Next Article in Journal
Dynamic Evolutionary Analysis of the Impact of Outward Foreign Direct Investment on Green Innovation Heterogeneity—From the Perspective of Binary Innovation
Previous Article in Journal
Passive Exoskeletons to Enhance Workforce Sustainability: Literature Review and Future Research Agenda
Previous Article in Special Issue
An Integrated Approach to Assess Smart Passive Bioventing as a Sustainable Strategy for the Remediation of a Polluted Site by Persistent Organic Pollutants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photocatalytic Degradation and Adsorptive Removal of Emerging Organic Pesticides Using Metal Oxide and Their Composites: Recent Trends and Future Perspectives

by
Haneen H. Shanaah
1,
Eman F. H. Alzaimoor
1,
Suad Rashdan
1,
Amina A. Abdalhafith
2 and
Ayman H. Kamel
1,3,*
1
Chemistry Department, College of Science, Sakhir 32038, Bahrain
2
Chemistry Department, Faculty of Arts and Sciences, University of Benghazi, Koufra, Benghazi P.O. Box 1308, Libya
3
Department of Chemistry, Faculty of Science, Ain Shams University, Cairo 11566, Egypt
*
Author to whom correspondence should be addressed.
Sustainability 2023, 15(9), 7336; https://0-doi-org.brum.beds.ac.uk/10.3390/su15097336
Submission received: 17 February 2023 / Revised: 10 April 2023 / Accepted: 21 April 2023 / Published: 28 April 2023

Abstract

:
For applications involving water cleanup, metal oxide nanoparticles are exceptionally successful. They are useful for the adsorption and photocatalytic destruction of organic pollutants due to their distinctive qualities, which include their wide surface/volume area, high number of active sites, porous structure, stability, recovery, and low toxicity. Metal oxide nanomaterials have drawn a lot of attention from researchers in the past ten years because of their various production pathways, simplicity in surface modification, abundance, and inexpensive cost. A wide range of metal oxides, such as iron oxides, MgO, TiO2, ZnO, WO3, CuO, Cu2O, metal oxides composites, and graphene–metal oxides composites, with variable structural, crystalline, and morphological features, are reviewed, emphasizing the recent development, challenges, and opportunities for adsorptive removal and photocatalytic degradation of organic pollutants such as dyes, pesticides, phenolic compounds, and so on. In-depth study of the photocatalytic mechanism of metal oxides, their composites, and photocatalytically important characteristics is also covered in this paper. Metal oxides are particularly effective photocatalysts for the degradation of organic pollutants due to their high photodegradation efficiency, economically sound methods for producing photo-catalytic materials, and precise band-gap engineering. Due to their detrimental effects on human health, pesticides—one of the highly hazardous organic pollutants—play a significant part in environmental contamination. Depending on where they come from and who they are targeting, they are categorized in various ways. Researchers focusing on metal oxides and their composites for the adsorptive and photocatalytic degradation of pesticides would find the review to be a beneficial resource. Detailed information on many pesticides, difficulties associated with pesticides, environmental concentration, and the necessity of degradation has been presented.

1. Introduction

Persistent organic chemicals (POPs) are highly hazardous to the ecosystem and living organisms [1]. Their non-biodegradability allows them to accumulate easily in the food chain, affecting both humans and wildlife [2]. Pesticides are one class of POPs with half-lives that can extend to years [3]. They have been used abundantly to control the growth of the crops by exterminating pests including insects, fungi, and microorganisms in agricultural farms [4]. However, the highly toxic nature of these material has become an alarming concern to humans and the environment since they can readily contaminate soil, air, and water through sewage water and industrial and domestic wastes [5]. Additionally, pesticides present in water can reach groundwater through leakages and leaching, surface water, and drinkable water [6,7,8].
The adverse effect of different pesticides on human health have been extensively reported and investigated [9]. For example, organochlorines (OCs) and organophosphates (OPs) accumulate acetylcholine in the central nervous system of humans, leading to serious brain disorder [10]. Carbamates are another class of pesticides that induce apoptosis for human cells and raise the risk of cancer [11]. Other pesticides such as simazine affect humans in trace amounts, leading to kidney failure and heart and lung diseases [12]. Accordingly, numerous attempts have been made to remediate these harmful toxicants from the environmental matrices, utilizing various removal techniques including chemical precipitation, membrane separation, electrocatalysis, advanced oxidation processes, and adsorption [13,14].
The adsorption process has become one of the most researched practices due to its ease of applicability and low cost, depending on the adsorbent [15]. Although adsorption is an easily accessible method for remediation, it generates secondary products and requires further treatment. Accordingly, photocatalytic degradation has attracted the attention of researchers as an alternative technique for the mineralization of toxicants. Photocatalytic degradation employs solar energy, UV, or/and visible light to irradiate a semiconducting material resulting, in the generation of a series of radicals, including hydroxide and superoxide radicals, which are responsible for photodegradation [16]. Studies have shown that photocatalytic activity can be enhanced when the material is adsorbed on the surface of the photocatalyst [17]. Photoactivity, photostability, chemical stability, and the band gap of the photocatalyst should be considered when choosing an adequate material for photocatalytic degradation [18,19]. Photocatalytic degradation can be regarded as a better approach towards the disposing of toxicants; however, it is a more complex process and requires the presence of oxygen as an oxidant [20]. Several studies have corroborated that the higher adsorption capacity of the photocatalyst leads to higher efficiency in degradation, where adsorption facilitates the contact between the pollutant and the photocatalyst. Therefore, a high photodegradation rate requires an effective adsorption process at optimum conditions [21].
Recently, various materials have been investigated for the effective removal of pesticides using adsorption and photocatalytic degradation. Metal oxide nanoparticles, including ferric oxides, cobalt oxides, copper oxides, zinc oxides, titanium oxides, magnesium oxides, cerium oxide, aluminum oxides, and other metal oxides have shown promising removal and degradation results with respect to pollutants such as heavy metals and POPs [22,23,24]. Likewise, they have been widely implemented as nano-adsorbents and photocatalysts for the remediation of pesticides [17]. Their exquisite physical and chemical properties, influenced by their size (1–100 nm), permit them to differ from their corresponding bulk material [25]. Their eminent properties, including large surface/volume area, high number of active sites, porous structure, stability, recovery, and low toxicity, make them valuable nano-adsorbents and photocatalysts [22,26,27].
Developing the structural properties of nano-adsorbents and photocatalysts has become a demand. Therefore, different strategies have been applied to achieve superior performance in physisorption, in chemisorption, and in reducing the band gaps of the photocatalysts to attain the slow recombination of the charge carriers and enhance photo degradation [28,29]. Doping photocatalysts with different metals [30,31], the multi-functionalization of the surface [32], developing heterostructures [33], developing nanocomposites such as metal oxides/metal-organic frameworks [34], metal oxides/polymers [35], and metal/metal oxides [36], are the main strategies followed to manipulate the characteristics of nano-adsorbents and photocatalysts. Many in vitro and in vivo studies have been conducted recently to better understand the toxicological effects and potential risks of various nanoparticle exposures on people and the environment. There is still a major gap in knowledge about the toxic effects of nanoparticle exposure.
Finding an effective remediation process for the harmful effects of pesticides requires immediate. This review provides a detailed collection of recent references to demonstrate the importance of metal oxides and their functionalized nanocomposites in the removal of different types of pesticides using adsorption and photocatalytic degradation.

2. Nanoarchitectures of Metal Oxides and Oxide Perovskites

2.1. Cobalt Oxide

Cobalt oxides are a type of inorganic metal oxides that exists abundantly in nature. Cobalt oxides have many properties that are favorable to environmental applications; they are highly stable, exert no toxicity, exhibit a magnetic behavior, highly resistant to corrosion and oxidation, and have high mechanical strength [37,38,39]. They are p-type semiconducting materials at room temperature that show good conductivity [40]. Cobalt oxides have differences in oxygen vacancies; thus, cobalt oxide exists in many oxidation states [41]. Some of the most used oxidation states include cobalt (II) oxide and cobalt (III) oxide [42]. The structure of Co3O4 NPs consists of a cubic spinel structure with Co (II) at the tetrahedral sites and Co (III) at the octahedral sites [38,43]. Figure 1 shows the structure of Co3O4 spinel structure, with Co (II) surrounded by four oxygen atoms and Co (III), surrounded by six oxygen atoms [44]. Several synthesis methods of cobalt oxide have been reported, including sol–gel, hydrothermal, and microwave-assisted methods. The variable oxidation states of cobalt have made the particles applicable to many fields [45]. Cobalt oxide plays a vital role in many applications, including pollutants sensing, degradation of harmful materials, drug delivery systems, supercapacitors, and storage devices [46,47,48,49,50,51,52,53,54].

2.2. Copper Oxide

Copper oxides exert various properties that make them easily appliable. Numerous structures of copper oxides can be synthesized, such as nanowires, nanorods, nanotubes, and nanoparticles [55]. They are abundant, inexpensive, and highly stable, with high catalytic and antibacterial activity [56]. Different oxidation states of copper oxides are available; however, cuprous oxide (Cu2O) and cupric oxide (CuO) are the most stable, while paramelaconite (Cu4O3) is metastable. Figure 2 shows the three crystal structures for copper oxides [57]. Cuprous oxide (Cu2O) crystallizes in a cubic structure and is a p-type semiconductor. On the other hand, cupric oxide (CuO) exists in a monoclinic crystal structure and is believed to be both a p- and an n-type semiconductor [58,59]. Figure 3 illustrates the difference between a p-type and an n-type semiconductors.
Copper oxide semiconductors are recognized for their role in the remediation of environmental contaminants due to their strong oxidation and reduction ability and environmental compatibility [60]. A difference between cobalt oxides and copper oxides is that cobalt oxides have higher stability at high temperatures. However, the precursor salts of cobalt oxides are more expensive than the precursor salts of copper oxide, making copper oxides more easily obtained and applied in many fields [61]. Copper oxides have many prospective applications in the fields of sensors, antimicrobial activity, catalysis, coatings, polymers, and electronics [62,63,64,65,66]. Moreover, cupric oxide shows promising results for the decontamination of water since it is relatively cheap, with high catalytic activity [67,68]. Many routes for the preparation of copper oxides are available. Chemical precipitation, the sonochemical method, hydrothermal synthesis, and synthesis via plant extracts and micro-organisms, are all examples for chemical and biological methods for the synthesis of copper oxide NPs [69,70].

2.3. Zinc Oxide

Zinc oxide is the second most abundant metal with different morphological shapes and sizes [71]. It is an inorganic multifunctional material that has favorable properties and characteristics [72]. Zinc oxide exhibits optical and piezoelectric properties that are attractive to important fields such as optoelectronics and transparent electronics [73]. It also offers a high surface area with a wide band gap of 3.37 eV, making it appropriate for photocatalysis [74]. Zinc oxides are widely used for medicinal applications since zinc is considered a dietary supplement, in addition to their antibacterial activity [75,76,77]. It has also been reported that the antibacterial activity of zinc oxides depends on their size and shape, which can be controlled through the synthesis route [78]. Zinc oxides can be synthesized by chemical, physical, and biological methods. These methods include microemulsion, precipitation, plasma and ultrasonic techniques, the sol–gel method, combustion, hydrothermal synthesis, and green synthesis from plants extracts [79,80,81]. Moreover, zinc oxides have gained attention for water purification and the removal of hazardous materials since they are biocompatible [82]. Polymorphs of zinc oxide consist of three phases shown in Figure 4: hexagonal wurtzite, cubic zinc blende, and cubic rock salt [83]. The wurtzite phase is the most thermodynamically stable, with every zinc atom tetrahedrally coordinated with four oxygen atoms. Zinc blende and rock salt are metastable [84,85]. More zinc oxides applications include solar cells [86,87,88], sensors [89,90], drug delivery [91,92], and the cosmetic industry [93].

2.4. Iron Oxide

Iron oxides have several forms which consist mainly of iron and oxygen. These forms include iron (II) oxide (wüstite, FeO), iron (II,III) oxide (magnetite, Fe3O4), and iron (III) oxide (ferric oxide, Fe2O3) [95], and are shown in Figure 5 [96]. Ferric oxide is the most common form and it has four polymorphs [97]: alpha phase hematite (α-Fe2O3), beta phase (β-Fe2O3), gamma phase maghemite (γ-Fe2O3), and epsilon phase (ε-Fe2O3) [98].
The magnetic properties of iron oxide NPs are highly affected by their size, dispersion, and surface [99]. Magnetite and maghemite are intrinsic ferrimagnetic materials, while hematite’s magnetic properties are thermally induced [100]. The main properties that allow IONPs to be of such interest in many fields is their superparamagnetic behavior, their high surface/volume ratio, non-toxicity, reusability, biocompatibility, high stability, and resistance to change [101,102].
IONPs are also widely utilized in biomedical applications such as drug delivery systems, where the particles can be carried to a specific site with high accuracy due to the use of an external magnetic field to direct the particles [103,104]. Iron oxide NPs are usefully applied in many fields including sensing [105,106,107], catalysis [108,109,110], photodegradation [111,112], and adsorption of pollutants [113,114,115]. The only disadvantage in IONPs is the aggregation of particles in aqueous media, which is unfavorable in water remediation applications. Therefore, IONPs can be further stabilized by surface modification, including coating with surfactants and polymers [116,117]. For example, a study coated iron oxide NPs with chitosan (made from chitin), which stabilized the particles and further functionalized them with amine groups, thus increasing the number of binding sites available [104].

2.5. Titanium Oxide

Titania or titanium dioxide (TiO2) is abundant in nature and has favorable advantages with respect to energy and environmental applications [118,119]. It has been shown to be promising for these applications due to its chemical stability, biological and chemical inertness, and non-toxicity [120]. TiO2 has long durability and transparency to visible light. It is active under UV light and functions as a semiconductor with a band gap around 3.2 eV [121]. Furthermore, titanium oxide exists in three crystalline forms, which are tetragonal anatase, tetragonal rutile, and orthorhombic brookite, with rutile being the most thermodynamically stable form [122]. Figure 6 shows the three polymorphs of titania [123]. Even though rutile is the most stable, anatase is more efficient when it comes to photocatalysis. However, some studies showed that rutile can possess good photocatalytic activity [124,125]. Anatase and rutile have tetragonal symmetry, while brookite has an orthorhombic crystalline structure [126]. The method of preparation for TiO2 nanomaterials controls their morphology; therefore, their performance in applications can be enhanced [127]. Tailoring particle size and crystal surfaces determines which facets are exposed in TiO2, affecting its photocatalytic activity tremendously. One of the most common methods for morphology control is the use of organic surfactants [128]. Moreover, a study synthesized different morphologies of titania by controlling the temperature in the solvothermal method [129]. Rose-like, chrysanthemum-like, and sea-urchin-like TiO2 nanostructures (shown in Figure 7) were successfully prepared and applied for the photocatalytic degradation of Rhodamine B, where each nanostructure had a different photocatalytic performance. The applications for titania are wide, including the removal of organic pollutants [130], medical applications [131], energy storage, and sustainable energy production [132,133].

2.6. Magnesium Oxide

Magnesium oxide (MgO) is a multifunctional inorganic material that holds great technological importance [134]. Magnesium oxide or magnesia exhibits a rock salt-type structure similar to simple NaCl [135], which is shown in Figure 8. It is known for its excellent optical, thermal, electrical, mechanical, and chemical properties [136]. Additionally, MgO has high thermal stability, with a melting point around (2852 °C) [137], and low heat capacity, making it a good insulator [138]. The particles are non-poisonous since MgO is considered an essential nutrient for plants and humans [139]. Magnesium oxides are known for their biocompatibility and stability; thus, they are frequently used in drug delivery systems and biomedical applications [140,141,142,143,144]. The method of synthesis of MgO highly influences the morphology and the physical structure of the nanoparticles, as in TiO2 [145]. A study synthesized MgO nanostructures with the microwave-assisted process using two different capping agents [139]. The structures obtained were MgO nanospheres and MgO nano-cubes. The nanostructures were then used for the remediation of ciprofloxacin from aqueous solutions, and MgO nanospheres exhibited higher adsorption capacity. Moreover, MgO NPs can be fabricated by several other methods such as biosynthesis [146], ultrasound-assisted methods [147,148], sol–gel methods [149,150], pyrolysis [151], hydrothermal [152], solution combustion [153], and the co-precipitation method [154,155]. Further properties of MgO include a wide bandgap of 7.8 eV, along with high porosity and high surface area, making MgO widely applicable in many technologies [156]. The technological fields involve optoelectronics [157], enhancement of energy conversion efficiency in perovskite solar cells [158], sensors [159], superconductors [160], and toxic waste remediation [161,162,163]. For remediation applications, MgO nanostructures are preferred due to their nano size, which allows them to have high surface area and high surface charge [164]. Lastly, MgO nanomaterials are broadly used in the field of catalysis [165,166,167,168,169].

2.7. Cerium Oxide

Cerium oxide (ceria) is a lanthanide rare earth metal oxide that attracted tremendous interest for its many applications [170]. Different types of cerium oxide nanomaterials are applied in various fields, such as biomedical fields [171,172,173,174,175], sensors [176,177], supercapacitors [178], fuel cells [179,180], adsorbents [181,182], photoprotective coating [183], solar cells [184], and the photodegradation of toxic pollutants [185,186,187]. Moreover, ceria differs from alkaline earth metals and post-transition metals due to its shielded 4f orbital electrons that affect its remarkable properties [188]. Cerium oxide has excellent chemical stability, is inexpensive, and is environmentally friendly [189]. It is highly conductive, with a large magnetic moment [190]. Cerium oxide has two oxidation states, Ce+3 and Ce+4, where cerium (IV) is more thermally stable than the reduced version of cerium. However, cerium can switch to its other oxidation state depending on its surrounding environment [191]. Ceria’s structure consists of a cubic fluorite-type oxide with many oxygen vacancies [192]. The cubic fluorite-type structure of ceria is shown in Figure 9. The surface of cerium oxide has Ce+3 ions as well [193]. Consequently, when cerium oxide nanoparticles (nanoceria) are synthesized, the reactivity of nanoceria increases with the increase of Ce+3 ions concentration, making the NPs good for catalysis applications [194,195]. Several methods are available for the synthesis of nanoceria including the sol–gel [196], biosynthesis [197], hydrothermal [198], sonochemical [199], microwave-assisted [200], and co-precipitation [201] methods.

2.8. Aluminum Oxide

Aluminum oxide (Alumina) is a promising candidate for many applications, it is used in catalysis [202,203], insulators [204], microelectronics [205,206,207], sensors [208], and remediation processes [209,210]. Alumina has unique acid/base characteristics, good mechanical strength, chemical inertness towards oxidation and reduction, excellent electrical insulation, sufficient thermal stability, high surface area, and a high melting point [211,212]. Although alumina is a good insulator, F. Tzompantzi proposed that it might also be effective for photocatalytic degradation [213]. Yanet Pina-Perez proposed that the hydroxyl groups on Al2O3’s surface might be the reason behind its photoactivity [214]. The photocatalytic activity of alumina can be increased by doping the metal ions with other metal oxides [215]. Alumina can be fabricated by various methods such as thermal decomposition [216], hydrothermal [217], combustion [218], co-precipitation [219], and sol–gel methods [220]. However, unlike other metal oxides, alumina needs high calcination temperatures (>1000 °C), which makes the fabrication costly process [221]. Alumina has many crystal structures. The most common one is α- Al2O3 since it is the most thermodynamically stable [222], and the structure is shown in Figure 10 [223]. Alumina has other polymorphs, including η, δ, κ, θ, γ, and ρ phases, which are metastable [224]. The type of phase produced depends highly on the method of synthesis followed. Some factors that affect the phase of alumina include the temperature, pH, pressure, and speed of stirring [225].

2.9. Other Metal Oxides

Besides the former metal oxides discussed, several metal oxides, including MnO2, WO3, and NiO, have been reported for application in the remediation of pesticides from water. Manganese dioxide nanoparticles are prepared using various synthetic routes, including hydrothermal, sol–gel, homogeneous hydrolysis, and sono-chemical methods [11,226,227,228]. They exist in different crystalline structures, such α-MnO2 (hollandite), β-MnO2 (pyrolusite), δ-MnO2 (birnessite), ε-MnO2 (akhtenskite), γ-MnO2 (ramsdellite), λ-MnO2 (defect spinels), and amorphous MnO2 [229,230,231]. The structures of MnO2 polymorphs are shown in Figure 11 [232,233]. They are widely applied as adsorbents and photocatalyst for the removal of different heavy metals and organic pollutants because they are cost effective, their structures are flexible, and they exert no toxicity [234]. Few studies were reported for the removal of pesticides using MnO2 NPs. A removal percentage of 66% within 2 h was found for the photodegradation of 2,4-dichlorophenoxyacetic acid using manganese-doped zinc oxide/graphene nanocomposite under LED light [235].
Various publications were considered for the fabrication of semi-conductor tungsten oxide nanoparticles using the hydrothermal method. For the remediation purposes, most of the time, tungsten oxide nanoparticles are coupled with zinc oxide and used as photocatalyst for the removal of different contaminants. This is attributed to the fact that tungsten oxide NPs have a narrow bandgap of 2.8 eV; therefore, they are weak as photocatalysts, and enhancements can be conducted by doping the material with other metals for high removal efficiency. The degradation efficiency increases with the increase in dopant percentage. At pH = 7, WO3-doped ZnO NPs immobilized on glass plates show 80% removal of 2,4-dichlorophenoxyacetic acid pesticide using UV light within 2 h. Additionally, they show 99% removal of diazinon within 3 h at pH = 7 [236,237]. Figure 12 represents the effect of doping percentage on the degradation efficiency of diazinon [237].
Nickel oxide nanoparticles are hierarchical porous structures, thermally stable and with large surface area [238]. Hence, they are efficient adsorbents for different pollutants [239,240]. Nickel oxide nanoparticles are fabricated using different methods, including sol–gel [241], thermal decomposition [242], biosynthesis [243], and laser ablation [244]. A study synthesized different nanostructures of NiO using the hydrothermal method, varying the reaction temperature, time, and the molar ratios of the precursors [245]. SEM images of the synthesized nanorods, nanoplates, and nanoparticles of NiO are shown in Figure 13 [245]. Furthermore, they are p-type semiconductors with a band gap of 3.6–4.0 eV, which is sufficient for photocatalytic degradation of pesticides [246]. A layered and flower-like structure of S-doped Ni–Co LDH with uniformly dispersed spherical Fe3O4 NPs has shown 92% degradation of chlorpyrifos using visible light at pH = 10 within 150 min [247].

