. 2008, 63, 515–582.
  22. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38.
  23. Xuan, J.; Xiao, W.J. Visible-Light Photoredox Catalysis. Angew. Chem. Int. Ed. 2012, 51, 6828–6838.
  24. Tu, W.; Zhou, Y.; Zou, Z. Photocatalytic Conversion of CO2 into Renewable Hydrocarbon Fuels: State-of-the-Art Accomplishment, Challenges, and Prospects. Adv. Mater. 2014, 26, 4607–4626.
  25. Yang, G.; Yan, W.; Zhang, Q.; Shen, S.; Ding, S. One-Dimensional CdS/ZnO Core/Shell Nanofibers via Single-Spinneret Electrospinning: Tunable Morphology and Efficient Photocatalytic Hydrogen Production. Nanoscale 2013, 5, 12432–12439.
  26. Wetchakun, N.; Chaiwichain, S.; Inceesungvorn, B.; Pingmuang, K.; Phanichphant, S.; Minett, A.I.; Chen, J. BiVO4/CeO2 Nanocomposites with High Visible-Light-Induced Photocatalytic Activity. ACS Appl. Mater. Interfaces 2012, 4, 3718–3723.
  27. Hu, S.; Zhou, F.; Wang, L.; Zhang, J. Preparation of Cu2O/CeO2 Heterojunction Photocatalyst for the Degradation of Acid Orange 7 under Visible Light Irradiation. Catal. Commun. 2011, 12, 794–797.
  28. Li, J.; Guo, S.; Zhai, Y.; Wang, E. Nafion–Graphene Nanocomposite Film as Enhanced Sensing Platform for Ultrasensitive Determination of Cadmium. Electrochem. Commun. 2009, 11, 1085–1088.
  29. Li, C.; Chen, R.; Zhang, X.; Shu, S.; Xiong, J.; Zheng, Y.; Dong, W. Electrospinning of CeO2–ZnO Composite Nanofibers and Their Photocatalytic Property. Mater. Lett. 2011, 65, 1327–1330.
  30. Li, K.; Lu, X.; Zhang, Y.; Liu, K.; Huang, Y.; Liu, H. Bi3TaO7/Ti3C2 Heterojunctions for Enhanced Photocatalytic Removal of Water-Borne Contaminants. Environ. Res. 2020, 185, 109409.
  31. Huang, Y.; Fan, W.; Long, B.; Li, H.; Zhao, F.; Liu, Z.; Tong, Y.; Ji, H. Visible light Bi2S3/Bi2O3/Bi2O2CO3 photocatalyst for effective degradation of organic pollutions. Appl. Catal. B Environ. 2016, 185, 68–76.
  32. Navalon, S.; Alvaro, M.; Garcia, H. Heterogeneous Fenton Catalysts Based on Clays, Silicas and Zeolites. Appl. Catal. B Environ. 2010, 99, 1–26.
  33. Baldrian, P. Wood-Inhabiting Ligninolytic Basidiomycetes in Soils: Ecology and Constraints for Applicability in Bioremediation. Fungal Ecol. 2008, 1, 4–12.
  34. Corma, A.; García, H. Lewis Acids: From Conventional Homogeneous to Green Homogeneous and Heterogeneous Catalysis. Chem. Rev. 2003, 103, 4307–4365.
  35. Pera-Titus, M.; García-Molina, V.; Baños, M.A.; Giménez, J.; Esplugas, S. Degradation of Chlorophenols by Means of Advanced Oxidation Processes: A General Review. Appl. Catal. B Environ. 2004, 47, 219–256.
  36. Astruc, D.; Lu, F.; Aranzaes, J.R. Nanoparticles as Recyclable Catalysts: The Frontier between Homogeneous and Heterogeneous Catalysis. Angew. Chem. Int. Ed. 2005, 44, 7852–7872.
  37. Nadeem Baig; Irshad Kammakakam; Wail Falath Nanomaterials: A Review of Synthesis Methods, Properties, Recent Progress, and Challenges. Mater. Adv. 2021, 2, 1821–1871.
  38. Thirunavukkarasu, A.; Muthukumaran, K.; Nithya, R. Adsorption of Acid Yellow 36 onto Green Nanoceria and Amine Functionalized Green Nanoceria: Comparative Studies on Kinetics, Isotherm, Thermodynamics, and Diffusion Analysis. J. Taiwan Inst. Chem. Eng. 2018, 93, 211–225.
  39. Rajarathinam, N.; Arunachalam, T.; Raja, S.; Selvasembian, R. Fenalan Yellow G Adsorption Using Surface-Functionalized Green Nanoceria: An Insight into Mechanism and Statistical Modelling. Environ. Res. 2020, 181, 108920.
  40. Mirjavadi, E.S.; Tehrani, R.M.A.; Khadir, A. Effective Adsorption of Zinc on Magnetic Nanocomposite of Fe3O4/Zeolite/Cellulose Nanofibers: Kinetic, Equilibrium, and Thermodynamic Study. Environ. Sci. Pollut. Res. 2019, 26, 33478–33493.
  41. Rangabhashiyam, S.; Suganya, E.; Lity, A.V.; Selvaraju, N. Equilibrium and Kinetics Studies of Hexavalent Chromium Biosorption on a Novel Green Macroalgae enteromorpha sp. Res. Chem. Intermed. 2015, 42, 1275–1294.
  42. Zach-Maor, A.; Semiat, R.; Shemer, H. Removal of Heavy Metals by Immobilized Magnetite Nano-Particles. New Pub Balaban 2012, 31, 64–70.
  43. Kumar, K.Y.; Muralidhara, H.B.; Nayaka, Y.A.; Balasubramanyam, J.; Hanumanthappa, H. Low-Cost Synthesis of Metal Oxide Nanoparticles and Their Application in Adsorption of Commercial Dye and Heavy Metal Ion in Aqueous Solution. Powder Technol. 2013, 246, 125–136.
  44. Madima, N.; Mishra, S.B.; Inamuddin, I.; Mishra, A.K. Carbon-Based Nanomaterials for Remediation of Organic and Inorganic Pollutants from Wastewater. A Review. Environ. Chem. Lett. 2020, 18, 1169–1191.
  45. Mashkoor, F.; Nasar, A. Inamuddin Carbon Nanotube-Based Adsorbents for the Removal of Dyes from Waters: A Review. Environ. Chem. Lett. 2020, 18, 605–629.
  46. Visakh, P.M. Introduction for Nanomaterials and Nanocomposites: State of Art, New Challenges, and Opportunities. In Nanomaterials and Nanocomposites: Zero to Three-Dimensional Materials and Their Composites; Wiley: Hoboken, NJ, USA, 2016; pp. 1–20.
  47. Ren, X.; Chen, C.; Nagatsu, M.; Wang, X. Carbon Nanotubes as Adsorbents in Environmental Pollution Management: A Review. Chem. Eng. J. 2011, 170, 395–410.
  48. Yang, K.; Xing, B. Adsorption of Organic Compounds by Carbon Nanomaterials in Aqueous Phase: Polanyi Theory and Its Application. Chem. Rev. 2010, 110, 5989–6008.
  49. Chen, W.; Duan, L.; Zhu, D. Adsorption of Polar and Nonpolar Organic Chemicals to Carbon Nanotubes. Environ. Sci. Technol. 2007, 41, 8295–8300.
  50. Yang, K.; Wu, W.; Jing, Q.; Zhu, L. Aqueous Adsorption of Aniline, Phenol, and Their Substitutes by Multi-Walled Carbon Nanotubes. Environ. Sci. Technol. 2008, 42, 7931–7936.
  51. Liu, L.; Liu, S.; Zhang, Q.; Li, C.; Bao, C.; Liu, X.; Xiao, P. Adsorption of Au(III), Pd(II), and Pt(IV) from Aqueous Solution onto Graphene Oxide. J. Chem. Eng. Data 2013, 58, 209–216.
  52. Machida, M.; Mochimaru, T.; Tatsumoto, H. Lead(II) Adsorption onto the Graphene Layer of Carbonaceous Materials in Aqueous Solution. Carbon 2006, 44, 2681–2688.
  53. Yang, S.T.; Chang, Y.; Wang, H.; Liu, G.; Chen, S.; Wang, Y.; Liu, Y.; Cao, A. Folding/Aggregation of Graphene Oxide and Its Application in Cu2+ Removal. J. Colloid Interface Sci. 2010, 351, 122–127.
  54. Wang, H.; Yuan, X.; Wu, Y.; Huang, H.; Zeng, G.; Liu, Y.; Wang, X.; Lin, N.; Qi, Y. Adsorption Characteristics and Behaviors of Graphene Oxide for Zn(II) Removal from Aqueous Solution. Appl. Surf. Sci. 2013, 279, 432–440.
  55. Bian, Y.; Bian, Z.Y.; Zhang, J.X.; Ding, A.Z.; Liu, S.L.; Wang, H. Effect of the Oxygen-Containing Functional Group of Graphene Oxide on the Aqueous Cadmium Ions Removal. Appl. Surf. Sci. 2015, 329, 269–275.
  56. Gao, W.; Majumder, M.; Alemany, L.B.; Narayanan, T.N.; Ibarra, M.A.; Pradhan, B.K.; Ajayan, P.M. Engineered Graphite Oxide Materials for Application in Water Purification. ACS Appl. Mater. Interfaces 2011, 3, 1821–1826.