2.10. Effect of Metal Oxide’s Crystalline Structure on the Photocatalytic/Sorption Performance

Transition Metal oxide nanoparticles (NPs) displayed remarkable surface properties, structural characteristics, and a significant specific surface area, which made them desirable candidates for adsorption processes [22]. When the size of the molecules reduces from bulk to nanoscale, it creates an exponential increment in surface-to-volume ratio. By decreasing the size and adding active edges on organic molecules’ surfaces for interaction, surface energy or adsorbent composites are improved. In contrast to their bulkier cousins, nanoparticles are far better at adsorbing organic pollutants from water. Moreover, MO NPs have lately shown a distinct potential as highly selective adsorbents intended for fast and effective removal of organic pollutants, whether used alone or in nanocomposites.
The metal oxides-based nanocomposites serve as large bandgap energy (Eg) semiconductors and have beneficial properties such non-toxicity and stability in water for the breakdown of organic contaminants. They also have correct structure, crystalline, and surface features. The fundamental process for the photocatalytic destruction of impurities from the surface of semiconducting materials is produced by oxygen. Oxygen vaccinations can benefit from semiconducting nanoparticles absorbing photons. By transforming organic pollutants into low/intermediate harmful yields, resulting in substances such as carbon dioxide, water, and inorganic ions, it supports environmental restoration. Once upon a time, photocatalytic treatment was thought to be the most environmentally friendly method of removing organic contaminants from wastewater. Using a short-range solar spectrum is thus a considerable obstacle to photocatalytic activity. The flaw could be fixed, for instance, by fabricating nanomaterials, doping hetero-atoms, and designing metal oxide nanocomposites through chemical and structural alterations. Worthwhile photocatalysts effectively delay electron-hole (e—h+) pair recombination, efficiently absorb the solar spectrum in the visible range, and function well as photocatalysts [248]. As photocatalysts and adsorbents, several metal oxide nanomaterials, such as Al2O3, CuO, CeO2, ZnO, and TiO2, have attracted a lot of interest [248]. To increase effectiveness and selectivity, several MO-based composites, including porous materials–reinforced metal oxides, magnetic metal oxides, metal–metal oxides, graphene–metal oxides, etc., were being studied. Adsorption events are controlled by these nanocomposites’ surface properties, size, and textural characteristics. Several NP morphologies provide flexible crystal defects, such as active edges on material surfaces for photocatalysis and adsorption applications.

3. Classification of Pesticides

The demand for categorizing pesticides has been raised significantly because of the increased number of pesticides, along with the variation in physical and chemical properties [249]. A considerable volume of literature has been published in this field. Recently, scientists classified pesticides based on origin and on target. Pesticides generally originate from organic, inorganic, and biological sources [250]. Table 1 elaborates on the organic class of pesticides, while Table 2 shows the classification of pesticides based on target. The pesticides’ chemical structures are shown in Figure 14.

4. Removal of Pesticides Using Functionalized Metal Oxide Nanomaterials by Adsorption

The hazards and consequences resulting from the massive use of pesticides raised the demand for efficient techniques to be employed for the removal of these contaminants. The adsorption technique has gained popularity as a simple, effective, insensitive, and flexible method [263]. It is a physiochemical method that occurs mostly in the solid–liquid form, though liquid–liquid and liquid–gas forms are also known [264,265,266].
In adsorption, the molecules of liquid or gases are bound to the surface of the solid. The material that provides the surface is called the adsorbent. The contaminants in the liquid or the gaseous phase are called adsorbates. Among the adsorbents reported in the literature, metal oxides have been proven as excellent adsorbents for the remediation of pesticides because of the large surface area provided for the adsorption of the pollutant [267]. The active sites and the functional groups, such as -OH, -COOH, and -C=OH, have a great impact on the efficiency of the adsorption process [268,269]. Moreover, metal oxides, having porous structures, thermal stability, low toxicity, and easy recovery, are all important for a good adsorbent. Two types of interaction between the adsorbent and the adsorbate are present: chemisorption and physisorption. Chemisorption is basically a chemical reaction between the adsorbent and the adsorbate, and it is an irreversible process. It is controlled by chemical bonds such as covalent, chelation, complex formation, proton displacement, and redox-reactions. On the other hand, physisorption, which is more dominant, is a reversible process controlled by Van der Wal’s bonds, dipole–dipole attraction, and London force, etc. [270]. Table 3 provides a comparison between the types of adsorption process [271].
The adsorption process depends on various parameters that need to be optimized, including pH, temperature, time, concentration of contaminant, and sorbent dosage. Table 4 represents the adsorption capacity Qmax (mg/g) and the percentage removal of targeted pesticides using metal oxide nanoparticles at different parameters. The adsorption capacity is calculated in (mg/g) using the formula in Equation (1):
Q m a x = C ° C e m × V
where Co is the initial concentration of the pesticide (mg/L), Ce is the pesticide concentration at equilibrium (mg/L), m is the mass of adsorbent (g), and V is the volume of the solution (L).
The adsorption isotherm and the adsorption kinetics are used to elucidate the adsorption process and to indicate the type of mechanism. The adsorption isotherm is expressed by Langmuir, Freundlich, Sips, Temkin, Redlich Peterson, Henry, and Dubinin–Astakhov (DA) models. Langmuir, Freundlich, and Dubinin–Astakhov models are most frequently used. Langmuir isotherm investigates a monolayer adsorption onto a homogeneous adsorbent, whereas Freundlich illustrates a multilayer adsorption onto a heterogeneous adsorbent. The Dubinin–Astakhov model is used to calculate the mean free adsorption energy E (J/mol). The physisorption mechanism gives an E value smaller than 8 J/mol. However, values of E from 16 J/mol to 40 J/mol indicate a chemisorption mechanism. The adsorption kinetics are equations that indicate the type of interactions between the adsorbent and the adsorbate (contaminant). Chemisorption interaction is described by a pseudo-second-order equation. The pseudo-first-order equation is applied for the physisorption interaction [272,273].
Despite the advantages of adsorption, there is one certain drawback associated with the use of this technique: it produces secondary pollutants which require highly advanced procedures for recycling and decomposing for them to be used in the industrial field [22].
Table 4. Adsorptive remediation of pesticides using metal oxides NPs.
Table 4. Adsorptive remediation of pesticides using metal oxides NPs.
Adsorbent aTargeted Pesticides bTarget Operation ParametersAdsorption ModellingRef.
Pesticide Conc.Adsorbent Dosage (g) or g/LpHTemp.
(K)
Time (min)Kinetics cIsotherm dMechanism eQmax (mg/g) or Percentage Removal (%)/Percentage Recovery
Co3O4/G-MCM-41Methyl parathion-----PFO, PSOL, F, DA-175.2[274]
NiO/Co@CChlorothalonil0.045 g/L0.01 g--15PSOLπ-CM, H62.2[275]
Tebuconazole0.045 g/L0.01 g--15PSOLπ-CM, H40.5
Chlorpyrifos0.045 g/L0.01 g--15PSOLπ-CM, H60.3
Butralin0.045 g/L0.01 g--15PSOLπ-CM, H50.2
Deltamethrin0.045 g/L0.01 g--15PSOLπ-CM, H54.1
Pyridaben0.045 g/L0.01 g--15PSOLπ-CM, H51.3
CeO22,4-Dichlorophenoxyacetic acid0.01 g/L0.025 g-308120PSOL, F, Sπ–π, e95.78[276]
Fe3O4@ZnAl-LDH@MIL-53(Al)Triadimefon5.0–600 mg kg−130 g/L6308.155PSOLπ–π, H, C, (π-CM), P46.08[277]
MgFe2O4Chlorpyrifos20 mg/L0.01 g/L10295360PSOL-4461[278]
Fe3O4Atrazine50 mg/L0.1 g229855PFOL-77.5[279]
Methoxychlor50 mg/L0.1 g229855PFOL-163.9
ZnONaphthalene25 mg/L0.012 g429840PSOL, F, T-66.8[280]
CTAB-ZnONaphthalene25 mg/L0.08 g429840PSOL, F, T-89.96
BMTF-IL-ZnONaphthalene25 mg/L0.06 g429840PSOL, F, T-148.3
ZnO/ZnFe2O4Atrazine50 mL aq. solution0.4 g/L72984320-D.Aπ–π, H, h, e--[281]
Fe3O4@SiO2@GO-2- phenylethylamineChlorpyrifos10 mL aq. Solution0.015 g729815PSOSπ–π, H88%[32]
Malathion10 mL aq. Solution0.015 g729815PSOSH76%
Parathion10 mL aq. Solution0.015 g729815PSOSπ–π, H85%
Fe3O4/MOF-99Dinotefuran0.3–1.5 ng/mL0.015 g--20--π–π88–107%[282]
Thiamethoxam0.3–1.5 ng/mL0.015 g--20--π–π88–107%
Fe3O4@SiO2@MOF/TiO2Triadimenol0.001 g/L0.04 g7298–313.151–60PSO-π–π90.2–104%[283]
Hexaconazole0.001 g/L0.04 g7298–313.151–60PSO-π–π90.2–104%
Diniconazole0.001 g/L0.04 g7298–313.151–60PSO-π–π90.2–104%
Fe3O4-GO@MOF-199.Flusilazole0.002 g/L0.02 g--15--h, π–π, H, e0.0356[284]
Fenbuconazole0.002 g/L0.02 g--15--h, π–π, H, e0.0342
Myclobutanil0.002 g/L0.02 g--15--h, π–π, H, e0.0324
Fe3O4–MWCNTs-ZIF-8Triazophos0.015 g0.002–0.08 g/L4RT15-F-3.12[285]
Diazinon0.015 g0.002–0.08 g/L4RT15-F-2.59
Phosalone0.015 g0.002–0.08 g/L4RT15-F-3.80
Profenofos0.015 g0.002–0.08 g/L4RT15-F-3.89
Methidathion0.015 g0.002–0.08 g/L4RT15-F-2.34
Ethoprop0.015 g0.002–0.08 g/L4RT15-F-2.18
Sulfotep0.015 g0.002–0.08 g/L4RT15-F-2.84
Isazofos0.015 g0.002–0.08 g/L4RT15-F-3
Chitosan–CuOThiophanate-methyl0.1 g/L0.1 g7RT25-L, Fh250[286]
Methomyl0.1 g/L0.1 g7RT25-L.F-20
Malathion0.02 g/L1 g/L2303960PSOL, F-322.6
Chitosan-ZnOThiophanate-methyl0.1 g/L0.1 g7RT25-L, Fh100
Methomyl0.1 g/L0.1 g7RT25-L, F-10
Permethrin0.0001 g/L0.5 g729890---99%[287]
Fe3O4/CuO/Activa-ted carbonImidacloprid0.01 g/L0.02 g729810PSOFC99%[288]
ZnO-IPPsChlorpyrifos0.01–0.6 g/L0.03 g2303–32330PSOL, F, T, D. A-47.846[289]
ZnO-CPMetribuzin0.033–0.1550.08 g3303–36380PSOF-200[290]
MOM-Fe3O4Triclosan0.005–0.2 g/L0.01–0.05 g/L4, 7, 10293, 303, 313600PFOL-103.45[291]
N-NiO@N-Fe3O4@N-ZnOAtrazine0.04 g/L0.1 g5-80PSOL-92%[292]
MgAl2O4Dimethomorph-0.5–2 g5.5-10---% Recovery = 90–94%[293]
Fe3O4 @PSLindane2, 10, 50, 200 µg/L0.02 g/L-RT<20PSOL-10.2[294]
Aldrin2, 10, 50, 200 µg/L0.02 g/L-RT<20PSOL-24.7
Dieldrin2, 10, 50, 200 µg/L2 × 10−5 g/L-RT<20PSOL-21.3
Endrin2, 10, 50, 200 µg/L2 × 10−5 g/L-RT<20PSOL-33.5
MgODiazinon0.30 g/L0.05 or 0.10 g--<5---21–37%[295]
Fenitrothion0.28 g/L0.05 or 0.10 g--5–60-- 27–47%
Fe3O4@nSiO2@mSiO2DDT0.0015 g/L0.05 g--15PSO--94%[296]
RT = room temperature; a Adsorbent: ZnONPs-IPPs = zinc oxide nanoparticles-impregnated pea peels; MOM-Fe3O4 = functionalized iron oxide nanoparticles with Moringa oleifera Lam. seeds; Fe3O4 @PS = magnetic nanosphere coated by polystyrene; ZnO-CP = zinc oxide with cucumber peel; CTAB-ZnO = cetyltrimethylammonium bromide functionalized zinc oxide; BMTF-IL-ZnO = 1-Butyl-3-methylimidazolium tetrafluoroborate functionalized zinc oxide; Hr-MgO = hierarchical magnesium oxide; b targeted pesticides fenitrothion = dimethoxy-(3-methyl-4-nitrophenoxy)-thioxophosphorane; DDT = dichloro-diphenyl-trichloroethane; Diazinone = diethoxy-[(2-isopropyl-6- methyl-4-pyrimidinyl)oxy]-thioxophosphorane; c Kinetic equation; PSO = pseudo-second order; PFO = pseudo-first order; d Isotherm equation; L = Langmuir; F = Freundlich; S = Sips; T = Temkin; DA = Dubinin–Astakhov; e Mechanisms: electrostatic interaction (e), hydrophobic interaction (h), π–π interaction (π–π), π-complex formation with cations (including metal or positive ion charge groups) (π-CM), hydrogen bond interaction (H), coordination or covalent bond (C).

5. Removal of Pesticides Using Functionalized Metal Oxide Nanomaterials by Photocatalytic Degradation

Photocatalytic degradation is an advanced oxidation process that destroys toxic substances into other harmless products. Unlike other remediation techniques, photocatalytic degradation completely mineralizes the toxicant, without the production of secondary waste [36]. The mechanism of photocatalytic degradation starts when the photocatalyst is irradiated under UV or visible light that has energy equal to or greater than its band gap [297]. The detailed mechanism of the reaction is shown in Equation (2) to Equation (8). Notably, photocatalytic degradation of organic molecules is carried out in a similar manner [21]. When the photocatalyst is irradiated, electrons are excited from the valence band of the photocatalyst to the conduction band generating electron/hole pairs (e/h+), as seen in Equation (2).
Oxygen in water becomes attracted to the positive holes generated by the radiation, and a proton leaves the water molecule, leaving hydroxyl ions adsorbed on the surface, which is shown in Equation (3). It is noted that *X resembles a species absorbed into the hole. Electrons act as reducing agents while positive holes act as oxidizing agents. Electrons reduce the oxygen adsorbed on the surface of the photocatalyst, generating a superoxide radical in Equation (4). Then, a superoxide and a proton react to produce a peroxide radical that is still adsorbed on the surface, and a hydrogen transfer from two peroxides occurs to produce hydrogen peroxide and oxygen (Equations (5) and (6)). Finally, hydrogen peroxide is irradiated to produce hydroxyl radicals in Equation (7), and hydroxyl radicals degrade the organic pesticide to water, carbon dioxide, and other products, depending on the type of pesticide (Equation (8)). Figure 15 illustrates a schematic mechanism for the photodegradation of a pesticide [298].
photocatalyst + hv → h+ + e
h+ + H2O → *OH + H+
*O2 + e → *O2
*O2 + H+ → *OOH
2*OOH → *O2 + H2O2
H2O2 + hv → 2 .OH
Pesticide + .OH → intermediates → H2O + CO2
Finding the optimum conditions for photocatalysis is extremely important to achieve maximum efficiency of degradation. The recent studies reporting on the photodegradation of different types of pesticides by metal oxide nanomaterials and their composites under UV or visible light have been cited in Table 5. The conditions that correspond to the maximum efficiency of degradation in the studies have been reported.
Several parameters should be considered when carrying out photocatalytic degradation [248]. The nature and type of the photocatalyst, concentration of the photocatalyst, concentration of the pesticide, pH, and irradiation time. Surface morphology, agglomeration, and size affect the behavior of the photocatalyst during the process. Moreover, the higher the concentration of the photocatalyst, the more efficient the degradation [299]. This is a result of having more active sites on the surface of the photocatalyst, thus generating more electron/hole pairs and, consequently, more hydroxyl radicals. However, it is worth mentioning that after very high dosages of the photocatalyst, the efficiency of the reaction decreases due to the blockage of light penetration [300]. Concerning the concentration of the pesticide, at high dosages of the pollutant, most studies reported a decrease in the efficiency of degradation, as reported in Table 4. Increasing the dosage of the pesticide allows for the adsorption of the pesticide on the active sites of the catalyst, preventing the generation of hydroxyl radicals [301]. Depending on the structure of both pesticide and the nano-photocatalyst, the pH can affect the reaction behavior between them. The reaction will be favorable in the pH that allows for the attraction of the photocatalyst and the pesticide, as well as the accelerated production of hydroxyl radicals [302]. The effect of irradiation time is directly proportional to the efficiency of degradation. The increase of irradiation time permits more excitation to occur, and consequently, more radicals are formed [303].
Metal oxide semiconductors, such as ZnO and TiO2 nanomaterials, are the most appropriate for photocatalytic degradation (Table 4) [298]. This is attributed to the fact that they can produce electron/hole pairs (e/h+) more when irradiated with light. Most photocatalysis research focuses on TiO2 nanomaterials [304,305,306]. The problem with ZnO NPs is the fast recombination of the generated electron/holes [301]. However, recently, it has been discovered that doping the semiconductors with other metals, or further functionalizing them, leads to better separation of charges [307].
Table 5. Reported studies for the photodegradation of pesticides by metal oxide nanomaterials and their composites.
Table 5. Reported studies for the photodegradation of pesticides by metal oxide nanomaterials and their composites.
PhotocatalystStructureTarget PesticideLight SourceConc. of PollutantConc. of PhotocatalystIrradiation Time (min)pHDegradation Efficiency (%)Ref.
Co3O4/MCM-41 NPsMCM-41 nanospheres decorated Co3O4.methyl parathionvisible100 mg/L0.25 g908100[274]
MCM-41/Co3O4
nanocomposite
Spherical shape.acephatevisible100 mg/L0.25 g708100[308]
Co3O4/MCM-41
nanocomposite
MCM-41 spherical grains decorated by Co3O4 NPs.omethoatevisible50 mg/L0.25 g30>6.5100[309]
Cu/ZnO nanocompositeSpherical and elliptical.monocrotophosvisible0.5 L0.5 g1807~90[310]
CuO/TiO2/PANI
nanocomposite
CuO/TiO2 spherical NPs embedded in tubular PANI.chlorpyrifosvisible5 mg/L45 mg90795[35]
ZnO/CuO nanocompositesShape depends on the synthesis conditions.triclopyrUV10 mg/L0.10 g/150 mL1004100[311]
CuO NPsSpherical and flower-like shape.lambda-cyhalothrinUV10 mg/L3 mg/L180799[312]
NiO NPsSpherical and flower-like shape.lambda-cyhalothrinUV10 mg/L4 mg/L180789[312]
Cu2O/BiVO4 compositesShape depends on the synthesis conditions.4-chlorophenolvisible50 mg/L5 g/L240-44[313]
Mn-doped zinc oxide/graphene nanocompositeSpherical particles distributed onto graphene sheets.2,4-dichlorophenoxyacetic acidLED25 mg/L2 g/L120566.2[235]
WO3 doped ZnO NPs immobilized on glass platesHeterogenous surface.2,4-dichlorophenoxyacetic acidUV25 mg/L-120780[236]
Nano hydroxyapatite modified CFGO/ZnO nanorod compositeA complex porous surface.chlorpyrifosvisible5 mg/L0.1 g303100[302]
WO3 doped ZnO NPs immobilized on glass platesHeterogenous surface.diazinonUV10 mg/L10 mg/cm2180799[237]
ZnO/rGO nanocompositerGO film with agglomerations of ZnO nanosheets.dimethoateUV5 mg/L50 mg180-~99[301]
ZnO NPsSpherical.monocrotophosUV500 mL aq. solution2 g120488[314]
ZnO NPs-methyl parathionUV-85 mg/L100>9~70[315]
ZnO NPs-parathionUV-85 mg/L100>9~65[315]
Cu-doped ZnO nanorodsNanorods.diazinonUV20 mg/L0.2 g/L120796.97[36]
ZnO nanorods nanorod incorporated carboxylic
GR/PANI composite
A complex porous surface.diuronvisible5 mg/L0.1 g403.0100[316]
Fe-ZnO nanocompositeRough surface due to Fe ions doped in ZnO.chlorpyrifosUV10 mg/L25 mg/L60-93.5[317]
Ag-ZnO nanocompositeUniform distribution of Ag onto ZnO surfaces.chlorpyrifosSunlight50 mg/L20 mg40-~90[318]
TiO2 NPsAggregated semi-spherical.imidaclopridUV100 mg100 mg/L207.588.15[319]
ZnO NPsAggregated semi-spherical.imidaclopridUV100 mg100 mg/L207.5~80[319]
rGO/Fe3O4/ZnO ternary nanohybridA complex layered surface.metalaxylvisible10 mg/L0.5 g/L120792.11[320]
La-ZnO-PAN fibersLa and ZnO embedded on PAN nanofibers.methyl parathionUV10 mg/L50 mg/L150<3100[321]
La-ZnO-PAN fibersLa and ZnO embedded on PAN nanofibers.atrazineUV10 mg/L30 mg/L60798[299]
rGO/ZnO nanocatalystZnO NPs uniformly distributed on rGO nanosheets.metalaxylUV10 mg/L0.75 g/L120790.25[322]
Cu-ZnO nanocompositeCu loaded on ZnO nanorods.methyl parathionUV500 mg/L20 mg/L80-99[323]
ZnO/CeO2 nanocompositeCeO2 NPs loaded onto ZnO hexagonal
nano-carrots.
triclopyrUV150 mL aq. solution100 mg70783.24[324]
ZnO nanofilmsNanoflowers.temephosSunlight simulator10 mg/L-12-100[325]
Fe/Ag@ZnO nanostructuresNanoflowers.2,4-dichlorophenoxyacetic acidUV/visible62 mg/L0.078 g/L63580[326]
ZnO/TiO2-Fe3O4
nanocomposite
Fe3O4 and TiO2 uniformly distributed on the porous nanostructure of ZnO.chlorpyrifosvisible8 mg/L60 mg501094.8[327]
PANI/ZnO-CoMoO4 nanocompositeSpherical CoMoO4 and ZnO NPs distributed on PANI.imidaclopridvisible4.5 mg/L163.5 mg180497[328]
Ag@ZnO nano-starsStar-like shape.methyl parathionvisible0.01 mg/L25 mg2007-[329]
Pd@ZnO nano-starsStar-like shape.methyl parathionvisible0.01 mg/L25 mg2007-[329]
Cu-ZnO nano heterojunction particlesCu NPs embedded onto ZnO surface.chlorpyrifossunlight200 mg/L250 mg240691[330]
Li dope ZnO nanostructuresAggregated spherical NPs.triclopyrUV100 mL aq. solution1 g/L1207~50[331]
ZnO@CdS nanostructuresCdS aggregated spherical NPs and ZnO nanoflowers.chlorpyrifossunlight2 mg/L25 mg/L360791[300]
ZnO@CdS nanostructuresCdS aggregated spherical NPs and ZnO nanoflowers.atrazinesunlight50 mg/L20 mg/L360789[300]
MgO NPs immobilized on concreteMgO NPs immobilized on concrete surface.diazinonUV5 mg/L-120799.46[332]
CeO2/TiO2/SiO2 nano-catalystNearly sphericalchlorpyrifosUV2 mg/L0.21 g/L905.481.1[333]
CeO2-SiO2 NPs-chlorpyrifosUV10 mL aq. solution7 mg1509~90[334]
Fe doped CeO2-SiO2 nanocompositeSpherical NPschlorpyrifosUV20 mg/L7 mg~230-81.31[335]
GO/Fe3O4/TiO2-NiO
nanocomposite
Spherical Fe3O4, TiO2, NiO dispersed on GO nanosheets.imidaclopridvisible5 mg/L0.08 g45997.47[303]
Au/Fe3O4 core/shell NPsSphericalmalathionUV10 mg/L10−4 mol/L90-76[336]
S-doped Ni–Co LDH/Fe3O4 nanocompositeA layered and flower-like structure with uniformly dispersed spherical Fe3O4 NPs.chlorpyrifosvisible2.5 mg/L60 mg1501092.5[247]
KIT-5/Bi2S3-Fe3O4 nanocompositeSpherical Bi2S3 and Fe3O4 NPs uniformly distributed on 3-D mesoporous cubic KIT-5 surface.parathionvisible4.5 mg/L55 mg55898.7[337]
GO- Fe3O4/TiO2 nanocompositeFe3O4 NPs and mesoporous TiO2 dispersed uniformly on GO nanosheets.chlorpyrifosvisible5 mg/L100 mg60~897[304]
KIT-6/WS2-Fe3O4 nanocompositeSpherical WS2 and Fe3O4 NPs uniformly distributed on 3-D mesoporous cubic KIT-6 surface.chlorpyrifosvisible7.2 mg/L50 mg52692.1[338]
Fe3O4/CdS–ZnS nanocompositeSpherical CdS, Fe3O4 and ZnS NPs.chlorpyrifosvisible10 mg/L0.01 g60794.55[339]
Fe3O4@WO3/SBA-15 nanocompositeAgglomerates of WO3 nanoplates on Fe3O4 NPs and uniform rods of hexagonal SBA-15.2,4-dichlorophenoxyacetic acidUV10−6 mol/L40 mg240-90.73[340]
TNP-Pd-Fe3O4/GO photocatalystFe3O4 NPs, Pd, and TiO2 nanoplates were dispersed uniformly on
GO sheets.
parathionvisible10 mg/L80 mg401098.5[341]
BiOBr/Fe3O4 photocatalystAgglomerated Fe3O4 NPs deposited on BiOBr microspheres.glyphosatevisible100 mg/L0.08 g60-97[342]
Ag2S doped nanostructures of Fe3O4 @Ag3PO4 ultrathin filmsAg2S and Fe3O4 NPs doped on Ag3PO4 ultrathin film.imidaclopridvisible2 mg/L30 mg904.3–998.9[343]
Ag2S doped nanostructures of Fe3O4 @Ag3PO4 ultrathin filmsAg2S and Fe3O4 NPs doped on Ag3PO4 ultrathin film.thiaclopridvisible2 mg/L30 mg60-90[343]
g-C3N4/Cu/TiO2 nanocompositeCu and TiO2 NPs dispersed on the irregular layered structure of graphitic-C3N4.endosulfanvisible5 mg/L40 mg806.860[344]
SBA-15/TiO2 nanocompositeTiO2 NPs dispersed on the hexagonal array of SBA-15.trifluralinUV60 mg/L0.2 g/L301090[345]
SBA-15/TiO2 nanocompositeTiO2 NPs dispersed on the hexagonal array of SBA-15.pendimethalinUV60 mg/L0.2 g/L301082.5[345]
TiO2 NPsIrregular agglomerated NPs.imidaclopridUV5 mg/L0.6 g/L3006.3599[305]
TiO2 nanostructures modified with CuHomogenous nano-porous structure of TiO2 with Cu dispersed on the surface.imidaclopridUV/vis25 mg/L-60--[306]
TiO2/CNT/Pd
photocatalyst
Heterostructure spherical Pd-doped TiO2 nanoparticles on carbon nanotubes.neonicotinoids thiaclopridsunlight5 mg/L0.1 g/L1807100[346]
TiO2/CNT/Pd photocatalystHeterostructure spherical Pd-doped TiO2 nanoparticles on carbon nanotubes.imidaclopridsunlight5 mg/L0.1 g/L180799.8[346]
TiO2/CNT/Pd photocatalystHeterostructure spherical Pd-doped TiO2 nanoparticles on carbon nanotubes.clothianidinsunlight5 mg/L0.1 g/L1807100[346]
TiO2 nanoparticlesSpherical with only a small quantity of hexagonal diameters.dimethoateUV5 mg/L300 mg/L320-100[347]
TiO2 nanoparticlesSpherical with only a small quantity of hexagonal diameters.methomylUV5 mg/L300 mg/L320-100[347]
CuS/TiO2 (CuST) nanoparticlesCoalesced and form a textured/porous nanostructure.4-chlorophenolUV20 mg/L100 mg150-87[348]
Pt@TiO2/rGO
nanocomposite
Monodisperse quasi-spherical Pt@TiO2 NPs deposited on the rGO nanosheets.diuronUV0.03 mmol/L7 mg-7100[349]
(CMC/Tryp/TiO2).Platelet-like crystallites.2,4-dichlorophenolUV200 mg/L0.5 g/L---[350]
SBA-16/TiO2 nanocompositesRutile phase.commercial paraquat (PQ) herbicideUV50 mg/L100 mg in 250 mL aq. Solution1440-70[351]
Ce-TiO2@RGO nanocompositeNon-uniform deposition of Ce-TiO2 with spherical crystalline TiO2 on a reduced graphene oxide sheet.quinalphosVisible-20 mg/L240-92[352]
Ce-TiO2@RGO
nanocomposite
Non-uniform deposition of Ce-TiO2 with spherical crystalline TiO2 on a reduced graphene oxide sheet.imidaclopridVisible-20 mg/L240-85[352]
Ag3PO4/TiO2 NPsCrystallized structure with cubed shape Ag3PO4 and anatase TiO22,4-dichlorophenoxyacetic acidVisible10 mg/L1 g/L60398.4[353]
2D/2D TiO2/MIL88(Fe) (TCS@MOF) nanocompositeStacked layer thin MIL-88(Fe) nanosheet with micro-sized TiO2 nano-granular spherical shape.monocrotophosvisible20 mg/L0.05 g/L305~98.79%[354]
TiO2 nanotubesNanotubesSimazineUV1 mg/L-54-48[355]
TiO2 NPsAgglomerated spherical shape.AcetamipridUV4.5 mg/L2000 mg/L240-100[356]
TiO2 NPs-ImidacloprideUV25 mg/L200 mg/L48-90[357]
TiO2 NPs-1,2-dichloroethaneUV100–200 mg/L100 mg/L360795[358]
N-doped TiO2 nanoparticlesAgglomerated small particles.dichlorodiphenyltrichloroethaneUV10,000 mg/L100048770[359]
lanthanide-doped TiO2 photocatalystsSolely anatase.metazachlorUV10 mg/L1000 mg/L300-85[360]

6. Challenges and Outlook

Despite the exquisite properties and the versatile applications of metal oxide nanomaterials and their composites, there are still inadequacies that cannot be ignored as the prepared material should be cost-effective, eco-friendly, and non-toxic. Recently, the use of functionalized metal oxides as adsorbents for the removal of pesticides has been riddled with many challenges. One challenge is the secondary waste produced from the adsorption process, which has not yet been addressed for the use of these materials as recycled materials in industries. Therefore, further studies on the implementation of the recycled metal oxides in the industrial field should be considered. Additionally, most of the published research neglects the fact that the water bodies are contaminated with multi-contaminants. Therefore, investigations should be conducted to assess the efficiency of metal oxides in the presence of multi-pollutants and a real representative matrix. Although the synthesized metal oxides have wide application, much of the research published does not include assessment on the toxicity of the material itself. It is very important to address and examine the toxicity of these materials and their composites, and to employ metal oxides as adsorbents and photocatalytic materials in commercial applications for the treatment of real samples.

7. Conclusions

Metal oxide nanomaterials and their composites have received considerable attention in recent years owing to their wide applications and eminent properties. Their porous structure, thermal stability, low toxicity, easy recovery, and large surface area make them extensively efficient for remediation applications as adsorbents and photocatalytic materials. Many publications have been collected on the removal of organic pesticides such as algaecides, fungicides, herbicides, insecticides, etc., using metal oxide nanomaterials and their nanocomposites, including metal oxides/metal-organic frameworks, metal oxides/polymers, metal/metal oxides other hybridized composites.
From the research reviewed, it can clearly be concluded that the prominent adsorptive interaction between metal oxides and pesticides is chemisorption. This finding is further supported by the type of mechanism and the type of kinetics, as the pseudo-second-order kinetic equation is used to express the chemisorption interaction. Additionally, the π–π interaction, π-complex interaction, and coordination or covalent bond are all types of chemical bonds. The adsorptive removal of pesticides using metal oxides has gained prominence due to its simplicity, effectivity, insensitivity, and flexibility. It has one limitation, in that it produces secondary products which need further recycling, decomposing, and management to be utilized in industries. Accordingly, photocatalytic degradation has emerged alternatively, which results in the complete mineralization of the pollutant to intermediates and H2O and CO2. Assessment of material toxicity should be focused on more, along with the by-products of adsorption. To scale up the material on an industrial scale, the investigated materials should be tested in real representative matrices that resemble the contaminated water.

Author Contributions

H.H.S., E.F.H.A. and A.H.K. prepared the manuscript, performing the literature survey and data interpretation. A.H.K., S.R. and A.A.A. revised the manuscript. A.A.A. provided the resources and financial support. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available but not put in the public domain.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

POPsPersistent organic chemicals
OCsOrganochlorines
OPsOrganophosphates
NPsNanoparticles
MOMetal oxides
IONPsIron oxide nanoparticles
RTRoom temperature
ZnONPs-IPPsZinc oxide nanoparticles impregnated Pea peels
MOM-Fe3O4Iron oxide nanoparticles with Moringa oleifera Lam. seeds
ZnO-CPZinc oxide with cucumber peel.
CTAB-ZnOCetyltrimethylammonium bromide functionalized Zinc oxide
BMTF-IL-ZnO 1-Butyl-3-methylimidazolium tetrafluoroborate functionalized Zinc oxide.
Hr-MgOHierarchical magnesium oxide
DDTDichloro-diphenyl-trichloroethane
PSOPseudo Second Order
PFOPseudo First Order
LLangmuir isotherm model.
FFreundlich isotherm model.
SSips isotherm model.
TTemkin isotherm model.
D-ADubinin–Astakhov isotherm model.
e-Electrostatic interaction
hHydrophobic interaction
π–ππ–π interaction
π-CMπ-complex formation
HHydrogen bond interaction
CCoordination or covalent bond