  57. Lingamdinne, L.P.; Koduru, J.R.; Choi, Y.-L.; Chang, Y.-Y.; Yang, J.-K. Studies on Removal of Pb(II) and Cr(III) Using Graphene Oxide Based Inverse Spinel Nickel Ferrite Nano-Composite as Sorbent. Hydrometallurgy 2016, 165, 64–72.
  58. Upadhyay, R.K.; Soin, N.; Roy, S.S. Role of Graphene/Metal Oxide Composites as Photocatalysts, Adsorbents and Disinfectants in Water Treatment: A Review. RSC Adv. 2014, 4, 3823–3851.
  59. Jayanthi, S.; Krishnarao Eswar, N.; Singh, S.A.; Chatterjee, K.; Madras, G.; Sood, A.K. Macroporous Three-Dimensional Graphene Oxide Foams for Dye Adsorption and Antibacterial Applications. RSC Adv. 2016, 6, 1231–1242.
  60. Idoia Martín-de-Lucía, C.; Campos-Mañas, M.; Agüera, A.; Rodea-Palomares, I.; Pulido-Reyes, G.; Leganés, F.; Fernández-Piñas, F. Roberto Rosal Reverse Trojan-Horse Effect Decreased Wastewater Toxicity in the Presence of Inorganic Nanoparticles. Environ. Sci. Nano 2017, 4, 1273–1282.
  61. Mahmoodi, N.M.; Maghsoodi, A. Kinetics and Isotherm of Cationic Dye Removal from Multicomponent System Using the Synthesized Silica Nanoparticle. Desalination Water Treat. 2014, 54, 562–571.
  62. Jhaveri, J.H.; Murthy, Z.V.P. A Comprehensive Review on Anti-Fouling Nanocomposite Membranes for Pressure Driven Membrane Separation Processes. Desalination 2016, 379, 137–154.
  63. Reddy, A.S.; Kalla, S.; Murthy, Z.V.P. Nano-Particles Enhanced Hydrophobic Membranes: High-Performance Study for Dye Wastewater Treatment Using Membrane Distillation. J. Water Process Eng. 2022, 46, 102610.
  64. Yang, B.; Geng, P.; Chen, G. One-Dimensional Structured IrO2 Nanorods Modified Membrane for Electrochemical Anti-Fouling in Filtration of Oily Wastewater. Sep. Purif. Technol. 2015, 156, 931–941.
  65. Pendergast, M.M.; Hoek, E.M.V. A Review of Water Treatment Membrane Nanotechnologies. Energy Environ. Sci. 2011, 4, 1946–1971.
  66. Forgacs, E.; Cserháti, T.; Oros, G. Removal of Synthetic Dyes from Wastewaters: A Review. Environ. Int. 2004, 30, 953–971.
  67. Singh, K.; Arora, S. Removal of Synthetic Textile Dyes from Wastewaters: A Critical Review on Present Treatment Technologies. Crit. Rev. Environ. Sci. Technol. 2011, 41, 807–878.
  68. Aliabadi, M.; Irani, M.; Ismaeili, J.; Piri, H.; Parnian, M.J. Electrospun Nanofiber Membrane of PEO/Chitosan for the Adsorption of Nickel, Cadmium, Lead and Copper Ions from Aqueous Solution. Chem. Eng. J. 2013, 220, 237–243.
  69. Deng, J.; Kang, X.; Chen, L.; Wang, Y.; Gu, Z.; Lu, Z. A Nanofiber Functionalized with Dithizone by Co-Electrospinning for Lead (II) Adsorption from Aqueous Media. J. Hazard. Mater. 2011, 196, 187–193.
  70. Wu, S.; Li, F.; Wang, H.; Fu, L.; Zhang, B.; Li, G. Effects of Poly (Vinyl Alcohol) (PVA) Content on Preparation of Novel Thiol-Functionalized Mesoporous PVA/SiO2 Composite Nanofiber Membranes and Their Application for Adsorption of Heavy Metal Ions from Aqueous Solution. Polymer 2010, 51, 6203–6211.
  71. Asouhidou, D.D.; Triantafyllidis, K.S.; Lazaridis, N.K.; Matis, K.A. Adsorption of Remazol Red 3BS from Aqueous Solutions Using APTES- and Cyclodextrin-Modified HMS-Type Mesoporous Silicas. Colloids Surfaces A Physicochem. Eng. Asp. 2009, 346, 83–90.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , , ,
View Times: 493
Entry Collection: Wastewater Treatment
Revisions: 2 times (View History)
Update Date: 14 Dec 2022
1000/1000