References

  1. Adebusuyi, A.T.; Sojinu, S.O.; Aleshinloye, A.O. The prevalence of persistent organic pollutants (POPs) in West Africa—A review. Environ. Chall. 2022, 7, 100486. [Google Scholar] [CrossRef]
  2. Ighalo, J.O.; Yap, P.-S.; Iwuozor, K.O.; Aniagor, C.O.; Liu, T.; Dulta, K.; Iwuchukwu, F.U.; Rangabhashiyam, S. Adsorption of persistent organic pollutants (POPs) from the aqueous environment by nano-adsorbents: A review. Environ. Res. 2022, 212, 113123. [Google Scholar] [CrossRef] [PubMed]
  3. Rani, M.; Shanker, U.; Jassal, V. Recent strategies for removal and degradation of persistent & toxic organochlorine pesticides using nanoparticles: A review. J. Environ. Manag. 2017, 190, 208–222. [Google Scholar] [CrossRef]
  4. Billet, L.S.; Belskis, A.; Hoverman, J.T. Temperature affects the toxicity of pesticides to cercariae of the trematode Echinostoma trivolvis. Aquat. Toxicol. 2022, 245, 106102. [Google Scholar] [CrossRef]
  5. Rasheed, T.; Rizwan, K.; Bilal, M.; Sher, F.; Iqbal, H.M.N. Tailored functional materials as robust candidates to mitigate pesticides in aqueous matrices—A review. Chemosphere 2021, 282, 131056. [Google Scholar] [CrossRef]
  6. Parra-Arroyo, L.; González-González, R.B.; Castillo-Zacarías, C.; Melchor Martínez, E.M.; Sosa-Hernández, J.E.; Bilal, M.; Iqbal, H.M.N.; Barceló, D.; Parra-Saldívar, R. Highly hazardous pesticides and related pollutants: Toxicological, regulatory, and analytical aspects. Sci. Total Environ. 2022, 807, 151879. [Google Scholar] [CrossRef]
  7. Cosgrove, S.; Jefferson, B.; Jarvis, P. Application of activated carbon fabric for the removal of a recalcitrant pesticide from agricultural run-off. Sci. Total Environ. 2022, 815, 152626. [Google Scholar] [CrossRef]
  8. Tyagi, H.; Chawla, H.; Bhandari, H.; Garg, S. Recent-enhancements in visible-light photocatalytic degradation of organochlorines pesticides: A review. Mater. Today Proc. 2022, 49, 3289–3305. [Google Scholar] [CrossRef]
  9. Islam, M.A.; Amin, S.M.N.; Rahman, M.A.; Juraimi, A.S.; Uddin, M.K.; Brown, C.L.; Arshad, A. Chronic effects of organic pesticides on the aquatic environment and human health: A review. Environ. Nanotechnol. Monit. Manag. 2022, 18, 100740. [Google Scholar] [CrossRef]
  10. Wang, W.; Wang, X.; Cheng, N.; Luo, Y.; Lin, Y.; Xu, W.; Du, D. Recent advances in nanomaterials-based electrochemical (bio)sensors for pesticides detection. TrAC Trends Anal. Chem. 2020, 132, 116041. [Google Scholar] [CrossRef]
  11. Wang, Y.; Wang, S.-L.; Xie, T.; Cao, J. Activated carbon derived from waste tangerine seed for the high-performance adsorption of carbamate pesticides from water and plant. Bioresour. Technol. 2020, 316, 123929. [Google Scholar] [CrossRef]
  12. Boruah, P.K.; Darabdhara, G.; Das, M.R. Polydopamine functionalized graphene sheets decorated with magnetic metal oxide nanoparticles as efficient nanozyme for the detection and degradation of harmful triazine pesticides. Chemosphere 2021, 268, 129328. [Google Scholar] [CrossRef]
  13. Bano, K.; Kaushal, S.; Singh, P.P. A review on photocatalytic degradation of hazardous pesticides using heterojunctions. Polyhedron 2021, 209, 115465. [Google Scholar] [CrossRef]
  14. Huang, H.; Xia, C.; Liang, D.; Xie, Y.; Kong, F.; Fu, J.; Dou, Z.; Yang, Q.; Suo, W.; Zhang, Q.; et al. Removal and magnetic recovery of heavy metals and pesticides from soil by layered double hydroxides modified biotite. Chem. Eng. J. 2022, 431, 134113. [Google Scholar] [CrossRef]
  15. Yeganeh, M.; Charkhloo, E.; Sobhi, H.R.; Esrafili, A.; Gholami, M. Photocatalytic processes associated with degradation of pesticides in aqueous solutions: Systematic review and meta-analysis. Chem. Eng. J. 2022, 428, 130081. [Google Scholar] [CrossRef]
  16. Koe, W.S.; Lee, J.W.; Chong, W.C.; Pang, Y.L.; Sim, L.C. An overview of photocatalytic degradation: Photocatalysts, mechanisms, and development of photocatalytic membrane. Environ. Sci. Pollut. Res. 2020, 27, 2522–2565. [Google Scholar] [CrossRef]
  17. Ni, J.; Lei, J.; Wang, Z.; Huang, L.; Zhu, H.; Liu, H.; Hu, F.; Qu, T.; Yang, H.; Yang, H.; et al. The Ultrahigh Adsorption Capacity and Excellent Photocatalytic Degradation Activity of Mesoporous CuO with Novel Architecture. Nanomaterials 2023, 13, 142. [Google Scholar] [CrossRef]
  18. Dhananjay, B.; Pangarkar, V.; Beenackers, A. Photocatalytic Degradation for Environmental Application-A Review. J. Chem. Technol. Biotechnol. 2001, 77, 102–116. [Google Scholar] [CrossRef]
  19. Jamshaid, M.; Nazir, M.A.; Najam, T.; Shah, S.S.A.; Khan, H.M.; Rehman, A.u. Facile synthesis of Yb3+-Zn2+ substituted M type hexaferrites: Structural, electric and photocatalytic properties under visible light for methylene blue removal. Chem. Phys. Lett. 2022, 805, 139939. [Google Scholar] [CrossRef]
  20. Ahmadi, A.; Hajilou, M.; Zavari, S.; Yaghmaei, S. A comparative review on adsorption and photocatalytic degradation of classified dyes with metal/non-metal-based modification of graphitic carbon nitride nanocomposites: Synthesis, mechanism, and affecting parameters. J. Clean. Prod. 2023, 382, 134967. [Google Scholar] [CrossRef]
  21. Qumar, U.; Hassan, J.Z.; Bhatti, R.A.; Raza, A.; Nazir, G.; Nabgan, W.; Ikram, M. Photocatalysis vs adsorption by metal oxide nanoparticles. J. Mater. Sci. Technol. 2022, 131, 122–166. [Google Scholar] [CrossRef]
  22. Gusain, R.; Gupta, K.; Joshi, P.; Khatri, O.P. Adsorptive removal and photocatalytic degradation of organic pollutants using metal oxides and their composites: A comprehensive review. Adv. Colloid Interface Sci. 2019, 272, 102009. [Google Scholar] [CrossRef] [PubMed]
  23. Hua, M.; Zhang, S.; Pan, B.; Zhang, W.; Lv, L.; Zhang, Q. Heavy metal removal from water/wastewater by nanosized metal oxides: A review. J. Hazard. Mater. 2012, 211–212, 317–331. [Google Scholar] [CrossRef] [PubMed]
  24. Farooq, U.; Ahmad, T.; Naaz, F.; Islam, S.u. Review on Metals and Metal Oxides in Sustainable Energy Production: Progress and Perspectives. Energy Fuels 2023, 37, 1577–1632. [Google Scholar] [CrossRef]
  25. Selmani, A.; Kovačević, D.; Bohinc, K. Nanoparticles: From synthesis to applications and beyond. Adv. Colloid Interface Sci. 2022, 303, 102640. [Google Scholar] [CrossRef] [PubMed]
  26. Bibi, S.; Ahmad, A.; Anjum, M.A.R.; Haleem, A.; Siddiq, M.; Shah, S.S.; Kahtani, A.A. Photocatalytic degradation of malachite green and methylene blue over reduced graphene oxide (rGO) based metal oxides (rGO-Fe3O4/TiO2) nanocomposite under UV-visible light irradiation. J. Environ. Chem. Eng. 2021, 9, 105580. [Google Scholar] [CrossRef]
  27. Khandare, D.; Mukherjee, S. A Review of Metal oxide Nanomaterials for Fluoride decontamination from Water Environment. Mater. Today Proc. 2019, 18, 1146–1155. [Google Scholar] [CrossRef]
  28. Farooq, U.; Naz, F.; Phul, R.; Pandit, N.A.; Jain, S.K.; Ahmad, T. Development of Heterostructured Ferroelectric SrZrO3/CdS Photocatalysts with Enhanced Surface Area and Photocatalytic Activity. J. Nanosci. Nanotechnol. 2020, 20, 3770–3779. [Google Scholar] [CrossRef]
  29. Jamshaid, M.; Khan, M.I.; Fernandez, J.; Shanableh, A.; Hussain, T.; Rehman, A.u. Synthesis of Ti4+ doped Ca-BiFO3 for the enhanced photodegradation of moxifloxacin. New J. Chem. 2022, 46, 19848–19856. [Google Scholar] [CrossRef]
  30. Shi, R.; Zhang, Z.; Luo, F. N-doped graphene-based CuO/WO3/Cu composite material with performances of catalytic decomposition 4-nitrophenol and photocatalytic degradation of organic dyes. Inorg. Chem. Commun. 2020, 121, 108246. [Google Scholar] [CrossRef]
  31. Wolski, L.; Grzelak, K.; Muńko, M.; Frankowski, M.; Grzyb, T.; Nowaczyk, G. Insight into photocatalytic degradation of ciprofloxacin over CeO2/ZnO nanocomposites: Unravelling the synergy between the metal oxides and analysis of reaction pathways. Appl. Surf. Sci. 2021, 563, 150338. [Google Scholar] [CrossRef]
  32. Wanjeri, V.W.O.; Sheppard, C.J.; Prinsloo, A.R.E.; Ngila, J.C.; Ndungu, P.G. Isotherm and kinetic investigations on the adsorption of organophosphorus pesticides on graphene oxide based silica coated magnetic nanoparticles functionalized with 2-phenylethylamine. J. Environ. Chem. Eng. 2018, 6, 1333–1346. [Google Scholar] [CrossRef]
  33. Singh, A.; Sikarwar, S.; Verma, A.; Chandra Yadav, B. The recent development of metal oxide heterostructures based gas sensor, their future opportunities and challenges: A review. Sens. Actuators A Phys. 2021, 332, 113127. [Google Scholar] [CrossRef]
  34. Wang, X.-L.; Sun, Y.-Y.; Xiao, Y.; Chen, X.-X.; Huang, X.-C.; Zhou, H.-L. Facile Solution-Refluxing Synthesis and Photocatalytic Dye Degradation of a Dynamic Covalent Organic Framework. Molecules 2022, 27, 8002. [Google Scholar] [CrossRef]
  35. Nekooie, R.; Shamspur, T.; Mostafavi, A. Novel CuO/TiO2/PANI nanocomposite: Preparation and photocatalytic investigation for chlorpyrifos degradation in water under visible light irradiation. J. Photochem. Photobiol. A Chem. 2021, 407, 113038. [Google Scholar] [CrossRef]
  36. Shirzad-Siboni, M.; Jonidi-Jafari, A.; Farzadkia, M.; Esrafili, A.; Gholami, M. Enhancement of photocatalytic activity of Cu-doped ZnO nanorods for the degradation of an insecticide: Kinetics and reaction pathways. J. Environ. Manag. 2017, 186, 1–11. [Google Scholar] [CrossRef] [PubMed]
  37. Hussain, I.; Lee, J.M.; Iqbal, S.; Kim, H.S.; Jang, S.W.; Jung, J.Y.; An, H.J.; Lamiel, C.; Mohamed, S.G.; Lee, Y.R.; et al. Preserved crystal phase and morphology: Electrochemical influence of copper and iron co-doped cobalt oxide and its supercapacitor applications. Electrochim. Acta 2020, 340, 135953. [Google Scholar] [CrossRef]
  38. Gopinath, S.; Mayakannan, M.; Vetrivel, S. Structural, optical, morphological properties of silver doped cobalt oxide nanoparticles by microwave irradiation method. Ceram. Int. 2022, 48, 6103–6115. [Google Scholar] [CrossRef]
  39. Rasheed, T.; Nabeel, F.; Bilal, M.; Iqbal, H.M.N. Biogenic synthesis and characterization of cobalt oxide nanoparticles for catalytic reduction of direct yellow-142 and methyl orange dyes. Biocatal. Agric. Biotechnol. 2019, 19, 101154. [Google Scholar] [CrossRef]
  40. Letsholathebe, D.; Thema, F.T.; Mphale, K.; Mohamed, H.E.A.; Holonga, K.J.; Ketlhwaafetse, R.; Chimidza, S. Optical and structural stability of Co3O4 nanoparticles for photocatalytic applications. Mater. Today Proc. 2021, 36, 499–503. [Google Scholar] [CrossRef]
  41. Dharaskar, S.; Kodgire, P.; Bansod, P. Enhanced diesel properties with energy efficient nano-aluminium oxide and nano-cobalt oxide particles. Mater. Today Proc. 2021, 62, 6927–6933. [Google Scholar] [CrossRef]
  42. Dey, S.; Dhal, G.C. The catalytic activity of cobalt nanoparticles for low-temperature oxidation of carbon monoxide. Mater. Today Chem. 2019, 14, 100198. [Google Scholar] [CrossRef]
  43. Uddin, M.K.; Baig, U. Synthesis of Co3O4 nanoparticles and their performance towards methyl orange dye removal: Characterisation, adsorption and response surface methodology. J. Clean. Prod. 2019, 211, 1141–1153. [Google Scholar] [CrossRef]
  44. Lima, A.F. Interpretation of the optical absorption spectrum of Co3O4 with normal spinel structure from first principles calculations. J. Phys. Chem. Solids 2014, 75, 148–152. [Google Scholar] [CrossRef]
  45. Asha, G.; Rajeshwari, V.; Stephen, G.; Gurusamy, S.; Carolin Jeniba Rachel, D. Eco-friendly synthesis and characterization of cobalt oxide nanoparticles by sativum species and its photo-catalytic activity. Mater. Today Proc. 2022, 48, 486–493. [Google Scholar] [CrossRef]
  46. Navarrete, È.; Bittencourt, C.; Noirfalise, X.; Umek, P.; González, E.; Güell, F.; Llobet, E. WO3 nanowires loaded with cobalt oxide nanoparticles, deposited by a two-step AACVD for gas sensing applications. Sens. Actuators B Chem. 2019, 298, 126868. [Google Scholar] [CrossRef]
  47. Jincy, C.S.; Meena, P. Evaluation of cytotoxic activity of Fe doped cobalt oxide nanoparticles. J. Trace Elem. Med. Biol. 2022, 70, 126916. [Google Scholar] [CrossRef]
  48. Gaikar, P.S.; Angre, A.P.; Wadhawa, G.; Ledade, P.V.; Mahmood, S.H.; Lambat, T.L. Green synthesis of cobalt oxide thin films as an electrode material for electrochemical capacitor application. Curr. Res. Green Sustain. Chem. 2022, 5, 100265. [Google Scholar] [CrossRef]
  49. Lakra, R.; Kumar, R.; Sahoo, P.K.; Sharma, D.; Thatoi, D.; Soam, A. Facile synthesis of cobalt oxide and graphene nanosheets nanocomposite for aqueous supercapacitor application. Carbon Trends 2022, 7, 100144. [Google Scholar] [CrossRef]
  50. Rakotonarivo, E.F.; Abouloula, C.N.; Narjis, A.; Nkhaili, L.; Brouillette, F.; Oueriagli, A. Optimization of the electrodeposition of the pure and cobalt doped copper oxide for solar cells and other applications. Phys. B Condens. Matter 2021, 609, 412783. [Google Scholar] [CrossRef]
  51. Jincy, C.S.; Meena, P. Synthesis, characterization, and NH3 gas sensing application of Zn doped cobalt oxide nanoparticles. Inorg. Chem. Commun. 2020, 120, 108145. [Google Scholar] [CrossRef]
  52. Ni, Q.; Ma, J.; Fan, C.; Kong, Y.; Peng, M.; Komarneni, S. Spinel-type cobalt-manganese oxide catalyst for degradation of Orange II using a novel heterogeneous photo-chemical catalysis system. Ceram. Int. 2018, 44, 19474–19480. [Google Scholar] [CrossRef]
  53. Feng, H.; Liang, L.; Ge, J.; Wu, W.; Huang, Z.; Liu, Y.; Li, L. Delicate manipulation of cobalt oxide nanodot clusterization on binder-free TiO2-nanorod photoanodes for efficient photoelectrochemical catalysis. J. Alloys Compd. 2020, 820, 153139. [Google Scholar] [CrossRef]
  54. Ullah, R.; Rasheed, M.A.; Abbas, S.; Rehman, K.-u.; Shah, A.; Ullah, K.; Khan, Y.; Bibi, M.; Ahmad, M.; Ali, G. Electrochemical sensing of H2O2 using cobalt oxide modified TiO2 nanotubes. Curr. Appl. Phys. 2022, 38, 40–48. [Google Scholar] [CrossRef]
  55. Wang, F.; Li, H.; Yuan, Z.; Sun, Y.; Chang, F.; Deng, H.; Xie, L.; Li, H. A highly sensitive gas sensor based on CuO nanoparticles synthetized via a sol–gel method. RSC Adv. 2016, 6, 79343–79349. [Google Scholar] [CrossRef]
  56. Censabella, M.; Iacono, V.; Scandurra, A.; Moulaee, K.; Neri, G.; Ruffino, F.; Mirabella, S. Low temperature detection of nitric oxide by CuO nanoparticles synthesized by pulsed laser ablation. Sens. Actuators B Chem. 2022, 358, 131489. [Google Scholar] [CrossRef]
  57. Heinemann, M.; Eifert, B.; Heiliger, C. Band structure and phase stability of the copper oxides Cu2O, CuO, and Cu4O3. Phys. Rev. B 2013, 87, 115111. [Google Scholar] [CrossRef]
  58. Wojcieszak, D.; Obstarczyk, A.; Mańkowska, E.; Mazur, M.; Kaczmarek, D.; Zakrzewska, K.; Mazur, P.; Domaradzki, J. Thermal oxidation impact on the optoelectronic and hydrogen sensing properties of p-type copper oxide thin films. Mater. Res. Bull. 2022, 147, 111646. [Google Scholar] [CrossRef]
  59. Kim, H.; Park, S.; Park, Y.; Choi, D.; Yoo, B.; Lee, C.S. Fabrication of a semi-transparent flexible humidity sensor using kinetically sprayed cupric oxide film. Sens. Actuators B Chem. 2018, 274, 331–337. [Google Scholar] [CrossRef]
  60. Moghanlou, A.O.; Sadr, M.H.; Bezaatpour, A.; Salimi, F.; Yosefi, M. RGO/Cu2O-CuO nanocomposite as a visible-light assisted photocatalyst for reduction of organic nitro groups to amines. Mol. Catal. 2021, 516, 111997. [Google Scholar] [CrossRef]
  61. Vázquez-Vargas, D.A.; Amézaga-Madrid, P.; Jáuregui-Martínez, L.E.; Esquivel-Pereyra, O.; Antúnez-Flores, W.; Pizá-Ruiz, P.; Miki-Yoshida, M. Synthesis and microstructural characterization of cupric oxide and cobalt oxide nanostructures for their application as selective solar coatings. Thin Solid Film. 2020, 706, 138046. [Google Scholar] [CrossRef]
  62. Maruthupandy, M.; Zuo, Y.; Chen, J.-S.; Song, J.-M.; Niu, H.-L.; Mao, C.-J.; Zhang, S.-Y.; Shen, Y.-H. Synthesis of metal oxide nanoparticles (CuO and ZnO NPs) via biological template and their optical sensor applications. Appl. Surf. Sci. 2017, 397, 167–174. [Google Scholar] [CrossRef]
  63. Mohammed Ali, M.J.; Radhy, M.M.; mashkoor, S.J.; Ali, E.M. Synthesis and characterization of copper oxide nanoparticles and their application for solar cell. Mater. Today Proc. 2021, 60, 917–921. [Google Scholar] [CrossRef]
  64. Pourbeyram, S.; Abdollahpour, J.; Soltanpour, M. Green synthesis of copper oxide nanoparticles decorated reduced graphene oxide for high sensitive detection of glucose. Mater. Sci. Eng. C 2019, 94, 850–857. [Google Scholar] [CrossRef]
  65. Nabila, M.I.; Kannabiran, K. Biosynthesis, characterization and antibacterial activity of copper oxide nanoparticles (CuO NPs) from actinomycetes. Biocatal. Agric. Biotechnol. 2018, 15, 56–62. [Google Scholar] [CrossRef]
  66. Kim, T.K.; VanSaders, B.; Moon, J.; Kim, T.; Liu, C.-H.; Khamwannah, J.; Chun, D.; Choi, D.; Kargar, A.; Chen, R.; et al. Tandem structured spectrally selective coating layer of copper oxide nanowires combined with cobalt oxide nanoparticles. Nano Energy 2015, 11, 247–259. [Google Scholar] [CrossRef]
  67. Nwanya, A.C.; Razanamahandry, L.C.; Bashir, A.K.H.; Ikpo, C.O.; Nwanya, S.C.; Botha, S.; Ntwampe, S.K.O.; Ezema, F.I.; Iwuoha, E.I.; Maaza, M. Industrial textile effluent treatment and antibacterial effectiveness of Zea mays L. Dry husk mediated bio-synthesized copper oxide nanoparticles. J. Hazard. Mater. 2019, 375, 281–289. [Google Scholar] [CrossRef]
  68. McDonald, K.J.; Reynolds, B.; Reddy, K.J. Intrinsic properties of cupric oxide nanoparticles enable effective filtration of arsenic from water. Sci. Rep. 2015, 5, 11110. [Google Scholar] [CrossRef]
  69. Srivastava, V.; Choubey, A.K. Investigation of adsorption of organic dyes present in wastewater using chitosan beads immobilized with biofabricated CuO nanoparticles. J. Mol. Struct. 2021, 1242, 130749. [Google Scholar] [CrossRef]
  70. Naaz, F.; Sharma, A.; Shahazad, M.; Ahmad, T. Hydrothermally Derived Hierarchical CuO Nanoflowers as an Efficient Photocatalyst and Electrocatalyst for Hydrogen Evolution. ChemistrySelect 2022, 7, e202201800. [Google Scholar] [CrossRef]
  71. Gupta, K.; Chundawat, T.S. Zinc oxide nanoparticles synthesized using Fusarium oxysporum to enhance bioethanol production from rice-straw. Biomass Bioenergy 2020, 143, 105840. [Google Scholar] [CrossRef]
  72. Bandeira, M.; Possan, A.L.; Pavin, S.S.; Raota, C.S.; Vebber, M.C.; Giovanela, M.; Roesch-Ely, M.; Devine, D.M.; Crespo, J.S. Mechanism of formation, characterization and cytotoxicity of green synthesized zinc oxide nanoparticles obtained from Ilex paraguariensis leaves extract. Nano-Struct. Nano-Objects 2020, 24, 100532. [Google Scholar] [CrossRef]
  73. Darvishi, E.; Kahrizi, D.; Arkan, E. Comparison of different properties of zinc oxide nanoparticles synthesized by the green (using Juglans regia L. leaf extract) and chemical methods. J. Mol. Liq. 2019, 286, 110831. [Google Scholar] [CrossRef]
  74. Sultana, K.A.; Islam, M.T.; Silva, J.A.; Turley, R.S.; Hernandez-Viezcas, J.A.; Gardea-Torresdey, J.L.; Noveron, J.C. Sustainable synthesis of zinc oxide nanoparticles for photocatalytic degradation of organic pollutant and generation of hydroxyl radical. J. Mol. Liq. 2020, 307, 112931. [Google Scholar] [CrossRef]
  75. Dkhil, M.A.; Diab, M.S.M.; Aljawdah, H.M.A.; Murshed, M.; Hafiz, T.A.; Al-Quraishy, S.; Bauomy, A.A. Neuro-biochemical changes induced by zinc oxide nanoparticles. Saudi J. Biol. Sci. 2020, 27, 2863–2867. [Google Scholar] [CrossRef] [PubMed]
  76. Brindhadevi, K.; Samuel, M.S.; Verma, T.N.; Vasantharaj, S.; Sathiyavimal, S.; Saravanan, M.; Pugazhendhi, A.; Duc, P.A. Zinc oxide nanoparticles (ZnONPs) -induced antioxidants and photocatalytic degradation activity from hybrid grape pulp extract (HGPE). Biocatal. Agric. Biotechnol. 2020, 28, 101730. [Google Scholar] [CrossRef]
  77. Dang, Z.; Sun, J.; Fan, J.; Li, J.; Li, X.; Chen, T. Zinc oxide spiky nanoparticles: A promising nanomaterial for killing tumor cells. Mater. Sci. Eng. C 2021, 124, 112071. [Google Scholar] [CrossRef]
  78. Shankar, S.; Rhim, J.-W. Effect of types of zinc oxide nanoparticles on structural, mechanical and antibacterial properties of poly(lactide)/poly(butylene adipate-co-terephthalate) composite films. Food Packag. Shelf Life 2019, 21, 100327. [Google Scholar] [CrossRef]
  79. Jayachandran, A.; Aswathy, T.R.; Nair, A.S. Green synthesis and characterization of zinc oxide nanoparticles using Cayratia pedata leaf extract. Biochem. Biophys. Rep. 2021, 26, 100995. [Google Scholar] [CrossRef]
  80. Bandeira, M.; Giovanela, M.; Roesch-Ely, M.; Devine, D.M.; da Silva Crespo, J. Green synthesis of zinc oxide nanoparticles: A review of the synthesis methodology and mechanism of formation. Sustain. Chem. Pharm. 2020, 15, 100223. [Google Scholar] [CrossRef]
  81. Ranjithkumar, B.; Ramalingam, H.B.; Kumar, E.R.; Srinivas, C.; Magesh, G.; Rahale, C.S.; El-Metwaly, N.M.; Shekar, B.C. Natural fuels (Honey and Cow urine) assisted combustion synthesis of zinc oxide nanoparticles for antimicrobial activities. Ceram. Int. 2021, 47, 14475–14481. [Google Scholar] [CrossRef]
  82. Archana, P.; Janarthanan, B.; Bhuvana, S.; Rajiv, P.; Sharmila, S. Concert of zinc oxide nanoparticles synthesized using Cucumis melo by green synthesis and the antibacterial activity on pathogenic bacteria. Inorg. Chem. Commun. 2022, 137, 109255. [Google Scholar] [CrossRef]
  83. Espitia, P.J.P.; Soares, N.d.F.F.; Coimbra, J.S.d.R.; de Andrade, N.J.; Cruz, R.S.; Medeiros, E.A.A. Zinc Oxide Nanoparticles: Synthesis, Antimicrobial Activity and Food Packaging Applications. Food Bioprocess Technol. 2012, 5, 1447–1464. [Google Scholar] [CrossRef]
  84. Rajbongshi, B.M.; Samdarshi, S.K. Cobalt-doped zincblende–wurtzite mixed-phase ZnO photocatalyst nanoparticles with high activity in visible spectrum. Appl. Catal. B Environ. 2014, 144, 435–441. [Google Scholar] [CrossRef]
  85. Doan Thi, T.U.; Nguyen, T.T.; Thi, Y.D.; Ta Thi, K.H.; Phan, B.T.; Pham, K.N. Green synthesis of ZnO nanoparticles using orange fruit peel extract for antibacterial activities. RSC Adv. 2020, 10, 23899–23907. [Google Scholar] [CrossRef]
  86. Kanimozhi, S.; Prabu, K.M.; Thambidurai, S.; Suresh, S. Dye-sensitized solar cell performance and photocatalytic activity enhancement using binary zinc oxide-copper oxide nanocomposites prepared via co-precipitation route. Ceram. Int. 2021, 47, 30234–30246. [Google Scholar] [CrossRef]
  87. Rakha, A.A.; Shahzad, M.; Ghaffar, A.; Javed, K.; Pervez, A.; Sarwar, N.; Munam, M. Power conversion efficiency of hybrid solar cells based on Camellia sinensis doped polyvinyl alcohol and ZnO nanoparticles. Opt. Mater. 2021, 120, 111434. [Google Scholar] [CrossRef]
  88. Abood, M.K.; Fayad, M.A.; Al Salihi, H.A.; Salbi, H.A.A. Effect of ZnO nanoparticles deposition on porous silicon solar cell. Mater. Today Proc. 2021, 42, 2935–2940. [Google Scholar] [CrossRef]
  89. Hanabaratti, R.M.; Tuwar, S.M.; Nandibewoor, S.T.; Gowda, J.I. Fabrication and characterization of zinc oxide nanoparticles modified glassy carbon electrode for sensitive determination of paracetamol. Chem. Data Collect. 2020, 30, 100540. [Google Scholar] [CrossRef]
  90. Hussein, H.T.; Kareem, M.H.; Abdul Hussein, A.M. Synthesis and characterization of carbon nanotube doped with zinc oxide nanoparticles CNTs-ZnO/PS as ethanol gas sensor. Optik 2021, 248, 168107. [Google Scholar] [CrossRef]
  91. Chelladurai, M.; Margavelu, G.; Vijayakumar, S.; González-Sánchez, Z.I.; Vijayan, K.; Sahadevan, R. Preparation and characterization of amine-functionalized mupirocin-loaded zinc oxide nanoparticles: A potent drug delivery agent in targeting human epidermoid carcinoma (A431) cells. J. Drug Deliv. Sci. Technol. 2022, 70, 103244. [Google Scholar] [CrossRef]
  92. Singh, T.A.; Das, J.; Sil, P.C. Zinc oxide nanoparticles: A comprehensive review on its synthesis, anticancer and drug delivery applications as well as health risks. Adv. Colloid Interface Sci. 2020, 286, 102317. [Google Scholar] [CrossRef]
  93. Paiva-Santos, A.C.; Herdade, A.M.; Guerra, C.; Peixoto, D.; Pereira-Silva, M.; Zeinali, M.; Mascarenhas-Melo, F.; Paranhos, A.; Veiga, F. Plant-mediated green synthesis of metal-based nanoparticles for dermopharmaceutical and cosmetic applications. Int. J. Pharm. 2021, 597, 120311. [Google Scholar] [CrossRef] [PubMed]
  94. Mustapha, S.; Ndamitso, M.M.; Abdulkareem, A.S.; Tijani, J.O.; Shuaib, D.T.; Ajala, A.O.; Mohammed, A.K. Application of TiO2 and ZnO nanoparticles immobilized on clay in wastewater treatment: A review. Appl. Water Sci. 2020, 10, 49. [Google Scholar] [CrossRef]
  95. Aksu Demirezen, D.; Yıldız, Y.Ş.; Yılmaz, Ş.; Demirezen Yılmaz, D. Green synthesis and characterization of iron oxide nanoparticles using Ficus carica (common fig) dried fruit extract. J. Biosci. Bioeng. 2019, 127, 241–245. [Google Scholar] [CrossRef]
  96. Yu, X.; Zhou, J. Grain boundary in oxide scale during high-temperature metal processing. Study Grain Bound. Character 2017, 59–77. [Google Scholar] [CrossRef]
  97. Machala, L.; Tuček, J.; Zbořil, R. Polymorphous Transformations of Nanometric Iron(III) Oxide: A Review. Chem. Mater. 2011, 23, 3255–3272. [Google Scholar] [CrossRef]
  98. Lassoued, A.; Lassoued, M.S.; Dkhil, B.; Ammar, S.; Gadri, A. Synthesis, structural, morphological, optical and magnetic characterization of iron oxide (α-Fe2O3) nanoparticles by precipitation method: Effect of varying the nature of precursor. Phys. E Low-Dimens. Syst. Nanostruct. 2018, 97, 328–334. [Google Scholar] [CrossRef]
  99. Ong, H.T.; Suppiah, D.D.; Muhd Julkapli, N. Fatty acid coated iron oxide nanoparticle: Effect on stability, particle size and magnetic properties. Colloids Surf. A Physicochem. Eng. Asp. 2020, 606, 125371. [Google Scholar] [CrossRef]
  100. Borges, R.; Mendonça-Ferreira, L.; Rettori, C.; Pereira, I.S.O.; Baino, F.; Marchi, J. New sol-gel-derived magnetic bioactive glass-ceramics containing superparamagnetic hematite nanocrystals for hyperthermia application. Mater. Sci. Eng. C 2021, 120, 111692. [Google Scholar] [CrossRef]
  101. Karaagac, O.; Köçkar, H. Improvement of the saturation magnetization of PEG coated superparamagnetic iron oxide nanoparticles. J. Magn. Magn. Mater. 2022, 551, 169140. [Google Scholar] [CrossRef]
  102. Al-Jabari, M.H.; Sulaiman, S.; Ali, S.; Barakat, R.; Mubarak, A.; Khan, S.A. Adsorption study of levofloxacin on reusable magnetic nanoparticles: Kinetics and antibacterial activity. J. Mol. Liq. 2019, 291, 111249. [Google Scholar] [CrossRef]
  103. Shi, S.-F.; Jia, J.-F.; Guo, X.-K.; Zhao, Y.-P.; Chen, D.-S.; Guo, Y.-Y.; Cheng, T.; Zhang, X.-L. Biocompatibility of chitosan-coated iron oxide nanoparticles with osteoblast cells. Int. J. Nanomed. 2012, 7, 5593. [Google Scholar]
  104. Unsoy, G.; Yalcin, S.; Khodadust, R.; Gündüz, G.; Gunduz, U. Synthesis optimization and characterization of chitosan-coated iron oxide nanoparticles produced for biomedical applications. J. Nanopart. Res. 2012, 14, 964. [Google Scholar] [CrossRef]
  105. Lopez, J.D.; Keley, M.; Dante, A.; Werneck, M.M. Optical fiber sensor coated with copper and iron oxide nanoparticles for hydrogen sulfide sensing. Opt. Fiber Technol. 2021, 67, 102731. [Google Scholar] [CrossRef]
  106. Pakapongpan, S.; Poo-arporn, Y.; Tuantranont, A.; Poo-arporn, R.P. A facile one-pot synthesis of magnetic iron oxide nanoparticles embed N-doped graphene modified magnetic screen printed electrode for electrochemical sensing of chloramphenicol and diethylstilbestrol. Talanta 2022, 241, 123184. [Google Scholar] [CrossRef] [PubMed]
  107. Tung, T.T.; Chien, N.V.; Van Duy, N.; Van Hieu, N.; Nine, M.J.; Coghlan, C.J.; Tran, D.N.H.; Losic, D. Magnetic iron oxide nanoparticles decorated graphene for chemoresistive gas sensing: The particle size effects. J. Colloid Interface Sci. 2019, 539, 315–325. [Google Scholar] [CrossRef]
  108. Roy, S.D.; Das, K.C.; Dhar, S.S. Conventional to green synthesis of magnetic iron oxide nanoparticles; its application as catalyst, photocatalyst and toxicity: A short review. Inorg. Chem. Commun. 2021, 134, 109050. [Google Scholar] [CrossRef]
  109. Moodley, P.; Scheijen, F.J.E.; Niemantsverdriet, J.W.; Thüne, P.C. Iron oxide nanoparticles on flat oxidic surfaces—Introducing a new model catalyst for Fischer–Tropsch catalysis. Catal. Today 2010, 154, 142–148. [Google Scholar] [CrossRef]
  110. Carrillo, A.I.; Serrano, E.; Luque, R.; García-Martínez, J. Microwave-assisted catalysis by iron oxide nanoparticles on MCM-41: Effect of the support morphology. Appl. Catal. A Gen. 2013, 453, 383–390. [Google Scholar] [CrossRef]
  111. Fatimah, I.; Amaliah, S.N.; Andrian, M.F.; Handayani, T.P.; Nurillahi, R.; Prakoso, N.I.; Wicaksono, W.P.; Chuenchom, L. Iron oxide nanoparticles supported on biogenic silica derived from bamboo leaf ash for rhodamine B photodegradation. Sustain. Chem. Pharm. 2019, 13, 100149. [Google Scholar] [CrossRef]
  112. Algarín, M.; Amaya, M.; Solano, R.; Patiño-Ruiz, D.; Herrera, A. Synthesis of a magnetic iron oxide/zinc oxide engineered nanocatalyst for enhanced visible-light photodegradation of Cartasol brilliant violet 5BFN in aqueous solution. Nano-Struct. Nano-Objects 2021, 26, 100730. [Google Scholar] [CrossRef]
  113. Alves, F.H.d.O.; Araújo, O.A.; de Oliveira, A.C.; Garg, V.K. Preparation and characterization of PAni(CA)/Magnetic iron oxide hybrids and evaluation in adsorption/photodegradation of blue methylene dye. Surf. Interfaces 2021, 23, 100954. [Google Scholar] [CrossRef]
  114. Xu, H.; Yuan, H.; Yu, J.; Lin, S. Study on the competitive adsorption and correlational mechanism for heavy metal ions using the carboxylated magnetic iron oxide nanoparticles (MNPs-COOH) as efficient adsorbents. Appl. Surf. Sci. 2019, 473, 960–966. [Google Scholar] [CrossRef]
  115. Lin, S.; Liu, L.; Yang, Y.; Lin, K. Study on preferential adsorption of cationic-style heavy metals using amine-functionalized magnetic iron oxide nanoparticles (MIONPs-NH2) as efficient adsorbents. Appl. Surf. Sci. 2017, 407, 29–35. [Google Scholar] [CrossRef]
  116. Javid, A.; Ahmadian, S.; Saboury, A.; Kalantar, M.; Rezaei-Zarchi, S. Chitosan-Coated Superparamagnetic Iron Oxide Nanoparticles for Doxorubicin Delivery: Synthesis and Anticancer Effect Against Human Ovarian Cancer Cells. Chem. Biol. Drug Des. 2013, 82, 296–306. [Google Scholar] [CrossRef]
  117. Jain, M.; Yadav, M.; Kohout, T.; Lahtinen, M.; Garg, V.K.; Sillanpää, M. Development of iron oxide/activated carbon nanoparticle composite for the removal of Cr(VI), Cu(II) and Cd(II) ions from aqueous solution. Water Resour. Ind. 2018, 20, 54–74. [Google Scholar] [CrossRef]
  118. Wei, Y.; Zhu, J.; Gan, Y.; Cheng, G. Titanium glycolate-derived TiO2 nanomaterials: Synthesis and applications. Adv. Powder Technol. 2018, 29, 2289–2311. [Google Scholar] [CrossRef]
  119. Bortamuly, R.; Naresh, V.; Das, M.R.; Kumar, V.K.; Muduli, S.; Martha, S.K.; Saikia, P. Titania supported bio-derived activated carbon as an electrode material for high-performance supercapacitors. J. Energy Storage 2021, 42, 103144. [Google Scholar] [CrossRef]
  120. Zhang, Y.; Yang, H.M.; Park, S.-J. Synthesis and characterization of nitrogen-doped TiO2 coatings on reduced graphene oxide for enhancing the visible light photocatalytic activity. Curr. Appl. Phys. 2018, 18, 163–169. [Google Scholar] [CrossRef]
  121. Pérez-Jiménez, L.E.; Solis-Cortazar, J.C.; Rojas-Blanco, L.; Perez-Hernandez, G.; Martinez, O.S.; Palomera, R.C.; Paraguay-Delgado, F.; Zamudio-Torres, I.; Morales, E.R. Enhancement of optoelectronic properties of TiO2 films containing Pt nanoparticles. Results Phys. 2019, 12, 1680–1685. [Google Scholar] [CrossRef]
  122. Bayan, E.M.; Lupeiko, T.G.; Pustovaya, L.E.; Volkova, M.G.; Butova, V.V.; Guda, A.A. Zn–F co-doped TiO2 nanomaterials: Synthesis, structure and photocatalytic activity. J. Alloys Compd. 2020, 822, 153662. [Google Scholar] [CrossRef]
  123. Haggerty, J.E.; Schelhas, L.T.; Kitchaev, D.A.; Mangum, J.S.; Garten, L.M.; Sun, W.; Stone, K.H.; Perkins, J.D.; Toney, M.F.; Ceder, G. High-fraction brookite films from amorphous precursors. Sci. Rep. 2017, 7, 15232. [Google Scholar] [CrossRef] [PubMed]
  124. Zhang, J.; Yan, S.; Fu, L.; Wang, F.; Yuan, M.; Luo, G.; Xu, Q.; Wang, X.; Li, C. Photocatalytic Degradation of Rhodamine B on Anatase, Rutile, and Brookite TiO2. Chin. J. Catal. 2011, 32, 983–991. [Google Scholar] [CrossRef]
  125. Kumar, S.G.; Rao, K.S.R.K. Comparison of modification strategies towards enhanced charge carrier separation and photocatalytic degradation activity of metal oxide semiconductors (TiO2, WO3 and ZnO). Appl. Surf. Sci. 2017, 391, 124–148. [Google Scholar] [CrossRef]
  126. Harlapur, S.F.; Rashmi, B.N.; Nagaswarupa, H.P.; Prashantha, S.C.; Shashishekar, T.R.; Anil Kumar, M.R. Photocatalytic studies of TiO2 nanomaterials prepared via facile wet chemical route. Mater. Today Proc. 2017, 4, 11713–11719. [Google Scholar] [CrossRef]
  127. Umar, A.A.; Rahman, M.Y.A.; Saad, S.K.M.; Salleh, M.M.; Oyama, M. Preparation of grass-like TiO2 nanostructure thin films: Effect of growth temperature. Appl. Surf. Sci. 2013, 270, 109–114. [Google Scholar] [CrossRef]
  128. D’Arienzo, M.; Scotti, R.; Di Credico, B.; Redaelli, M. Chapter 13—Synthesis and Characterization of Morphology-Controlled TiO2 Nanocrystals: Opportunities and Challenges for their Application in Photocatalytic Materials. In Studies in Surface Science and Catalysis; Fornasiero, P., Cargnello, M., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; Volume 177, pp. 477–540. [Google Scholar]
  129. Fan, Z.; Meng, F.; Zhang, M.; Wu, Z.; Sun, Z.; Li, A. Solvothermal synthesis of hierarchical TiO2 nanostructures with tunable morphology and enhanced photocatalytic activity. Appl. Surf. Sci. 2016, 360, 298–305. [Google Scholar] [CrossRef]
  130. Narzary, S.; Alamelu, K.; Raja, V.; Jaffar Ali, B.M. Visible light active, magnetically retrievable Fe3O4@SiO2@g-C3N4/TiO2 nanocomposite as efficient photocatalyst for removal of dye pollutants. J. Environ. Chem. Eng. 2020, 8, 104373. [Google Scholar] [CrossRef]
  131. Qiao, H.; Xiao, H.; Huang, Y.; Yuan, C.; Zhang, X.; Bu, X.; Wang, Z.; Han, S.; Zhang, L.; Su, Z.; et al. SiO2 loading into polydopamine-functionalized TiO2 nanotubes for biomedical applications. Surf. Coat. Technol. 2019, 364, 170–179. [Google Scholar] [CrossRef]
  132. Malevu, T.D. Ball Milling synthesis and characterization of highly crystalline TiO2-ZnO hybrids for photovoltaic applications. Phys. B Condens. Matter 2021, 621, 413291. [Google Scholar] [CrossRef]
  133. Ullah, F.; Qureshi, M.T.; Sultana, K.; Saleem, M.; Al Elaimi, M.; Abdel Hameed, R.; Haq, S.u.; Ismail, H.S.; Anwar, M.S. Structural and dielectric studies of MgAl2O4–TiO2 composites for energy storage applications. Ceram. Int. 2021, 47, 30665–30670. [Google Scholar] [CrossRef]
  134. Abinaya, S.; Kavitha, H.P.; Prakash, M.; Muthukrishnaraj, A. Green synthesis of magnesium oxide nanoparticles and its applications: A review. Sustain. Chem. Pharm. 2021, 19, 100368. [Google Scholar] [CrossRef]
  135. Nagappa, B.; Chandrappa, G.T. Mesoporous nanocrystalline magnesium oxide for environmental remediation. Microporous Mesoporous Mater. 2007, 106, 212–218. [Google Scholar] [CrossRef]
  136. Mehta, M.; Mukhopadhyay, M.; Christian, R.; Mistry, N. Synthesis and characterization of MgO nanocrystals using strong and weak bases. Powder Technol. 2012, 226, 213–221. [Google Scholar] [CrossRef]
  137. Abdulkhaleq, N.A.; Nayef, U.M.; Albarazanchi, A.K.H. MgO nanoparticles synthesis via laser ablation stationed on porous silicon for photoconversion application. Optik 2020, 212, 164793. [Google Scholar] [CrossRef]
  138. Saito, A.; Obata, S.; Nishina, Y. Uniform coating of magnesium oxide crystal with reduced graphene oxide achieves moisture barrier performance. Appl. Surf. Sci. 2022, 573, 151483. [Google Scholar] [CrossRef]
  139. Sivaselvam, S.; Premasudha, P.; Viswanathan, C.; Ponpandian, N. Enhanced removal of emerging pharmaceutical contaminant ciprofloxacin and pathogen inactivation using morphologically tuned MgO nanostructures. J. Environ. Chem. Eng. 2020, 8, 104256. [Google Scholar] [CrossRef]
  140. Alkhudhayri, A.; Thagfan, F.A.; Al-Quraishy, S.; Abdel-Gaber, R.; Dkhil, M.A. Assessment of the oxidative status and goblet cell response during eimeriosis and after treatment of mice with magnesium oxide nanoparticles. Saudi J. Biol. Sci. 2022, 29, 1234–1238. [Google Scholar] [CrossRef]
  141. El-Sawy, N.M.; Raafat, A.I.; Badawy, N.A.; Mohamed, A.M. Radiation development of pH-responsive (xanthan-acrylic acid)/MgO nanocomposite hydrogels for controlled delivery of methotrexate anticancer drug. Int. J. Biol. Macromol. 2020, 142, 254–264. [Google Scholar] [CrossRef]
  142. Zheng, X.; Feng, S.; Wang, X.; Shi, Z.; Mao, Y.; Zhao, Q.; Wang, S. MSNCs and MgO-MSNCs as drug delivery systems to control the adsorption kinetics and release rate of indometacin. Asian J. Pharm. Sci. 2019, 14, 275–286. [Google Scholar] [CrossRef]
  143. Yang, S.; Liang, L.; Liu, L.; Yin, Y.; Liu, Y.; Lei, G.; Zhou, K.; Huang, Q.; Wu, H. Using MgO nanoparticles as a potential platform to precisely load and steadily release Ag ions for enhanced osteogenesis and bacterial killing. Mater. Sci. Eng. C 2021, 119, 111399. [Google Scholar] [CrossRef]
  144. Nigam, A.; Saini, S.; Rai, A.K.; Pawar, S.J. Structural, optical, cytotoxicity, and antimicrobial properties of MgO, ZnO and MgO/ZnO nanocomposite for biomedical applications. Ceram. Int. 2021, 47, 19515–19525. [Google Scholar] [CrossRef]
  145. Karthik, K.; Dhanuskodi, S.; Gobinath, C.; Prabukumar, S.; Sivaramakrishnan, S. Fabrication of MgO nanostructures and its efficient photocatalytic, antibacterial and anticancer performance. J. Photochem. Photobiol. B Biol. 2019, 190, 8–20. [Google Scholar] [CrossRef]
  146. Nguyen, D.T.C.; Dang, H.H.; Vo, D.-V.N.; Bach, L.G.; Nguyen, T.D.; Tran, T.V. Biogenic synthesis of MgO nanoparticles from different extracts (flower, bark, leaf) of Tecoma stans (L.) and their utilization in selected organic dyes treatment. J. Hazard. Mater. 2021, 404, 124146. [Google Scholar] [CrossRef]
  147. Štengl, V.; Bakardjieva, S.; Maříková, M.; Bezdička, P.; Šubrt, J. Magnesium oxide nanoparticles prepared by ultrasound enhanced hydrolysis of Mg-alkoxides. Mater. Lett. 2003, 57, 3998–4003. [Google Scholar] [CrossRef]
  148. Karuppusamy, I.; Samuel, M.S.; Selvarajan, E.; Shanmugam, S.; Sahaya Murphin Kumar, P.; Brindhadevi, K.; Pugazhendhi, A. Ultrasound-assisted synthesis of mixed calcium magnesium oxide (CaMgO2) nanoflakes for photocatalytic degradation of methylene blue. J. Colloid Interface Sci. 2021, 584, 770–778. [Google Scholar] [CrossRef]
  149. Possato, L.G.; Gonçalves, R.G.L.; Santos, R.M.M.; Chaves, T.F.; Briois, V.; Pulcinelli, S.H.; Martins, L.; Santilli, C.V. Sol-gel synthesis of nanocrystalline MgO and its application as support in Ni/MgO catalysts for ethanol steam reforming. Appl. Surf. Sci. 2021, 542, 148744. [Google Scholar] [CrossRef]
  150. Possato, L.G.; Pereira, E.; Gonçalves, R.G.L.; Pulcinelli, S.H.; Martins, L.; Santilli, C.V. Controlling the porosity and crystallinity of MgO catalysts by addition of surfactant in the sol-gel synthesis. Catal. Today 2020, 344, 52–58. [Google Scholar] [CrossRef]
  151. Dong, B.; Qin, W.; Su, Y.; Wang, X. Magnetic properties of FeSiCr@MgO soft magnetic composites prepared by magnesium acetate pyrolysis for high-frequency applications. J. Magn. Magn. Mater. 2021, 539, 168350. [Google Scholar] [CrossRef]
  152. Barati Dalenjan, M.; Rashidi, A.; Khorasheh, F.; Ardjmand, M. Effect of Ni ratio on mesoporous Ni/MgO nanocatalyst synthesized by one-step hydrothermal method for thermal catalytic decomposition of CH4 to H2. Int. J. Hydrogen Energy 2022, 47, 11539–11551. [Google Scholar] [CrossRef]
  153. Sherikar, B.N.; Sahoo, B.; Umarji, A.M. Effect of fuel and fuel to oxidizer ratio in solution combustion synthesis of nanoceramic powders: MgO, CaO and ZnO. Solid State Sci. 2020, 109, 106426. [Google Scholar] [CrossRef]
  154. Rao, L.S.; Rao, T.V.; Naheed, S.; Rao, P.V. Structural and optical properties of zinc magnesium oxide nanoparticles synthesized by chemical co-precipitation. Mater. Chem. Phys. 2018, 203, 133–140. [Google Scholar] [CrossRef]
  155. Rajendran, V.; Deepa, B.; Mekala, R. Studies on structural, morphological, optical and antibacterial activity of Pure and Cu-doped MgO nanoparticles synthesized by co-precipitation method. Mater. Today Proc. 2018, 5, 8796–8803. [Google Scholar] [CrossRef]
  156. Borgohain, X.; Boruah, A.; Sarma, G.K.; Rashid, M.H. Rapid and extremely high adsorption performance of porous MgO nanostructures for fluoride removal from water. J. Mol. Liq. 2020, 305, 112799. [Google Scholar] [CrossRef]
  157. Lei, P.; Ding, X.; Bai, Y.; Yu, X.; Jiang, G.; Li, T.; Sun, B.; Zhang, X.; Li, X.; Wu, L.; et al. 1.8-μm MgO: PPLN optical parametric oscillator pumped by Nd: YVO4/YVO4 2nd-Stokes Raman laser. Results Phys. 2021, 29, 104703. [Google Scholar] [CrossRef]
  158. Huang, S.; Kang, B.; Duan, L.; Zhang, D. Highly efficient inverted polymer solar cells by using solution processed MgO/ZnO composite interfacial layers. J. Colloid Interface Sci. 2021, 583, 178–187. [Google Scholar] [CrossRef]
  159. Chetankumar, K.; Swamy, B.E.K.; Naik, H.S.B. MgO and MWCNTs amplified electrochemical sensor for guanine, adenine and epinephrine. Mater. Chem. Phys. 2021, 267, 124610. [Google Scholar] [CrossRef]
  160. Sakoda, M.; Aibara, M.; Mede, K.; Kikuchi, M.; Naito, M. Superconducting tunnel junctions on MgB2 using MgO and CaF2 as a barrier. Phys. C Supercond. Its Appl. 2016, 530, 82–86. [Google Scholar] [CrossRef]
  161. Guo, T.; Bulin, C. Facile fabrication of MgO/graphene oxide composite as an efficient adsorbent for rapid removal of aqueous organic dyes: Performance evaluation and mechanistic investigation. J. Phys. Chem. Solids 2021, 158, 110251. [Google Scholar] [CrossRef]
  162. Nga, N.K.; Thuy Chau, N.T.; Viet, P.H. Preparation and characterization of a chitosan/MgO composite for the effective removal of reactive blue 19 dye from aqueous solution. J. Sci. Adv. Mater. Devices 2020, 5, 65–72. [Google Scholar] [CrossRef]
  163. Cai, Y.; Li, C.; Wu, D.; Wang, W.; Tan, F.; Wang, X.; Wong, P.K.; Qiao, X. Highly active MgO nanoparticles for simultaneous bacterial inactivation and heavy metal removal from aqueous solution. Chem. Eng. J. 2017, 312, 158–166. [Google Scholar] [CrossRef]
  164. Prado, D.C.; Fernández, I.; Rodríguez-Páez, J.E. MgO nanostructures: Synthesis, characterization and tentative mechanisms of nanoparticles formation. Nano-Struct. Nano-Objects 2020, 23, 100482. [Google Scholar] [CrossRef]
  165. Yang, D.; Xu, P.; Guan, E.; Browning, N.D.; Gates, B.C. Rhodium pair-sites on magnesium oxide: Synthesis, characterization, and catalysis of ethylene hydrogenation. J. Catal. 2016, 338, 12–20. [Google Scholar] [CrossRef]
  166. Maruthai, J.; Ramachandran, K.; Muthukumarasamy, A.; Chidambaram, S.; Gaidi, M.; Dauodi, K. Bio fabrication of 2D MgO/Ag nanocomposite for effective environmental utilization in antibacterial, anti-oxidant and catalytic applications. Surf. Interfaces 2022, 30, 101921. [Google Scholar] [CrossRef]
  167. Dong, H.; Liu, M.; Yan, X.; Liu, M.; Qian, Z.; Xie, Y.; Luo, W.; Lei, C.; Zhou, Z. Pyrolysis gas from biomass and plastics over X-Mo@MgO (X = Ni, Fe, Co) catalysts into functional carbon nanocomposite: Gas reforming reaction and proper process mechanisms. Sci. Total Environ. 2022, 831, 154751. [Google Scholar] [CrossRef]
  168. Sharmin, F.; Chandra Roy, D.; Basith, M.A. Photocatalytic water splitting ability of Fe/MgO-rGO nanocomposites towards hydrogen evolution. Int. J. Hydrogen Energy 2021, 46, 38232–38246. [Google Scholar] [CrossRef]
  169. Shao, Y.; Zhao, D.-Y.; Lu, W.; Long, Y.; Zheng, W.; Zhao, J.; Hu, Z.-T. MgO/Carbon nanocomposites synthesized in molten salts for catalytic isomerization of glucose to fructose in aqueous media. Green Chem. Eng. 2021, 3, 359–366. [Google Scholar] [CrossRef]
  170. Hamidian, K.; Saberian, M.R.; Miri, A.; Sharifi, F.; Sarani, M. Doped and un-doped cerium oxide nanoparticles: Biosynthesis, characterization, and cytotoxic study. Ceram. Int. 2021, 47, 13895–13902. [Google Scholar] [CrossRef]
  171. Sebastiammal, S.; Bezy, N.A.; Somaprabha, A.; Henry, J.; Biju, C.S.; Fathima, A.L. Chemical and sweet basil leaf mediated synthesis of cerium oxide (CeO2) nanoparticles: Antibacterial action toward human pathogens. Phosphorus Sulfur Silicon Relat. Elem. 2022, 197, 237–243. [Google Scholar] [CrossRef]
  172. He, L.; Su, Y.; Lanhong, J.; Shi, S. Recent advances of cerium oxide nanoparticles in synthesis, luminescence and biomedical studies: A review. J. Rare Earths 2015, 33, 791–799. [Google Scholar] [CrossRef]
  173. Niemiec, S.M.; Hilton, S.A.; Wallbank, A.; Azeltine, M.; Louiselle, A.E.; Elajaili, H.; Allawzi, A.; Xu, J.; Mattson, C.; Dewberry, L.C.; et al. Cerium oxide nanoparticle delivery of microRNA-146a for local treatment of acute lung injury. Nanomed. Nanotechnol. Biol. Med. 2021, 34, 102388. [Google Scholar] [CrossRef]
  174. Asgharzadeh, F.; Hashemzadeh, A.; Rahmani, F.; Yaghoubi, A.; Nazari, S.E.; Avan, A.; Mehr, S.M.H.; Soleimanpour, S.; Khazaei, M. Cerium oxide nanoparticles acts as a novel therapeutic agent for ulcerative colitis through anti-oxidative mechanism. Life Sci. 2021, 278, 119500. [Google Scholar] [CrossRef] [PubMed]
  175. Hasanzadeh, L.; Darroudi, M.; Ramezanian, N.; Zamani, P.; Aghaee-Bakhtiari, S.H.; Nourmohammadi, E.; Kazemi Oskuee, R. Polyethylenimine-associated cerium oxide nanoparticles: A novel promising gene delivery vector. Life Sci. 2019, 232, 116661. [Google Scholar] [CrossRef] [PubMed]
  176. Wan, G.X.; Feng, Y.; Zhang, F.F.; Sun, X.W.; Li, Y.X.; Xue, R.J. Hydrothermal synthesis of Ag doped ZnO/CeO2 nanocomposites and its application in ethanol sensing. Mater. Lett. 2022, 318, 132191. [Google Scholar] [CrossRef]
  177. Negi, K.; Kumar, M.; Singh, G.; Chauhan, S.; Chauhan, M.S. Nanostructured CeO2 for selective-sensing and smart photocatalytic applications. Ceram. Int. 2018, 44, 15281–15289. [Google Scholar] [CrossRef]
  178. Veeresha, G.; Krishnamurthy, G.; Shivakumar, M.S. Cobalt nanocrystals doped on CeO2/RGO nanocomposite for supercapacitor applications. Inorg. Chem. Commun. 2022, 138, 109232. [Google Scholar] [CrossRef]
  179. Miller, H.A.; Bellini, M.; Dekel, D.R.; Vizza, F. Recent developments in Pd-CeO2 nano-composite electrocatalysts for anodic reactions in anion exchange membrane fuel cells. Electrochem. Commun. 2022, 135, 107219. [Google Scholar] [CrossRef]
  180. Wang, P.; Cai, C.; Tan, J.; Pan, M. Effect of the CeO2 nanoparticles in microporous layers on the durability of proton exchange membrane fuel cells. Int. J. Hydrogen Energy 2021, 46, 34867–34873. [Google Scholar] [CrossRef]
  181. Li, C.-M.; Zhang, Y.-S.; Wang, X.-P.; Yin, X.-B.; Luo, N.-N.; Khayambashi, A.; Wei, Y.-Z. The synthesis and characterization of hydrous cerium oxide nanoparticles loaded on porous silica micro-sphere as novel and efficient adsorbents to remove phosphate radicals from water. Microporous Mesoporous Mater. 2019, 279, 73–81. [Google Scholar] [CrossRef]
  182. Yu, Y.; Zhang, C.; Yang, L.; Paul Chen, J. Cerium oxide modified activated carbon as an efficient and effective adsorbent for rapid uptake of arsenate and arsenite: Material development and study of performance and mechanisms. Chem. Eng. J. 2017, 315, 630–638. [Google Scholar] [CrossRef]
  183. Bansal, R.; Nair, S.; Pandey, K.K. UV resistant wood coating based on zinc oxide and cerium oxide dispersed linseed oil nano-emulsion. Mater. Today Commun. 2022, 30, 103177. [Google Scholar] [CrossRef]
  184. Gayathri, R.; Raja, G.; Rajeswaran, P. A simple and one step low cost microwave induced low cost grapheme modified CeO2 photo electrodes for high-efficiency dye-sensitized solar cells. Inorg. Chem. Commun. 2020, 120, 108132. [Google Scholar] [CrossRef]
  185. Latha, P.; Prakash, K.; Karuthapandian, S. Effective Photodegradation of CR & MO dyes by morphologically controlled Cerium oxide nanocubes under visible light Illumination. Optik 2018, 154, 242–250. [Google Scholar] [CrossRef]
  186. Abbasi, M.A.; Amin, K.M.; Ali, M.; Ali, Z.; Atif, M.; Ensinger, W.; Khalid, W. Synergetic effect of adsorption-photocatalysis by GO−CeO2nanocomposites for photodegradation of doxorubicin. J. Environ. Chem. Eng. 2022, 10, 107078. [Google Scholar] [CrossRef]
  187. Iqbal, A.; Ahmed, A.S.; Ahmad, N.; Shafi, A.; Ahamad, T.; Khan, M.Z.; Srivastava, S. Biogenic synthesis of CeO2 nanoparticles and its potential application as an efficient photocatalyst for the degradation of toxic amido black dye. Environ. Nanotechnol. Monit. Manag. 2021, 16, 100505. [Google Scholar] [CrossRef]
  188. Fang, X.; Song, H. Synthesis of cerium oxide nanoparticles loaded on chitosan for enhanced auto-catalytic regenerative ability and biocompatibility for the spinal cord injury repair. J. Photochem. Photobiol. B Biol. 2019, 191, 83–87. [Google Scholar] [CrossRef]
  189. Mazloum-Ardakani, M.; Mohammadian-Sarcheshmeh, H.; Naderi, H.; Farbod, F.; Sabaghian, F. Fabrication of a high-performance hybrid supercapacitor using a modified graphene aerogel/cerium oxide nanoparticle composite. J. Energy Storage 2019, 26, 100998. [Google Scholar] [CrossRef]
  190. Sridharan, M.; Kamaraj, P.; Vennilaraj; Arockiaselvi, J.; Pushpamalini, T.; Vivekanand, P.A.; Hari Kumar, S. Synthesis, characterization and evaluation of biosynthesized Cerium oxide nanoparticle for its anticancer activity on breast cancer cell (MCF 7). Mater. Today Proc. 2021, 36, 914–919. [Google Scholar] [CrossRef]
  191. Shekunova, T.O.; Lapkina, L.A.; Shcherbakov, A.B.; Meshkov, I.N.; Ivanov, V.K.; Tsivadze, A.Y.; Gorbunova, Y.G. Deactivation of singlet oxygen by cerium oxide nanoparticles. J. Photochem. Photobiol. A Chem. 2019, 382, 111925. [Google Scholar] [CrossRef]
  192. Selvaraj Janaki, N.J.; Ivan Jebakumar, D.S.; Sumithraj Premkumar, P. Studies on the physical properties of green synthesized cerium oxide nanoparticles using Melia dubia leaf extract. Mater. Today Proc. 2021, 58, 850–854. [Google Scholar] [CrossRef]
  193. Hancock, M.L.; Yokel, R.A.; Beck, M.J.; Calahan, J.L.; Jarrells, T.W.; Munson, E.J.; Olaniyan, G.A.; Grulke, E.A. The characterization of purified citrate-coated cerium oxide nanoparticles prepared via hydrothermal synthesis. Appl. Surf. Sci. 2021, 535, 147681. [Google Scholar] [CrossRef]
  194. Celardo, I.; Traversa, E.; Ghibelli, L. Cerium oxide nanoparticles: A promise for applications in therapy. J. Exp. Ther. Oncol. 2011, 9, 47–51. [Google Scholar] [PubMed]
  195. Orge, C.A.; Órfão, J.J.M.; Pereira, M.F.R.; Duarte de Farias, A.M.; Neto, R.C.R.; Fraga, M.A. Ozonation of model organic compounds catalysed by nanostructured cerium oxides. Appl. Catal. B Environ. 2011, 103, 190–199. [Google Scholar] [CrossRef]
  196. Akbari, A.; Khammar, M.; Taherzadeh, D.; Rajabian, A.; Khorsand Zak, A.; Darroudi, M. Zinc-doped cerium oxide nanoparticles: Sol-gel synthesis, characterization, and investigation of their in vitro cytotoxicity effects. J. Mol. Struct. 2017, 1149, 771–776. [Google Scholar] [CrossRef]
  197. Muthuvel, A.; Jothibas, M.; Mohana, V.; Manoharan, C. Green synthesis of cerium oxide nanoparticles using Calotropis procera flower extract and their photocatalytic degradation and antibacterial activity. Inorg. Chem. Commun. 2020, 119, 108086. [Google Scholar] [CrossRef]
  198. Syed Khadar, Y.A.; Balamurugan, A.; Devarajan, V.P.; Subramanian, R.; Dinesh Kumar, S. Synthesis, characterization and antibacterial activity of cobalt doped cerium oxide (CeO2:Co) nanoparticles by using hydrothermal method. J. Mater. Res. Technol. 2019, 8, 267–274. [Google Scholar] [CrossRef]
  199. Kishor Kumar, M.J.; Kalathi, J.T. Low-temperature sonochemical synthesis of high dielectric Lanthanum doped Cerium oxide nanopowder. J. Alloys Compd. 2018, 748, 348–354. [Google Scholar] [CrossRef]
  200. Soren, S.; Bessoi, M.; Parhi, P. A rapid microwave initiated polyol synthesis of cerium oxide nanoparticle using different cerium precursors. Ceram. Int. 2015, 41, 8114–8118. [Google Scholar] [CrossRef]
  201. Ahmed, S.H.; Bakiro, M.; Aljasmi, F.I.A.; Albreiki, A.M.O.; Bayane, S.; Alzamly, A. Investigation of the band gap and photocatalytic properties of CeO2/rGO composites. Mol. Catal. 2020, 486, 110874. [Google Scholar] [CrossRef]
  202. Faghihinezhad, M.; Baghdadi, M.; Shahin, M.S.; Torabian, A. Catalytic ozonation of real textile wastewater by magnetic oxidized g-C3N4 modified with Al2O3 nanoparticles as a novel catalyst. Sep. Purif. Technol. 2022, 283, 120208. [Google Scholar] [CrossRef]
  203. Bahari, M.B.; Mamat, C.R.; Jalil, A.A.; Hassan, N.S.; Nabgan, W.; Setiabudi, H.D.; Vo, D.-V.N.; Phuong Thuy, B.T. Mesoporous alumina: A comprehensive review on synthesis strategies, structure, and applications as support for enhanced H2 generation via CO2-CH4 reforming. Int. J. Hydrogen Energy 2022, 47, 41507–41526. [Google Scholar] [CrossRef]
  204. Contreras, J.E.; Taha-Tijerina, J.; López-Perales, J.F.; Banda-Muñoz, F.; Díaz-Tato, L.; Rodríguez, E.A. Enhancing the quartz-clay-feldspar system by nano-Al2O3 addition for electrical insulators: From laboratory to prototype scale. Mater. Chem. Phys. 2021, 263, 124389. [Google Scholar] [CrossRef]
  205. Chang, Y.-J.; Xie, G.-H.; Zhou, Y.-M.; Wang, J.-X.; Wang, Z.-X.; Guo, H.-J.; You, B.-Z.; Yan, G.-C. Enhancing storage performance of P2-type Na2/3Fe1/2Mn1/2O2 cathode materials by Al2O3 coating. Trans. Nonferrous Met. Soc. China 2022, 32, 262–272. [Google Scholar] [CrossRef]
  206. Khaleduzzaman, S.S.; Sohel, M.R.; Saidur, R.; Selvaraj, J. Stability of Al2O3-water Nanofluid for Electronics Cooling System. Procedia Eng. 2015, 105, 406–411. [Google Scholar] [CrossRef]
  207. Goyal, R.K.; Tiwari, A.N.; Mulik, U.P.; Negi, Y.S. Novel high performance Al2O3/poly(ether ether ketone) nanocomposites for electronics applications. Compos. Sci. Technol. 2007, 67, 1802–1812. [Google Scholar] [CrossRef]
  208. Moutcine, A.; Ifguis, O.; Amine Samaini, M.; Ennachete, M.; Sâadane, H.; Laghlimi, C.; Chtaini, A. Simultaneous electrochemical determination of heavy metals by an electrode modified CPE-NP-Al2O3. Mater. Today Proc. 2022, 53, 404–407. [Google Scholar] [CrossRef]
  209. Saadi, Z.; Saadi, R.; Fazaeli, R. Fixed-bed adsorption dynamics of Pb (II) adsorption from aqueous solution using nanostructured γ-alumina. J. Nanostruct. Chem. 2013, 3, 48. [Google Scholar] [CrossRef]
  210. Liu, X.; Niu, C.; Zhen, X.; Wang, J.; Su, X. Novel approach for synthesis of boehmite nanostructures and their conversion to aluminum oxide nanostructures for remove Congo red. J. Colloid Interface Sci. 2015, 452, 116–125. [Google Scholar] [CrossRef]
  211. Hossain, A.M.S.; Balbín, A.; Erami, R.S.; Prashar, S.; Fajardo, M.; Gómez-Ruiz, S. Synthesis and study of the catalytic applications in C–C coupling reactions of hybrid nanosystems based on alumina and palladium nanoparticles. Inorg. Chim. Acta 2017, 455, 645–652. [Google Scholar] [CrossRef]
  212. Tan, H.-b.; Guo, C.-s. Preparation of long alumina fibers by sol-gel method using malic acid. Trans. Nonferrous Met. Soc. China 2011, 21, 1563–1567. [Google Scholar] [CrossRef]
  213. Tzompantzi, F.; Piña, Y.; Mantilla, A.; Aguilar-Martínez, O.; Galindo-Hernández, F.; Bokhimi, X.; Barrera, A. Hydroxylated sol–gel Al2O3 as photocatalyst for the degradation of phenolic compounds in presence of UV light. Catal. Today 2014, 220–222, 49–55. [Google Scholar] [CrossRef]
  214. Piña-Pérez, Y.; Tzompantzi-Morales, F.; Pérez-Hernández, R.; Arroyo-Murillo, R.; Acevedo-Peña, P.; Gómez-Romero, R. Photocatalytic activity of Al2O3 improved by the addition of Ce3+/Ce4+ synthesized by the sol-gel method. Photodegradation of phenolic compounds using UV light. Fuel 2017, 198, 11–21. [Google Scholar] [CrossRef]
  215. Polat Gonullu, M. Design and characterization of single bilayer ZnO/Al2O3 film by ultrasonically spray pyrolysis and its application in photocatalysis. Superlattices Microstruct. 2021, 164, 107113. [Google Scholar] [CrossRef]
  216. Tan, L.; Dong, W.; Liu, K.; Luo, T.; Gu, X. Thermal decomposition in-situ preparation of gray rutile TiO2−x/Al2O3 composite and its enhanced visible-light-driven photocatalytic properties. Opt. Mater. 2021, 111, 110716. [Google Scholar] [CrossRef]
  217. Veksha, A.; Bin Mohamed Amrad, M.Z.; Chen, W.Q.; Binte Mohamed, D.K.; Tiwari, S.B.; Lim, T.-T.; Lisak, G. Tailoring Fe2O3–Al2O3 catalyst structure and activity via hydrothermal synthesis for carbon nanotubes and hydrogen production from polyolefin plastics. Chemosphere 2022, 297, 134148. [Google Scholar] [CrossRef] [PubMed]
  218. Chong, X.; Xiao, G.; Ding, D.; Luo, J.; Zheng, X. Combustion synthesis of SiC/Al2O3 composite powders with SiC nanowires and their growth mechanism. Ceram. Int. 2022, 48, 1778–1788. [Google Scholar] [CrossRef]
  219. Rahbar Shamskar, F.; Meshkani, F.; Rezaei, M. Ultrasound assisted co-precipitation synthesis and catalytic performance of mesoporous nanocrystalline NiO-Al2O3 powders. Ultrason. Sonochem. 2017, 34, 436–447. [Google Scholar] [CrossRef]
  220. Grishina, E.P.; Kudryakova, N.O.; Ramenskaya, L.M. Thin-film Al2O3 coating on low carbon steel obtained by the sol–gel method with different peptizing acids: Corrosion investigation. Thin Solid Film. 2022, 746, 139125. [Google Scholar] [CrossRef]
  221. Mirjalili, F.; Mohamad, H.; Luqman Chuah, A. Preparation of nano-scale alpha-Al2O3 powder by the sol-gel method. Ceram. Silik. 2011, 55, 378–383. [Google Scholar]
  222. Mavrič, A.; Valant, M.; Cui, C.; Wang, Z.M. Advanced applications of amorphous alumina: From nano to bulk. J. Non-Cryst. Solids 2019, 521, 119493. [Google Scholar] [CrossRef]
  223. Peintinger, M.F.; Kratz, M.J.; Bredow, T. Quantum-chemical study of stable, meta-stable and high-pressure alumina polymorphs and aluminum hydroxides. J. Mater. Chem. A 2014, 2, 13143–13158. [Google Scholar] [CrossRef]
  224. Mallakpour, S.; Khadem, E. Recent development in the synthesis of polymer nanocomposites based on nano-alumina. Prog. Polym. Sci. 2015, 51, 74–93. [Google Scholar] [CrossRef]
  225. Niero, D.F.; Montedo, O.R.K.; Bernardin, A.M. Synthesis and characterization of nano α-alumina by an inorganic sol–gel method. Mater. Sci. Eng. B 2022, 280, 115690. [Google Scholar] [CrossRef]
  226. Xiao, W.; Wang, D.; Lou, X.W. Shape-Controlled Synthesis of MnO2 Nanostructures with Enhanced Electrocatalytic Activity for Oxygen Reduction. J. Phys. Chem. C 2010, 114, 1694–1700. [Google Scholar] [CrossRef]
  227. Yan, D.; Yan, P.X.; Yue, G.H.; Liu, J.Z.; Chang, J.B.; Yang, Q.; Qu, D.M.; Geng, Z.R.; Chen, J.T.; Zhang, G.A.; et al. Self-assembled flower-like hierarchical spheres and nanobelts of manganese oxide by hydrothermal method and morphology control of them. Chem. Phys. Lett. 2007, 440, 134–138. [Google Scholar] [CrossRef]
  228. Rao, B.C.N.R.; Govindaraj, A.; Vivekchand, S.R.C. Inorganic nanomaterials: Current status and future prospects. Annu. Rep. Sect. A (Inorg. Chem.) 2006, 102, 20–45. [Google Scholar] [CrossRef]
  229. McKenzie, R.M. The synthesis of birnessite, cryptomelane, and some other oxides and hydroxides of manganese. Mineral. Mag. 2018, 38, 493–502. [Google Scholar] [CrossRef]
  230. Shih, Y.-J.; Huang, R.; Huang, Y.H. Adsorptive removal of arsenic using a novel akhtenskite coated waste Goethite. J. Clean. Prod. 2015, 87, 897–905. [Google Scholar] [CrossRef]
  231. Indra, A.; Menezes, P.W.; Schuster, F.; Driess, M. Significant role of Mn(III) sites in eg1 configuration in manganese oxide catalysts for efficient artificial water oxidation. J. Photochem. Photobiol. B Biol. 2015, 152, 156–161. [Google Scholar] [CrossRef]
  232. Shin, J.; Seo, J.K.; Yaylian, R.; Huang, A.; Meng, Y.S. A review on mechanistic understanding of MnO2 in aqueous electrolyte for electrical energy storage systems. Int. Mater. Rev. 2020, 65, 356–387. [Google Scholar] [CrossRef]
  233. Juran, T.R.; Young, J.; Smeu, M. Density Functional Theory Modeling of MnO2 Polymorphs as Cathodes for Multivalent Ion Batteries. J. Phys. Chem. C 2018, 122, 8788–8795. [Google Scholar] [CrossRef]
  234. Šťastný, M.; Stengl, V.; Henych, J.; Tolasz, J.; Vomáčka, P.; Ederer, J. Mesoporous manganese oxide for the degradation of organophosphates pesticides. J. Mater. Sci. 2016, 51, 2634–2642. [Google Scholar] [CrossRef]
  235. Ebrahimi, R.; Mohammadi, M.; Maleki, A.; Jafari, A.; Shahmoradi, B.; Rezaee, R.; Safari, M.; Daraei, H.; Giahi, O.; Yetilmezsoy, K.; et al. Photocatalytic Degradation of 2,4-Dichlorophenoxyacetic Acid in Aqueous Solution Using Mn-doped ZnO/Graphene Nanocomposite Under LED Radiation. J. Inorg. Organomet. Polym. Mater. 2020, 30, 923–934. [Google Scholar] [CrossRef]
  236. Zandsalimi, Y.; Maleki, A.; Shahmoradi, B.; Dehestani, S.; Rezaee, R.; McKay, G. Photocatalytic removal of 2,4-Dichlorophenoxyacetic acid from aqueous solution using tungsten oxide doped zinc oxide nanoparticles immobilised on glass beads. Environ. Technol. 2022, 43, 631–645. [Google Scholar] [CrossRef]
  237. Maleki, A.; Moradi, F.; Shahmoradi, B.; Rezaee, R.; Lee, S.-M. The photocatalytic removal of diazinon from aqueous solutions using tungsten oxide doped zinc oxide nanoparticles immobilized on glass substrate. J. Mol. Liq. 2020, 297, 111918. [Google Scholar] [CrossRef]
  238. Nagpal, M.; Kakkar, R. Use of metal oxides for the adsorptive removal of toxic organic pollutants. Sep. Purif. Technol. 2019, 211, 522–539. [Google Scholar] [CrossRef]
  239. Sheela, T.; Nayaka, Y.A. Kinetics and thermodynamics of cadmium and lead ions adsorption on NiO nanoparticles. Chem. Eng. J. 2012, 191, 123–131. [Google Scholar] [CrossRef]
  240. Ahmad Bhat, S.; Zafar, F.; Ullah Mirza, A.; Hossain Mondal, A.; Kareem, A.; Haq, Q.M.R.; Nishat, N. NiO nanoparticle doped-PVA-MF polymer nanocomposites: Preparation, Congo red dye adsorption and antibacterial activity. Arab. J. Chem. 2020, 13, 5724–5739. [Google Scholar] [CrossRef]
  241. Pooyandeh, S.; Shahidi, S.; Khajehnezhad, A.; Mongkholrattanasit, R. In situ deposition of NiO nano particles on cotton fabric using sol–gel method- photocatalytic activation properties. J. Mater. Res. Technol. 2021, 12, 1–14. [Google Scholar] [CrossRef]
  242. Ibrahim, E.M.M.; Abdel-Rahman, L.H.; Abu-Dief, A.M.; Elshafaie, A.; Hamdan, S.K.; Ahmed, A.M. The synthesis of CuO and NiO nanoparticles by facile thermal decomposition of metal-Schiff base complexes and an examination of their electric, thermoelectric and magnetic Properties. Mater. Res. Bull. 2018, 107, 492–497. [Google Scholar] [CrossRef]
  243. Mohandesi, M.; Tavakolian, M.; Rahimpour, M.R. Eggplant as an appreciable bio-template for green synthesis of NiO nanoparticles: Study of physical and photocatalytic properties. Ceram. Int. 2022, 48, 22820–22826. [Google Scholar] [CrossRef]
  244. Mostafa, A.M.; Mwafy, E.A. The effect of laser fluence for enhancing the antibacterial activity of NiO nanoparticles by pulsed laser ablation in liquid media. Environ. Nanotechnol. Monit. Manag. 2020, 14, 100382. [Google Scholar] [CrossRef]
  245. Kumar, J.P.; Giri, S.D.; Sarkar, A. Mesoporous NiO with different morphology: Synthesis, characterization and their evaluation for oxygen evolution reaction. Int. J. Hydrogen Energy 2018, 43, 15639–15649. [Google Scholar] [CrossRef]
  246. Hosny, N.M. Synthesis, characterization and optical band gap of NiO nanoparticles derived from anthranilic acid precursors via a thermal decomposition route. Polyhedron 2011, 30, 470–476. [Google Scholar] [CrossRef]
  247. Rashidimoghaddam, M.; Saljooqi, A.; Shamspur, T.; Mostafavi, A. Constructing S-doped Ni–Co LDH intercalated with Fe3O4 heterostructure photocatalysts for enhanced pesticide degradation. New J. Chem. 2020, 44, 15584–15592. [Google Scholar] [CrossRef]
  248. 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. [Google Scholar] [CrossRef]
  249. Hassaan, M.A.; El Nemr, A. Pesticides pollution: Classifications, human health impact, extraction and treatment techniques. Egypt. J. Aquat. Res. 2020, 46, 207–220. [Google Scholar] [CrossRef]
  250. Abubakar, Y.; Tijjani, H.; Egbuna, C.; Adetunji, C.O.; Kala, S.; Kryeziu, T.L.; Ifemeje, J.C.; Patrick-Iwuanyanwu, K.C. Chapter 3—Pesticides, History, and Classification. In Natural Remedies for Pest, Disease and Weed Control; Egbuna, C., Sawicka, B., Eds.; Academic Press: Cambridge, MA, USA, 2020; pp. 29–42. [Google Scholar] [CrossRef]
  251. Sousa, R.M.O.F.; Cunha, A.C.; Fernandes-Ferreira, M. The potential of Apiaceae species as sources of singular phytochemicals and plant-based pesticides. Phytochemistry 2021, 187, 112714. [Google Scholar] [CrossRef]
  252. George, D.; Finn, R.; Graham, K.; Sparagano, O. Present and future potential of plant-derived products to control arthropods of veterinary and medical significance. Parasites Vectors 2014, 7, 28. [Google Scholar] [CrossRef]
  253. Quijano, L.; Yusà, V.; Font, G.; Pardo, O. Chronic cumulative risk assessment of the exposure to organophosphorus, carbamate and pyrethroid and pyrethrin pesticides through fruit and vegetables consumption in the region of Valencia (Spain). Food Chem. Toxicol. 2016, 89, 39–46. [Google Scholar] [CrossRef]
  254. Pirsaheb, M.; Fattahi, N.; Rahimi, R.; Sharafi, K.; Ghaffari, H.R. Evaluation of abamectin, diazinon and chlorpyrifos pesticide residues in apple product of Mahabad region gardens: Iran in 2014. Food Chem. 2017, 231, 148–155. [Google Scholar] [CrossRef]
  255. Blankson, G.K.; Osei-Fosu, P.; Adeendze, E.A.; Ashie, D. Contamination levels of organophosphorus and synthetic pyrethroid pesticides in vegetables marketed in Accra, Ghana. Food Control 2016, 68, 174–180. [Google Scholar] [CrossRef]
  256. Yuan, Y.; Chen, C.; Zheng, C.; Wang, X.; Yang, G.; Wang, Q.; Zhang, Z. Residue of chlorpyrifos and cypermethrin in vegetables and probabilistic exposure assessment for consumers in Zhejiang Province, China. Food Control 2014, 36, 63–68. [Google Scholar] [CrossRef]
  257. Ciasca, B.; Pecorelli, I.; Lepore, L.; Paoloni, A.; Catucci, L.; Pascale, M.; Lattanzio, V.M.T. Rapid and reliable detection of glyphosate in pome fruits, berries, pulses and cereals by flow injection—Mass spectrometry. Food Chem. 2020, 310, 125813. [Google Scholar] [CrossRef] [PubMed]
  258. Mojsak, P.; Łozowicka, B.; Kaczyński, P. Estimating acute and chronic exposure of children and adults to chlorpyrifos in fruit and vegetables based on the new, lower toxicology data. Ecotoxicol. Environ. Saf. 2018, 159, 182–189. [Google Scholar] [CrossRef] [PubMed]
  259. Li, Z.; Zhang, Y.; Zhao, Q.; Wang, C.; Cui, Y.; Li, J.; Chen, A.; Liang, G.; Jiao, B. Occurrence, temporal variation, quality and safety assessment of pesticide residues on citrus fruits in China. Chemosphere 2020, 258, 127381. [Google Scholar] [CrossRef]
  260. Hirano, T.; Suzuki, N.; Ikenaka, Y.; Hoshi, N.; Tabuchi, Y. Neurotoxicity of a pyrethroid pesticide deltamethrin is associated with the imbalance in proteolytic systems caused by mitophagy activation and proteasome inhibition. Toxicol. Appl. Pharmacol. 2021, 430, 115723. [Google Scholar] [CrossRef]
  261. Nurdin, M.; Maulidiyah, M.; Salim, L.O.A.; Muzakkar, M.Z.; Umar, A.A. High performance cypermethrin pesticide detection using anatase TiO2-carbon paste nanocomposites electrode. Microchem. J. 2019, 145, 756–761. [Google Scholar] [CrossRef]
  262. Xia, Y.; Bian, Q.; Xu, L.; Cheng, S.; Song, L.; Liu, J.; Wu, W.; Wang, S.; Wang, X. Genotoxic effects on human spermatozoa among pesticide factory workers exposed to fenvalerate. Toxicology 2004, 203, 49–60. [Google Scholar] [CrossRef]
  263. Gupta, D.; Jamwal, D.; Rana, D.; Katoch, A. 26—Microwave synthesized nanocomposites for enhancing oral bioavailability of drugs. In Applications of Nanocomposite Materials in Drug Delivery; Inamuddin, A.M.A., Mohammad, A., Eds.; Woodhead Publishing: Cambridge, UK, 2018; pp. 619–632. [Google Scholar] [CrossRef]
  264. Dabrowski, A. Adsorption--from theory to practice. Adv Colloid Interface Sci 2001, 93, 135–224. [Google Scholar] [CrossRef]
  265. del Mar Orta, M.; Martín, J.; Medina-Carrasco, S.; Santos, J.L.; Aparicio, I.; Alonso, E. Adsorption of propranolol onto montmorillonite: Kinetic, isotherm and pH studies. Appl. Clay Sci. 2019, 173, 107–114. [Google Scholar] [CrossRef]
  266. Derylo-Marczewska, A.; Blachnio, M.; Marczewski, A.W.; Swiatkowski, A.; Buczek, B. Adsorption of chlorophenoxy pesticides on activated carbon with gradually removed external particle layers. Chem. Eng. J. 2017, 308, 408–418. [Google Scholar] [CrossRef]
  267. Sharma, L.; Kakkar, R. Hierarchically structured magnesium based oxides: Synthesis strategies and applications in organic pollutant remediation. CrystEngComm 2017, 19, 6913–6926. [Google Scholar] [CrossRef]
  268. Batzill, M. Fundamental aspects of surface engineering of transition metal oxide photocatalysts. Energy Environ. Sci. 2011, 4, 3275–3286. [Google Scholar] [CrossRef]
  269. Mahmoud, A.M.; Ibrahim, F.A.; Shaban, S.A.; Youssef, N.A. Adsorption of heavy metal ion from aqueous solution by nickel oxide nano catalyst prepared by different methods. Egypt. J. Pet. 2015, 24, 27–35. [Google Scholar] [CrossRef]
  270. Crini, G.; Lichtfouse, E.; Wilson, L.D.; Morin-Crini, N. Adsorption-Oriented Processes Using Conventional and Non-conventional Adsorbents for Wastewater Treatment. In Green Adsorbents for Pollutant Removal: Fundamentals and Design; Crini, G., Lichtfouse, E., Eds.; Springer International Publishing: Cham, Switzerland, 2018; pp. 23–71. [Google Scholar] [CrossRef]
  271. Králik, M. Adsorption, chemisorption, and catalysis. Chem. Pap. 2014, 68, 1625–1638. [Google Scholar] [CrossRef]
  272. Edet, U.A.; Ifelebuegu, A.O. Kinetics, Isotherms, and Thermodynamic Modeling of the Adsorption of Phosphates from Model Wastewater Using Recycled Brick Waste. Processes 2020, 8, 665. [Google Scholar] [CrossRef]
  273. Ahmed, M.J.; Hameed, B.H. Insights into the isotherm and kinetic models for the coadsorption of pharmaceuticals in the absence and presence of metal ions: A review. J. Environ. Manag. 2019, 252, 109617. [Google Scholar] [CrossRef]
  274. Adeola, A.O.; Abiodun, B.A.; Adenuga, D.O.; Nomngongo, P.N. Adsorptive and photocatalytic remediation of hazardous organic chemical pollutants in aqueous medium: A review. J. Contam. Hydrol. 2022, 248, 104019. [Google Scholar] [CrossRef]
  275. Salam, M.A.; AbuKhadra, M.R.; Mohamed, A.S. Effective oxidation of methyl parathion pesticide in water over recycled glass based-MCM-41 decorated by green Co3O4 nanoparticles. Environ. Pollut. 2020, 259, 113874. [Google Scholar] [CrossRef]
  276. Zhao, J.; Huang, P.; Wang, X.; Yang, J.; Zhou, Z.; Du, X.; Lu, X. Efficient adsorption removal of organic nitrogen pesticides: Insight into a new hollow NiO/Co@C magnetic nanocomposites derived from metal-organic framework. Sep. Purif. Technol. 2022, 287, 120608. [Google Scholar] [CrossRef]
  277. Abdelillah Ali Elhussein, E.; Şahin, S.; Bayazit, Ş.S. Preparation of CeO2 nanofibers derived from Ce-BTC metal-organic frameworks and its application on pesticide adsorption. J. Mol. Liq. 2018, 255, 10–17. [Google Scholar] [CrossRef]
  278. Zhou, D.-D.; Lu, Z.-H.; Chen, M.; Zhuang, L.-Y.; Cao, Y.-W.; Liu, X.; Senosy, I.A.; Yang, Z.-H. A novel magnetic double MOF composite is synthesized for removing azole fungicides economically and efficiently. Appl. Surf. Sci. 2022, 594, 153441. [Google Scholar] [CrossRef]
  279. Sharma, L.; Kakkar, R. Magnetically retrievable one-pot fabrication of mesoporous magnesium ferrite (MgFe2O4) for the remediation of chlorpyrifos and real pesticide wastewater. J. Environ. Chem. Eng. 2018, 6, 6891–6903. [Google Scholar] [CrossRef]
  280. Nayak, A.; Bhushan, B.; Sharma, P.K.; Gupta, V. Development of Magnetic Nanoparticles from Poplar Sawdust for Removal of Pesticides from Aqueous Solution. J. Graph. Era Univ. 2018, 6, 55–70. [Google Scholar]
  281. Kaur, Y.; Bhatia, Y.; Chaudhary, S.; Chaudhary, G.R. Comparative performance of bare and functionalize ZnO nanoadsorbents for pesticide removal from aqueous solution. J. Mol. Liq. 2017, 234, 94–103. [Google Scholar] [CrossRef]
  282. Chen, D.; Chen, C.; Shen, W.; Quan, H.; Chen, S.; Xie, S.; Luo, X.; Guo, L. MOF-derived magnetic porous carbon-based sorbent: Synthesis, characterization, and adsorption behavior of organic micropollutants. Adv. Powder Technol. 2017, 28, 1769–1779. [Google Scholar] [CrossRef]
  283. Cao, X.; Liu, G.; She, Y.; Jiang, Z.; Jin, F.; Jin, M.; Du, P.; Zhao, F.; Zhang, Y.; Wang, J. Preparation of magnetic metal organic framework composites for the extraction of neonicotinoid insecticides from environmental water samples. RSC Adv. 2016, 6, 113144–113151. [Google Scholar] [CrossRef]
  284. Su, H.; Lin, Y.; Wang, Z.; Wong, Y.L.; Chen, X.; Chan, T.W. Magnetic metal-organic framework-titanium dioxide nanocomposite as adsorbent in the magnetic solid-phase extraction of fungicides from environmental water samples. J. Chromatogr. A 2016, 1466, 21–28. [Google Scholar] [CrossRef]
  285. Liu, G.; Li, L.; Huang, X.; Zheng, S.; Xu, D.; Xu, X.; Zhang, Y.; Lin, H. Determination of triazole pesticides in aqueous solution based on magnetic graphene oxide functionalized MOF-199 as solid phase extraction sorbents. Microporous Mesoporous Mater. 2018, 270, 258–264. [Google Scholar] [CrossRef]
  286. Liu, G.; Li, L.; Huang, X.; Zheng, S.; Xu, X.; Liu, Z.; Zhang, Y.; Wang, J.; Lin, H.; Xu, D. Adsorption and removal of organophosphorus pesticides from environmental water and soil samples by using magnetic multi-walled carbon nanotubes @ organic framework ZIF-8. J. Mater. Sci. 2018, 53, 10772–10783. [Google Scholar] [CrossRef]
  287. Badawy, M.E.I.; El-Nouby, M.A.M.; Marei, A.E.-S.M. Development of a Solid-Phase Extraction (SPE) Cartridge Based on Chitosan-Metal Oxide Nanoparticles (Ch-MO NPs) for Extraction of Pesticides from Water and Determination by HPLC. Int. J. Anal. Chem. 2018, 2018, 3640691. [Google Scholar] [CrossRef] [PubMed]
  288. Moradi Dehaghi, S.; Rahmanifar, B.; Moradi, A.M.; Azar, P.A. Removal of permethrin pesticide from water by chitosan–zinc oxide nanoparticles composite as an adsorbent. J. Saudi Chem. Soc. 2014, 18, 348–355. [Google Scholar] [CrossRef]
  289. Parsaie, A.; Rahbar, N.; Baezat, M. A New Fe3O4/CuO/AC Nanocomposite for Imidacloprid Removal: Characterization, Optimization, and Adsorption Modeling. Res. Sq. 2021. [Google Scholar] [CrossRef]
  290. ul Haq, A.; Saeed, M.; Usman, M.; Naqvi, S.A.R.; Bokhari, T.H.; Maqbool, T.; Ghaus, H.; Tahir, T.; Khalid, H. Sorption of chlorpyrifos onto zinc oxide nanoparticles impregnated Pea peels (Pisum sativum L): Equilibrium, kinetic and thermodynamic studies. Environ. Technol. Innov. 2020, 17, 100516. [Google Scholar] [CrossRef]
  291. Haq, A.; Saeed, M.; Muneer, M.; Jamal, M.; Maqbool, T.; Tahir, T. Biosorption of metribuzin pesticide by Cucumber (Cucumis sativus) peels-zinc oxide nanoparticles composite. Sci. Rep. 2022, 12, 5840. [Google Scholar] [CrossRef]
  292. Cusioli, L.F.; Quesada, H.B.; Barbosa de Andrade, M.; Gomes, R.G.; Bergamasco, R. Application of a novel low-cost adsorbent functioned with iron oxide nanoparticles for the removal of triclosan present in contaminated water. Microporous Mesoporous Mater. 2021, 325, 111328. [Google Scholar] [CrossRef]
  293. Allam, E.A.; Ali, A.S.M.; Elsharkawy, R.M.; Mahmoud, M.E. Framework of nano metal oxides N-NiO@N-Fe3O4@N-ZnO for adsorptive removal of atrazine and bisphenol-A from wastewater: Kinetic and adsorption studies. Environ. Nanotechnol. Monit. Manag. 2021, 16, 100481. [Google Scholar] [CrossRef]
  294. Rao, T.N.; Balaji, A.P.B.; Panagal, M.; Parvatamma, B.; Selvaraj, B.; Panneerselvam, S.; Aruni, W.; Subramanian, K.; Sampath Renuga, P.; Pandian, S. Nanoremediation of dimethomorph in water samples using magnesium aluminate nanoparticles. Environ. Technol. Innov. 2020, 20, 101176. [Google Scholar] [CrossRef]
  295. Lan, J.; Cheng, Y.; Zhao, Z. Effective organochlorine pesticides removal from aqueous systems by magnetic nanospheres coated with polystyrene. J. Wuhan Univ. Technol.-Mater. Sci. Ed. 2014, 29, 168–173. [Google Scholar] [CrossRef]
  296. Armaghan, M.; Amini, M.M. Adsorption of diazinon and fenitrothion on nanocrystalline magnesium oxides. Arab. J. Chem. 2017, 10, 91–99. [Google Scholar] [CrossRef]
  297. Liu, F.; Tian, H.; He, J. Adsorptive performance and catalytic activity of superparamagnetic Fe3O4@nSiO2@mSiO2 core–shell microspheres towards DDT. J. Colloid Interface Sci. 2014, 419, 68–72. [Google Scholar] [CrossRef] [PubMed]
  298. Vaya, D.; Surolia, P.K. Semiconductor based photocatalytic degradation of pesticides: An overview. Environ. Technol. Innov. 2020, 20, 101128. [Google Scholar] [CrossRef]
  299. Krishnasamy, L.; Krishna, K.; Subpiramaniyam, S. Photocatalytic Degradation of Atrazine in Aqueous Solution Using La-Doped ZnO/PAN Nanofibers. Environ. Sci. Pollut. Res. 2022, 29, 54282–54291. [Google Scholar] [CrossRef]
  300. Rani, M.; Yadav, J.; Keshu; Shanker, U. Green synthesis of sunlight responsive zinc oxide coupled cadmium sulfide nanostructures for efficient photodegradation of pesticides. J. Colloid Interface Sci. 2021, 601, 689–703. [Google Scholar] [CrossRef]
  301. Zhu, Z.; Guo, F.; Xu, Z.; Di, X.; Zhang, Q. Photocatalytic degradation of an organophosphorus pesticide using a ZnO/rGO composite. RSC Adv. 2020, 10, 11929–11938. [Google Scholar] [CrossRef]
  302. Anirudhan, T.S.; Shainy, F.; Sekhar, V.C.; Athira, V.S. Highly efficient photocatalytic degradation of chlorpyrifos in aqueous solutions by nano hydroxyapatite modified CFGO/ZnO nanorod composite. J. Photochem. Photobiol. A Chem. 2021, 418, 113333. [Google Scholar] [CrossRef]
  303. Soltani-nezhad, F.; Saljooqi, A.; Shamspur, T.; Mostafavi, A. Photocatalytic degradation of imidacloprid using GO/Fe3O4/TiO2-NiO under visible radiation: Optimization by response level method. Polyhedron 2019, 165, 188–196. [Google Scholar] [CrossRef]
  304. Zangiabadi, M.; Shamspur, T.; Saljooqi, A.; Mostafavi, A. Evaluating the efficiency of the GO-Fe3O4/TiO2 mesoporous photocatalyst for degradation of chlorpyrifos pesticide under visible light irradiation. Appl. Organomet. Chem. 2019, 33, e4813. [Google Scholar] [CrossRef]
  305. Andronic, L.; Vladescu, A.; Enesca, A. Synthesis, Characterisation, Photocatalytic Activity, and Aquatic Toxicity Evaluation of TiO2 Nanoparticles. Nanomaterials 2021, 11, 3197. [Google Scholar] [CrossRef] [PubMed]
  306. Čižmar, T.; Panžić, I.; Capan, I.; Gajović, A. Nanostructured TiO2 photocatalyst modified with Cu for improved imidacloprid degradation. Appl. Surf. Sci. 2021, 569, 151026. [Google Scholar] [CrossRef]
  307. Khan, S.H.; Pathak, B. Zinc oxide based photocatalytic degradation of persistent pesticides: A comprehensive review. Environ. Nanotechnol. Monit. Manag. 2020, 13, 100290. [Google Scholar] [CrossRef]
  308. AbuKhadra, M.R.; Mohamed, A.S.; El-Sherbeeny, A.M.; Elmeligy, M.A. Enhanced photocatalytic degradation of acephate pesticide over MCM-41/Co3O4 nanocomposite synthesized from rice husk silica gel and Peach leaves. J. Hazard. Mater. 2020, 389, 122129. [Google Scholar] [CrossRef]
  309. Mohamed, A.S.; Abukkhadra, M.R.; Abdallah, E.A.; El-Sherbeeny, A.M.; Mahmoud, R.K. The photocatalytic performance of silica fume based Co3O4/MCM-41 green nanocomposite for instantaneous degradation of Omethoate pesticide under visible light. J. Photochem. Photobiol. A Chem. 2020, 392, 112434. [Google Scholar] [CrossRef]
  310. Hanh, N.T.; Le Minh Tri, N.; Van Thuan, D.; Thanh Tung, M.H.; Pham, T.-D.; Minh, T.D.; Trang, H.T.; Binh, M.T.; Nguyen, M.V. Monocrotophos pesticide effectively removed by novel visible light driven Cu doped ZnO photocatalyst. J. Photochem. Photobiol. A Chem. 2019, 382, 111923. [Google Scholar] [CrossRef]
  311. Kumari, V.; Sharma, S.; Sharma, A.; Kumari, K.; Kumar, N. Hydrothermal synthesis conditions effect on hierarchical ZnO/CuO hybrid materials and their photocatalytic activity. J. Mater. Sci. Mater. Electron. 2021, 32, 9596–9610. [Google Scholar] [CrossRef]
  312. Iqbal, A.; Haq, A.u.; Cerrón-Calle, G.A.; Naqvi, S.A.R.; Westerhoff, P.; Garcia-Segura, S. Green synthesis of flower-shaped copper oxide and nickel oxide nanoparticles via capparis decidua leaf extract for synergic adsorption-photocatalytic degradation of pesticides. Catalysts 2021, 11, 806. [Google Scholar] [CrossRef]
  313. Aguilera-Ruiz, E.; De La Garza-Galván, M.; Zambrano-Robledo, P.; Ballesteros-Pacheco, J.C.; Vazquez-Arenas, J.; Peral, J.; García-Pérez, U.M. Facile synthesis of visible-light-driven Cu2O/BiVO4 composites for the photomineralization of recalcitrant pesticides. RSC Adv. 2017, 7, 45885–45895. [Google Scholar] [CrossRef]
  314. Bharti; Jangwan, J.S.; Kumar, G.; Kumar, V.; Kumar, A. Abatement of organic and inorganic pollutants from drinking water by using commercial and laboratory-synthesized zinc oxide nanoparticles. SN Appl. Sci. 2021, 3, 311. [Google Scholar] [CrossRef]
  315. Sharma, A.K.; Tiwari, R.K.; Gaur, M.S. Nanophotocatalytic UV degradation system for organophosphorus pesticides in water samples and analysis by Kubista model. Arab. J. Chem. 2016, 9, S1755–S1764. [Google Scholar] [CrossRef]
  316. Anirudhan, T.S.; Shainy, F.; Manasa Mohan, A. Fabrication of zinc oxide nanorod incorporated carboxylic graphene/polyaniline composite and its photocatalytic activity for the effective degradation of diuron from aqueous solutions. Sol. Energy 2018, 171, 534–546. [Google Scholar] [CrossRef]
  317. Khan, S.H.; Pathak, B.; Fulekar, M.H. Synthesis, characterization and photocatalytic degradation of chlorpyrifos by novel Fe: ZnO nanocomposite material. Nanotechnol. Environ. Eng. 2018, 3, 13. [Google Scholar] [CrossRef]
  318. Choudhary, M.K.; Kataria, J.; Bhardwaj, V.K.; Sharma, S. Green biomimetic preparation of efficient Ag–ZnO heterojunctions with excellent photocatalytic performance under solar light irradiation: A novel biogenic-deposition-precipitation approach. Nanoscale Adv. 2019, 1, 1035–1044. [Google Scholar] [CrossRef]
  319. Yari, K.; Seidmohammadi, A.; Khazaei, M.; Bhatnagar, A.; Leili, M. A comparative study for the removal of imidacloprid insecticide from water by chemical-less UVC, UVC/TiO2 and UVC/ZnO processes. J. Environ. Health Sci. Eng. 2019, 17, 337–351. [Google Scholar] [CrossRef] [PubMed]
  320. Dehghan, S.; Jafari, A.J.; FarzadKia, M.; Esrafili, A.; Kalantary, R.R. Visible-light-driven photocatalytic degradation of Metalaxyl by reduced graphene oxide/Fe3O4/ZnO ternary nanohybrid: Influential factors, mechanism and toxicity bioassay. J. Photochem. Photobiol. A Chem. 2019, 375, 280–292. [Google Scholar] [CrossRef]
  321. Lakshmi, K.; Kadirvelu, K.; Mohan, P.S. Chemically modified electrospun nanofiber for high adsorption and effective photocatalytic decontamination of organophosphorus compounds. J. Chem. Technol. Biotechnol. 2019, 94, 3190–3200. [Google Scholar] [CrossRef]
  322. Dehghan, S.; Tahergorabi, M.; Norzaee, S.; Boorboor Azimi, E.; Hasham Firooz, M.; Dadban Shahamat, Y. Preparation and photocatalytic performance of reduced graphene oxide/ZnO nanocatalyst for degradation of metalaxyl from aqueous solution: Effect of operational parameters, mineralisation and toxicity bioassay. Int. J. Environ. Anal. Chem. 2022, 102, 7112–7134. [Google Scholar] [CrossRef]
  323. Aulakh, M.K.; Kaur, S.; Pal, B.; Singh, S. Morphological influence of ZnO nanostructures and their Cu loaded composites for effective photodegradation of methyl parathion. Solid State Sci. 2020, 99, 106045. [Google Scholar] [CrossRef]
  324. Kumari, V.; Yadav, S.; Mittal, A.; Sharma, S.; Kumari, K.; Kumar, N. Hydrothermally synthesized nano-carrots ZnO with CeO2 heterojunctions and their photocatalytic activity towards different organic pollutants. J. Mater. Sci. Mater. Electron. 2020, 31, 5227–5240. [Google Scholar] [CrossRef]
  325. Serrano-Lázaro, A.; Verdín-Betancourt, F.A.; Jayaraman, V.K.; López-González, M.d.L.; Hernández-Gordillo, A.; Sierra-Santoyo, A.; Bizarro, M. Efficient photocatalytic elimination of Temephos pesticide using ZnO nanoflowers. J. Photochem. Photobiol. A Chem. 2020, 393, 112414. [Google Scholar] [CrossRef]
  326. Javadimanesh, L.; Khazaei, M.; Leili, M.; Samarghandi, M.R. Using Fe/Ag nanostructures covered by Zn for the degradation of 2, 4 dichlorophenoxyacetic acid assisting with light irradiation and ozonation; optimisation and kinetic studies. Int. J. Environ. Anal. Chem. 2021, 1–14. [Google Scholar] [CrossRef]
  327. Saljooqi, A.; Shamspur, T.; Mostafavi, A. Synthesis and photocatalytic activity of porous ZnO stabilized by TiO2 and Fe3O4 nanoparticles: Investigation of pesticide degradation reaction in water treatment. Environ. Sci. Pollut. Res. 2021, 28, 9146–9156. [Google Scholar] [CrossRef] [PubMed]
  328. Adabavazeh, H.; Saljooqi, A.; Shamspur, T.; Mostafavi, A. Synthesis of polyaniline decorated with ZnO and CoMoO4 nanoparticles for enhanced photocatalytic degradation of imidacloprid pesticide under visible light. Polyhedron 2021, 198, 115058. [Google Scholar] [CrossRef]
  329. Veerakumar, P.; Sangili, A.; Saranya, K.; Pandikumar, A.; Lin, K.-C. Palladium and silver nanoparticles embedded on zinc oxide nanostars for photocatalytic degradation of pesticides and herbicides. Chem. Eng. J. 2021, 410, 128434. [Google Scholar] [CrossRef]
  330. Pathania, D.; Sharma, A.; Kumar, S.; Srivastava, A.K.; Kumar, A.; Singh, L. Bio-synthesized Cu–ZnO hetro-nanostructure for catalytic degradation of organophosphate chlorpyrifos under solar illumination. Chemosphere 2021, 277, 130315. [Google Scholar] [CrossRef]
  331. Yadav, S.; Jindal, J.; Mittal, A.; Sharma, S.; Kumari, K.; Kumar, N. Facile solution combustion synthesized, Li doped ZnO nanostructures for removal of abiotic contaminants. J. Phys. Chem. Solids 2021, 157, 110217. [Google Scholar] [CrossRef]
  332. Ahmadifard, T.; Heydari, R.; Tarrahi, M.J.; Khorramabadi, G.S. Photocatalytic degradation of diazinon in aqueous solutions using immobilized MgO nanoparticles on concrete. Int. J. Chem. React. Eng. 2019, 17, 20180154. [Google Scholar] [CrossRef]
  333. Mansourian, R.; Mousavi, S.M.; Alizadeh, S.; Sabbaghi, S. CeO2/TiO2/SiO2 nanocatalyst for the photocatalytic and sonophotocatalytic degradation of chlorpyrifos. Can. J. Chem. Eng. 2022, 100, 451–464. [Google Scholar] [CrossRef]
  334. Farrukh, M.A.; Butt, K.M.; Altaf, A. Influence of pH and temperature on structural, optical and catalytical investigations of CeO2-SiO2 nanoparticles. Silicon 2019, 11, 2591–2598. [Google Scholar] [CrossRef]
  335. Farrukh, M.A.; Butt, K.M.; Chong, K.-K.; Chang, W.S. Photoluminescence emission behavior on the reduced band gap of Fe doping in CeO2-SiO2 nanocomposite and photophysical properties. J. Saudi Chem. Soc. 2019, 23, 561–575. [Google Scholar] [CrossRef]
  336. Fouad, D.M.; El-Said, W.A.; Mohamed, M.B. Spectroscopic characterization of magnetic Fe3O4@Au core shell nanoparticles. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 140, 392–397. [Google Scholar] [CrossRef] [PubMed]
  337. Adabavazeh, H.; Saljooqi, A.; Shamspur, T.; Mostafavi, A. Synthesis of KIT-5 decorated by Bi2S3-Fe3O4 photocatalyst for degradation of parathion pesticide in aqueous media: Offering a degradation model and optimization using response surface methodology (RSM). Appl. Organomet. Chem. 2020, 34, e5345. [Google Scholar] [CrossRef]
  338. Merci, S.; Saljooqi, A.; Shamspur, T.; Mostafavi, A. Investigation of photocatalytic chlorpyrifos degradation by a new silica mesoporous material immobilized by WS2 and Fe3O4 nanoparticles: Application of response surface methodology. Appl. Organomet. Chem. 2020, 34, e5343. [Google Scholar] [CrossRef]
  339. Soltani-Nezhad, F.; Saljooqi, A.; Mostafavi, A.; Shamspur, T. Synthesis of Fe3O4/CdS–ZnS nanostructure and its application for photocatalytic degradation of chlorpyrifos pesticide and brilliant green dye from aqueous solutions. Ecotoxicol. Environ. Saf. 2020, 189, 109886. [Google Scholar] [CrossRef] [PubMed]
  340. Lima, M.S.; Cruz-Filho, J.F.; Noleto, L.F.; Silva, L.J.; Costa, T.M.; Luz, G.E., Jr. Synthesis, characterization and catalytic activity of Fe3O4@ WO3/SBA-15 on photodegradation of the acid dichlorophenoxyacetic (2, 4-D) under UV irradiation. J. Environ. Chem. Eng. 2020, 8, 104145. [Google Scholar] [CrossRef]
  341. Saljooqi, A.; Shamspur, T.; Mostafavi, A. Synthesis of titanium nanoplate decorated by Pd and Fe3O4 nanoparticles immobilized on graphene oxide as a novel photocatalyst for degradation of parathion pesticide. Polyhedron 2020, 179, 114371. [Google Scholar] [CrossRef]
  342. Cao, L.; Ma, D.; Zhou, Z.; Xu, C.; Cao, C.; Zhao, P.; Huang, Q. Efficient photocatalytic degradation of herbicide glyphosate in water by magnetically separable and recyclable BiOBr/Fe3O4 nanocomposites under visible light irradiation. Chem. Eng. J. 2019, 368, 212–222. [Google Scholar] [CrossRef]
  343. Shi, E.; Xu, Z.; Wang, W.; Xu, Y.; Zhang, Y.; Yang, X.; Liu, Q.; Zeng, T.; Song, S.; Jiang, Y.; et al. Ag2S-doped core-shell nanostructures of Fe3O4@Ag3PO4 ultrathin film: Major role of hole in rapid degradation of pollutants under visible light irradiation. Chem. Eng. J. 2019, 366, 123–132. [Google Scholar] [CrossRef]
  344. Nekooie, R.; Ghasemi, J.B.; Badiei, A.; Shamspur, T.; Mostafavi, A.; Moradian, S. Design and synthesis of g-C3N4/(Cu/TiO2) nanocomposite for the visible light photocatalytic degradation of endosulfan in aqueous solutions. J. Mol. Struct. 2022, 1258, 132650. [Google Scholar] [CrossRef]
  345. Ebrahimpour, M.; Hassaninejad-Darzi, S.K.; Mousavi, H.Z.; Samadi-Maybodi, A. Simultaneous monitoring of the photocatalytic degradation process of trifluralin and pendimethalin herbicides by SBA-15/TiO2 nanocomposite. Environ. Nanotechnol. Monit. Manag. 2022, 18, 100678. [Google Scholar] [CrossRef]
  346. de la Flor, M.P.; Camarillo, R.; Martínez, F.; Jiménez, C.; Quiles, R.; Rincón, J. Synthesis and characterization of TiO2/CNT/Pd: An effective sunlight photocatalyst for neonicotinoids degradation. J. Environ. Chem. Eng. 2021, 9, 106278. [Google Scholar] [CrossRef]
  347. Massoud, A.; Derbalah, A.; El-Mehasseb, I.; Allah, M.S.; Ahmed, M.S.; Albrakati, A.; Elmahallawy, E.K. Photocatalytic Detoxification of Some Insecticides in Aqueous Media Using TiO2 Nanocatalyst. Int. J. Environ. Res. Public Health 2021, 18, 9278. [Google Scholar] [CrossRef]
  348. Valadez-Renteria, E.; Barrera-Rendon, E.; Oliva, J.; Rodriguez-Gonzalez, V. Flexible CuS/TiO2 based composites made with recycled bags and polystyrene for the efficient removal of the 4-CP pesticide from drinking water. Sep. Purif. Technol. 2021, 270, 118821. [Google Scholar] [CrossRef]
  349. Ben Saber, N.; Mezni, A.; Alrooqi, A.; Altalhi, T. Ternary Pt@TiO2/rGO Nanocomposite to Boost Photocatalytic Activity for Environmental and Energy Use. J. Inorg. Organomet. Polym. Mater. 2021, 31, 3802–3809. [Google Scholar] [CrossRef]
  350. Hasanin, M.; Abdelhameed, R.M.; Dacrory, S.; Abou-Yousef, H.; Kamel, S. Photocatalytic degradation of pesticide intermediate using green eco-friendly amino functionalized cellulose nanocomposites. Mater. Sci. Eng. B 2021, 270, 115231. [Google Scholar] [CrossRef]
  351. Vanichvattanadecha, C.; Jaroenworaluck, A.; Henpraserttae, P.; Wimuktiwan, P.; Manpetch, P.; Singhapong, W. Ordered mesoporous silica (SBA-16) supporting titania (TiO2) nanoparticles for photodegradation of paraquat (PQ) herbicide. J. Porous Mater. 2021, 28, 1137–1153. [Google Scholar] [CrossRef]
  352. Behera, L.; Barik, B.; Mohapatra, S. Improved photodegradation and antimicrobial activity of hydrothermally synthesized 0.2Ce-TiO2/RGO under visible light. Colloids Surf. A Physicochem. Eng. Asp. 2021, 620, 126553. [Google Scholar] [CrossRef]
  353. Amiri, F.; Dehghani, M.; Amiri, Z.; Yousefinejad, S.; Azhdarpoor, A. Photocatalytic degradation of 2,4-dichlorophenoxyacetic acid from aqueous solutions by Ag3PO4/TiO2 nanoparticles under visible light: Kinetic and thermodynamic studies. Water Sci. Technol. 2021, 83, 3110–3122. [Google Scholar] [CrossRef]
  354. Vigneshwaran, S.; Sirajudheen, P.; Karthikeyan, P.; Nikitha, M.; Ramkumar, K.; Meenakshi, S. Immobilization of MIL-88(Fe) anchored TiO2-chitosan(2D/2D) hybrid nanocomposite for the degradation of organophosphate pesticide: Characterization, mechanism and degradation intermediates. J. Hazard. Mater. 2021, 406, 124728. [Google Scholar] [CrossRef]
  355. Meriam Suhaimy, S.H.; Lai, C.W.; Tajuddin, H.A.; Samsudin, E.M.; Johan, M.R. Impact of TiO2 Nanotubes’ Morphology on the Photocatalytic Degradation of Simazine Pollutant. Materials 2018, 11, 2066. [Google Scholar] [CrossRef] [PubMed]
  356. Joice, J.A.I.; Aishwarya, S.; Sivakumar, T. Nano Structured Ni and Ru Impregnated TiO2 Photocatalysts: Synthesis, Characterization and Photocatalytic Degradation of Neonicotinoid Insecticides. J. Nanosci. Nanotechnol. 2019, 19, 2575–2589. [Google Scholar] [CrossRef] [PubMed]
  357. Ahmari, H.; Heris, S.Z.; Khayyat, M.H. The effect of titanium dioxide nanoparticles and UV irradiation on photocatalytic degradation of Imidaclopride. Environ. Technol. 2018, 39, 536–547. [Google Scholar] [CrossRef] [PubMed]
  358. Eydivand, S.; Nikazar, M. Degradation of 1,2-Dichloroethane in Simulated Wastewater Solution: A Comprehensive Study by Photocatalysis Using TiO2 and ZnO Nanoparticles. Chem. Eng. Commun. 2015, 202, 102–111. [Google Scholar] [CrossRef]
  359. Ananpattarachai, J.; Kajitvichyanukul, P. Photocatalytic degradation of p,p′-DDT under UV and visible light using interstitial N-doped TiO2. J. Environ. Sci. Health Part B 2015, 50, 247–260. [Google Scholar] [CrossRef]
  360. Lopes Colpani, G.; Zanetti, M.; Zeferino, R.; Silva, L.; Mello, J.; Riella, H.; Padoin, N.; Fiori, M.; Soares, C. Lanthanides Effects on TiO2 Photocatalysts; IntechOpen: London, UK, 2018. [Google Scholar] [CrossRef]
Figure 1. Spinel structure of Co3O4 with Co (II) at the tetrahedral sites and Co(III) at the octahedral sites. Reproduced with permission from [44].
Figure 1. Spinel structure of Co3O4 with Co (II) at the tetrahedral sites and Co(III) at the octahedral sites. Reproduced with permission from [44].
Sustainability 15 07336 g001
Figure 2. The different structures of copper oxides: (a) cubic Cu2O, (b) monoclinic CuO, and (c) tetragonal Cu4O3. Reproduced with permission from [57].
Figure 2. The different structures of copper oxides: (a) cubic Cu2O, (b) monoclinic CuO, and (c) tetragonal Cu4O3. Reproduced with permission from [57].
Sustainability 15 07336 g002
Figure 3. Difference between an n-type semiconductor and a p-type semiconductor.
Figure 3. Difference between an n-type semiconductor and a p-type semiconductor.
Sustainability 15 07336 g003
Figure 4. Polymorphs of zinc oxide: (a) zinc blende, (b) hexagonal wurtzite, and (c) cubic rock-salt structures. Reproduced with permission from [94].
Figure 4. Polymorphs of zinc oxide: (a) zinc blende, (b) hexagonal wurtzite, and (c) cubic rock-salt structures. Reproduced with permission from [94].
Sustainability 15 07336 g004
Figure 5. Forms of Iron oxides: (a) wüstite, (b) magnetite, and (c) Ferric oxide (hematite phase). Reproduced with permission from [96].
Figure 5. Forms of Iron oxides: (a) wüstite, (b) magnetite, and (c) Ferric oxide (hematite phase). Reproduced with permission from [96].
Sustainability 15 07336 g005
Figure 6. Polymorphs of TiO2. Reproduced with permission from [123].
Figure 6. Polymorphs of TiO2. Reproduced with permission from [123].
Sustainability 15 07336 g006
Figure 7. SEM images for TiO2 nanostructures synthesized at (a) 140 °C rose-like structure, (b) 170 °C chrysanthemum- like structure, and (c) 200 °C sea-urchin-like structure. Reproduced with permission from [129].
Figure 7. SEM images for TiO2 nanostructures synthesized at (a) 140 °C rose-like structure, (b) 170 °C chrysanthemum- like structure, and (c) 200 °C sea-urchin-like structure. Reproduced with permission from [129].
Sustainability 15 07336 g007
Figure 8. Rock salt-type structure of MgO.
Figure 8. Rock salt-type structure of MgO.
Sustainability 15 07336 g008
Figure 9. Cubic fluorite-type structure of CeO2.
Figure 9. Cubic fluorite-type structure of CeO2.
Sustainability 15 07336 g009
Figure 10. Rhombohedral α-Al2O3 structure. Reproduced with permission from [223].
Figure 10. Rhombohedral α-Al2O3 structure. Reproduced with permission from [223].
Sustainability 15 07336 g010
Figure 11. MnO2 polymorphs. Reprinted with permission from [233]. Copyright @2018 American Chemical Society.
Figure 11. MnO2 polymorphs. Reprinted with permission from [233]. Copyright @2018 American Chemical Society.
Sustainability 15 07336 g011
Figure 12. The effect of different dopant percentages of WO3ZnO on the photodegradation efficiency of diazinon. Reproduced with permission from [237].
Figure 12. The effect of different dopant percentages of WO3ZnO on the photodegradation efficiency of diazinon. Reproduced with permission from [237].
Sustainability 15 07336 g012
Figure 13. SEM images of NiO nanostructures synthesized with different temperatures and molar ratios of precursors; (A) NiO nanorods and nanoplates, (B) NiO nanoplates, and (C) NiO NPs. Reproduced with permission from [245].
Figure 13. SEM images of NiO nanostructures synthesized with different temperatures and molar ratios of precursors; (A) NiO nanorods and nanoplates, (B) NiO nanoplates, and (C) NiO NPs. Reproduced with permission from [245].
Sustainability 15 07336 g013
Figure 14. The chemical structures of some pesticides: (a) Pyrethroids, (b) Organophosphates, (c) Carbamates, and (d) Organochlorines.
Figure 14. The chemical structures of some pesticides: (a) Pyrethroids, (b) Organophosphates, (c) Carbamates, and (d) Organochlorines.
Sustainability 15 07336 g014
Figure 15. A schematic mechanism for the photodegradation of a pesticide.
Figure 15. A schematic mechanism for the photodegradation of a pesticide.
Sustainability 15 07336 g015
Table 1. Classification of organic pesticides based on origin.
Table 1. Classification of organic pesticides based on origin.
OriginSourceClassExampleFeaturesRefs.
OrganicNaturalPlants PhytochemicalsEssential oil, plant extracts, and leftover oilseed cakes.Low Toxicity, limited persistence in the environment, and complicated structures that prevent resistance in pests.[251,252]
SyntheticPyrethroidsPhenthion,
Diazinon,
Cypermethrin, Deltamethrin, Cyfluthrin, and Cypermethrin
Effect the sodium channel in insects, resulting in paralysis of the organism; highly toxic to insects and fish but less to mammals; unstable upon the exposure of light; and commonly used in food.[253,254,255,256]
OrganophosphatesAldrin, Dieldrin, Glyphosate, and Chlorpyrifos. Cause paralysis, resulting in death, and dominant for variety of pests.[257,258]
CarbamatesFenvalerate, Permethrin, Cyhalothrin, and Carbofuran. Effect the nerve system of the pests, resulting in poisoning and death, and low pollution is caused upon degradation.[259,260,261,262]
OrganochlorineChlorothalonil and Endrin Aldehyde.Used for insects, long persistent in environment, affecting the nerve system and causing paralysis and death of the pests.
Table 2. Classification of pesticides based on target.
Table 2. Classification of pesticides based on target.
ClassTarget PestsExampleChemical StructuresRef.
AcaricidesMitesBifonazoleSustainability 15 07336 i001[250]
AvicidesBirdsAvitrolSustainability 15 07336 i002
FungicidesFungiAzoxystrobinSustainability 15 07336 i003
HerbicidesWeedsAtrazineSustainability 15 07336 i004
InsecticidesInsectsAldicarbSustainability 15 07336 i005
LarvicidesLarvaeMethopreneSustainability 15 07336 i006
MolluscicidesSnailMetaldehydeSustainability 15 07336 i007
NematicidesNematodesAldicarbSustainability 15 07336 i008
OvicidesEgg (prevents hatching of eggs in insects and mites)BenzoxazineSustainability 15 07336 i009
PiscicidesFishesRotenoneSustainability 15 07336 i010
RepellentsInsectsMethiocarbSustainability 15 07336 i011
RodenticidesRodentsWarfarinSustainability 15 07336 i012
TermiticidesKills termitesFipronilSustainability 15 07336 i013
ViricidesVirusesScytovirin
Table 3. The advantages and disadvantages of chemisorption and physisorption.
Table 3. The advantages and disadvantages of chemisorption and physisorption.
PhysisorptionChemisorption
Advantages
  • Reversible in nature
  • Low adsorption enthalpy
  • Favors low temperature
  • Low activation energy
  • Strong interaction between the adsorbent and the adsorbate by chemical bonds
  • Higher selectivity
Disadvantages
  • Weak interaction between the adsorbate and the adsorbent
  • The extent of adsorption is inversely proportional to temperature.
  • Low selectivity
  • Irreversible in nature
  • High adsorption enthalpy
  • Favors high temperatures
  • High activation energy
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

Shanaah, H.H.; Alzaimoor, E.F.H.; Rashdan, S.; Abdalhafith, A.A.; Kamel, A.H. Photocatalytic Degradation and Adsorptive Removal of Emerging Organic Pesticides Using Metal Oxide and Their Composites: Recent Trends and Future Perspectives. Sustainability 2023, 15, 7336. https://0-doi-org.brum.beds.ac.uk/10.3390/su15097336

AMA Style

Shanaah HH, Alzaimoor EFH, Rashdan S, Abdalhafith AA, Kamel AH. Photocatalytic Degradation and Adsorptive Removal of Emerging Organic Pesticides Using Metal Oxide and Their Composites: Recent Trends and Future Perspectives. Sustainability. 2023; 15(9):7336. https://0-doi-org.brum.beds.ac.uk/10.3390/su15097336

Chicago/Turabian Style

Shanaah, Haneen H., Eman F. H. Alzaimoor, Suad Rashdan, Amina A. Abdalhafith, and Ayman H. Kamel. 2023. "Photocatalytic Degradation and Adsorptive Removal of Emerging Organic Pesticides Using Metal Oxide and Their Composites: Recent Trends and Future Perspectives" Sustainability 15, no. 9: 7336. https://0-doi-org.brum.beds.ac.uk/10.3390/su15097336

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop