Next Article in Journal
Decolorization of Lactose-6-Phosphate Solutions Using Activated Carbon
Previous Article in Journal
Hierarchical Porous Carbon Aerogel Derived from Sodium Alginate for High Performance Electrochemical Capacitor Electrode
Previous Article in Special Issue
Current Trends and Future Perspectives in the Remediation of Polluted Water, Soil and Air—A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Metal Oxide Nanoparticles’ Green Synthesis by Plants: Prospects in Phyto- and Bioremediation and Photocatalytic Degradation of Organic Pollutants

by
Mohamed Ashour
1,*,
Abdallah Tageldein Mansour
2,3,*,
Abdelwahab M. Abdelwahab
2,4 and
Ahmed E. Alprol
1
1
National Institute of Oceanography and Fisheries (NIOF), Cairo 11516, Egypt
2
Animal and Fish Production Department, College of Agricultural and Food Sciences, King Faisal University, P.O. Box 420, Al Hofuf 31982, Saudi Arabia
3
Fish and Animal Production Department, Faculty of Agriculture (Saba Basha), Alexandria University, Alexandria 21531, Egypt
4
Department of Animal Production, Faculty of Agriculture, Fayoum University, Faiyum 2933110, Egypt
*
Authors to whom correspondence should be addressed.
Processes 2023, 11(12), 3356; https://doi.org/10.3390/pr11123356
Submission received: 9 October 2023 / Revised: 29 November 2023 / Accepted: 30 November 2023 / Published: 2 December 2023
(This article belongs to the Special Issue Remediation Process of Environmental Pollution)

Abstract

:
Over the past few decades, the production of metal oxide nanoparticles (MONPs) has developed into an exciting and sophisticated research area. Green metal oxide nanoparticles have played an extremely imperative role in various fields, including biomedical, environmental, energy, agricultural applications, catalytic, bioactive, antibacterial, poisonous, and biocompatible. To achieve sustainability and adopt environmentally friendly practices, the production of MONPs is now increasingly focused on exploring green chemistry and alternative pathways. When made using green synthesis techniques, the metal oxide nanoparticles are especially important because they do not require external stabilizers, capping agents, dangerous chemicals, or harsh operating conditions (high pressure and temperature). Plant-mediated synthesis of different MONPs using either whole cells or extracts has several advantages, including rapid synthesis (compared with other biogenic processes (using fungi and bacteria)), being more stable than other types, being available in nature, and being non-toxic. This review provides a comprehensive overview of the green synthesis of MONPs using plant parts, factors affecting the synthesis, and the characterization of synthesized NPs. Additionally, it highlights the potential of these environmentally friendly nanoparticles that are widely used to treat environmental pollutants, including the removal of heavy metals, antibacterials, and the degradation of organic pollutants.

1. Introduction

For the development of green technology, science and technology have advanced most quickly [1,2,3,4]. The word “nanomaterial” (NM) refers to materials that have a size between 1 and 100 nm, at least in one dimension, and distinct properties in terms of porosity, form, size, and other functions [5,6]. It has gained lots of attention mainly owing to its electrical, magnetic, chemical, distinctive optical, and mechanical properties that set it apart from its bulk counterparts. The dimensions and shape of a nanoparticle play a key role in determining its essential characteristics and properties, including its optical, electrical, catalytic, bioactive, antibacterial, poisonous, and biocompatible traits [7,8,9].
Metal oxide nanoparticles are primarily synthesized using physical and chemical methods. However, these techniques involve the use of highly reactive and toxic reducing agents like sodium borohydride and hydrazine hydrate, which have harmful effects on the environment, plants, and animals. Furthermore, these methods require sophisticated equipment, complex procedures, and strict experimental conditions, posing significant challenges. Consequently, physical methods result in the production of heterogeneous nanoparticles with high energy consumption, while chemical methods rely on synthetic agents for capping, reducing, and stabilizing and generate non-eco-friendly byproducts, although they yield the desired homogeneous metallic nanoparticles. Additionally, the nanoparticles produced through physical and chemical processes cannot be used for medical purposes due to their adverse effects on human health [3].
Green synthesis has become a more popular term to describe these biological techniques. The term “green synthesis” refers to methodologies, approaches, and processes that synthesize nanoparticles with the least amount of energy consumption while avoiding harmful byproducts through regulation, control, remediation, and cleanup. Waste prevention and minimization, a decrease in byproducts and pollution, the use of greener (or nontoxic) solvents and auxiliaries, and renewable feedstock are the guiding principles for green synthesis [10]. Numerous methods of green synthesis of nanoparticles include mild reaction conditions, mild reducing agents, microwaves, microbes, solar energy, and ultrasound methods [11]
Many biological sources, including yeast, fungi, bacteria, actinomycetes, microalgae, seaweeds, waste aquaculture, plant extracts, disposed-of food products, and horticultural food materials [12], have already been used as substances to produce steady and well-functionalized nanomaterials [13,14,15] (Figure 1). Plant parts, such as leaves, roots, etc., are considered an environmentally friendly and dependable method for producing metal oxide nanoparticles, and synthesizing them is a popular choice for eco-conscious synthesis [16,17]. Also, biological methods have gained much interest lately due to their simple nature, ease of usage, and ecological safety. Alkaloids, phenolic acids, terpenoids, polyphenols, and polysaccharides are a few phytochemicals found in the extracts of plants that function as bioresistant and capping agents in the biogenic production of metal oxide nanoparticles [18,19].
Plant-based synthesis of green MONPs has gained popularity as a phrase to designate these biological methods for the advancement of green technology. Recently, scientists have put a lot of effort into developing chemical processes that are reliable, simple, and environmentally friendly. Sundrarajan et al. [20] produced titanium dioxide nanomaterials (TiO2NPs) from Morinda citrifolia leaf extract. SEM, XRD, FTIR, and EDX approaches were applied to characterize the produced nanomaterial. Interestingly, 10 nm-sized average crystallites of tetragonal rutile phase TiO2 were obtained. TiO2NPs have antimicrobial properties with more efficiency.
Using Annona squamosa leaf extract as a bioreductant, a noble ZnO-NP synthesis has demonstrated intriguing antibiotic and anticancer activities with a focused action by securing the natural microbiota [21]. In another study, Madan, et al. [22] conducted research into the eco-friendly production of multifunctional zinc oxide nanoparticles with diverse morphologies in the hexagonal wurtzite structure. To achieve this, they employed a low-temperature solution combustion method by utilizing the extract of Azadirachta indica (Neem) leaf as a model, used as an antibacterial and photocatalytic for organic pollutants.
The study by Momeni et al. [23] employed a regenerative and nontoxic leaf extract of Euphorbia prolifera as a stabilizing agent and a mild reducing agent. This plant was used to produce Cu/ZnONPs with a size range of 5–17 nm. When used to break down dyes such as Congo red and methylene blue in their aqueous solutions while being exposed to NaBH4, the produced composite nanomaterial demonstrated outstanding catalytic activity. Using Aloe barbadensis leaf extracts with lowering and capping potential produced ZnO-NP as an efficient, cost-effective, and biogenic fabrication technique [24].
Heavy metals, organic dyes, organic wastewater, and other organic pollutants have been found in various harmful forms in the environment in recent years [25,26,27,28]; they are difficult to eradicate using old technologies, which necessitates the development of new advanced technologies [29,30].
Due to their minimal environmental risks, plant extracts are the primary focus of this study’s recent breakthroughs in the biosynthesis of MONPs. MONPs are nanoscale particles that come from anthropogenic and natural processes and contain both artificially created and naturally occurring particles [31]. As a result, they are better suited for usage in several applications, for example, lithium-ion batteries, sensors, catalysis, and environmental ones. Green MONPs are also widely employed as adsorbents, photocatalysts, and antimicrobials to remove various contaminants [32]. As a result, this work aims to encourage green chemistry through the quick, inexpensive, and ecologically friendly green synthesis of metal oxide nanoparticles by employing plant extracts; a relatively new technique that results in reliable green chemistry and highlights the use of such synthetically created ecologically friendly metal oxide nanoparticles, particularly in wastewater treatment.

2. Green Synthesis Definition

The adoption of green fabrication as an alternative to traditional chemical methods of producing MONPs is gaining momentum as an environmentally friendly option. Green synthesis includes technologies, strategies, and procedures aimed at reducing energy consumption and eliminating harmful by-products through regulation, control, processing, and cleaning. Utilizing ecologically friendly or non-toxic solvents and additives, as well as renewable raw materials, is the fundamental tenet of green synthesis, together with waste reduction and minimization [10]. Because reduction and capping agents are needed through the synthesis methods, metallic nanoparticles can be made more effectively utilizing green synthesis approaches. Aside from cyanobacteria and plant parts (fruit, root, flower, and leaves), natural sources for the ecologically friendly synthesis of nanoparticles include actinomycetes, bacteria, yeast, fungal organisms, and cyanobacteria [33]. As a reducing medium to create nanoparticles and as a capping medium to stabilize those nanoparticles, the whole cell of an algae, a microbe of a plant, or an extract can be employed [34]. It is necessary to produce chemicals and molecules based on “green synthesis” principles, which call for environmentally friendly procedures. Green chemistry adheres to the 12 green chemistry principles that were previously discussed, which is why it has been given that name [35].

3. Metal Oxide Nanoparticles

Nanomaterials are elements having at least one dimension between 1 and 100 nm, with distinct properties in terms of pores, dimensions, form, and other elements [36]. The reaction between a metal cation and oxygen gas results in the creation of metal oxides. Metal oxides are a fascinating group of inorganic substances that have been widely investigated and studied due to their diverse forms and properties. Transition metal oxide nanoparticles are particularly significant because of their attractive electrical, optical, and magnetic properties. These inorganic materials, known as metal oxides, have piqued the interest of researchers, attributed to their diverse forms and properties. Transition metal oxide nanoparticles, in particular, possess attractive electrical, optical, and magnetic properties, which make them well-suited for an extensive range of uses, including environmental applications, catalysis, lithium-ion batteries, and sensors. As a result, there is a growing need to explore their unique features and properties in various applications [31]. Metal oxides have bonds that can be virtually ionic, extremely covalent, or even metallic, and synthesis is more complicated than pure metals [37]. Similarly, MONPs are used in a variety of sectors, including magnetic storage media, catalytic processes, sensors, electronics, and solar energy conversion, thanks to their varied characteristics. Among various metal oxide types, TiO2, Fe3O4, CuO, and ZnO are the most widely researched for the synthesis of MONPs using plant-based methods [38].

4. Factors Affecting the Synthesis of Inorganic Nanoparticles

pH, exposure duration, reaction time, temperature, and stirring rate are a few variables that affect the synthesis of NPs. The number of bioactive compounds in the plant extract is another consideration. The synthesis of nanoparticles as well as their shape, properties, size, and uses are all influenced by these variables [39].

4.1. pH

The form and size of the MONPs produced may change in the reaction’s pH. The buffer strength (also known as pH), which controls NPs’ shapes and sizes, must remain constant throughout the NPs’ generation procedure. Larger particles are created by comparing lower acidic pH values to higher acidic pH values [13]. In contrast, a high pH environment favors the growth of spherical NPs, and synthesis is excellent for the synthesis of small-sized MONPs. The surface plasmon resonance (SPR) peak develops wide and deflects towards a longer wavelength region, which results in a wide range of NPs (usually triangular and cylindrical shaped), compared with synthesizing small-sized NPs when the pH is low [40].

4.2. Temperature

Temperature increase has shown catalytic behavior by boosting the reaction rate and efficiency of nanoparticle synthesis while using biological entities to formulate nanoparticles. Several chemical processes, such as electrochemical, solvothermal, or templating procedures, are significantly influenced by the reaction’s temperature. In general, temperatures of close to 100 °C are required for the synthesis of NPs utilizing green techniques. While chemical synthesis requires mild temperatures, physical reactions require temperatures higher than 350 °C [41]. The average size of the group shrinks as their conversion rate rises. Researchers have postulated that a crucial enzyme included in the synthesis of nanoparticles may have high temperatures [13].

4.3. Concentration of Extracts

Constituents of plant extracts serve as both reductants and stabilizers of generated nanoparticles. The number of bioreducing agents in a plant extract influences the characteristics of nanoparticles by regulating the particle’s size and shape. High plant extract content accelerates nanoparticle synthesis by increasing nanoparticle nucleation. As the concentration of the reducing agent in the plant extract increases, the nucleation rate increases. However, smaller nanoparticles are produced by rapid synthesis rates at high extract concentrations [42].

4.4. Reaction Times

The time taken for a reaction plays a vital role in the synthesis of nanoparticles. In fact, conducting the same experiment while varying the reaction time can lead to variations in the size of the particles produced. The period during which metal salts are exposed to a plant extract is referred to as the “contact time” and is used as a measure of the reaction duration. The concentration of the plant extract also has a noteworthy influence on both the reaction and contact time. Plant extracts containing higher concentrations of reductive groups tend to facilitate faster nanoparticle formation [42].

4.5. Stirring Speed

The stirring speed plays a crucial role in regulating the rate and size of nanoparticle synthesis. Bhattacharya et al. [43] revealed that, when the stirring speed was increased to 900 rpm, the size of CuO NPs decreased from 42.54 nm at 500 rpm to 3.6 nm. Moreover, the smaller size may be accredited to the bigger nanoparticles being broken up into smaller ones at a faster stirring speed. Increased nanoparticle synthesis rate brought on by uniform distribution of metal salts and bioreductant species of the plant extract is another benefit of faster stirring. Homogeneous distribution speeds up the reduction of metal salts to nanoparticles by enhancing the interaction among species of bioreductant and metal salts [11].

4.6. Choice of Species

The cost-effectiveness of NP biosynthesis depends on the physical–chemical factors listed above, as well as (a) choosing the right organism or species based on crucial intrinsic traits, including metabolic pathways, growth rate, and enzyme activity; (b) the amount of its inoculum; and (c) the choice of biocatalysts, which are vital to speed up the rate of reaction (i.e., reduction) (although this should be taken into account) [3]. Enzymes, raw cells, or whole cells can all be employed as biocatalysts (e.g., NADPH, FAD). Because coenzymes are expensive, they can be recycled along the pathway, and they also have a large amount of effectiveness [44].

5. Various Techniques for Synthesizing MONPs

MONPs can be created using either a “top-down” process or a “bottom-up” strategy. A top-down strategy is used to manufacture MONPs by breaking up bulk materials into tiny particles and reducing their thickness [45]. Also, bottom-up synthesis (referred to as “self-assembly”) is a method of producing nanoparticles from smaller units, including atoms, molecules, and smaller connecting unit particles. In contrast, the top-down technique involves breaking down larger structures using manufacturing methods, such as cutting, grinding, and shaping, to achieve the desired nanoscale dimensions [46]. Contrarily, the bottom-up approach uses chemical interactions to create nanoscale structures [47]. For this procedure, physical and chemical techniques can be utilized, including mechanical methods (such as crushing and grinding) [45], sputtering [48], chemical etching [48], thermal evaporation [49], pulsed laser ablation [13], and photoreduction methods [50]. The top-down technique has a key disadvantage in that the surface structure is not perfect [51].
In order to mitigate toxicity and enhance the effectiveness of chemically synthesized nanoparticles (NPs), scientists have utilized plant metabolites as adjuvants. This approach involves utilizing biological extracts derived from microorganisms or plants as reducing agents to convert metal ions into nanoparticles. This reduction process can occur either outside the cell (extracellularly) or within the cell (intracellularly). By employing plant metabolites in this manner, researchers aim to improve the safety and efficiency of the synthesized nanoparticles. The extract components operate as capping agents and assist in producing homogenous stable MONPs by inhibiting aggregation [11]. This synthetic method is used to create MONPs that are biocompatible, scalable, non-toxic, repeatable, and helpful in medicine on a big scale. Furthermore, it is anticipated that metabolites consumed during the reduction process will adhere to the surface of NPs and boost activity [48]. The manufacture of NPs using biological entities is simple (one-pot synthesis), economical, quick, and environmentally friendly [52]. Nanoparticles have a basic advantage over biological ones in that their size and form can be controlled, and they can be produced in greater numbers utilizing chemical and physical processes [53]. The plant parts that follow will describe how to synthesis MONPs utilizing a variety of biological materials found in nature.

6. Biosynthesis of MONPs by Plants

Due to its environmentally friendly nature, using plant systems for the synthesis of MONPs has been hailed as a “green” method and a dependable technique [54]. To synthesize MONPs through biogenic means, various phytochemicals found in plant stems, leaves, and flowers, including polyphenols, alkaloids, tannins, terpenoids, phenolic acids, and polysaccharides, serve as both reductants and capping agents [55]. Many areas of plant parts can collect heavy metals. As a result, methods for the synthesis of nanoparticles using plants have gained popularity, as they are rapid, efficient, practical, and economical. They are a fantastic substitute for traditional preparation methods (Figure 2). In a “one-pot” synthesis approach, a variety of plants can be used for the reduction and maintenance of metallic nanoparticles. Proteins, polysaccharides, and coenzymes are examples of biomolecules that plants can use to transform metal ions into nanoparticles. The synthesis of nanoparticles is based on several factors, including the type of plant extract used, the quantity of the extract and metal salt, temperature, pH level, and time of contact. Plant nanoparticle synthesis is cost-effective and environmentally friendly and it is safe for humans. In another study, Sharma et al. [56] utilized the Neem, Azadirachta indica, leaf extract as a reducing agent in the synthesis of nanoscale tetragonal hausmannite crystals of Mn3O4. This study created a very effective chemical sensor for identifying the presence of 2-butanone using the generated nanomaterial as a working electrode.
By reducing the thermal decomposition temperature of ammonium perchlorate to 175 °C in a single decomposition phase, the sensor was able to function effectively. The study by Al-Ruqeishi et al. [57] used polyphenols contained in Omani mango (disambiguation) tree leaves to create a green synthesis of Fe2O3-NP nanorods. The average diameter and length of the Fe2O3-NPs generated were 15 nm and 3 nm, respectively. The study by Kumar et al. [58] described the production of zinc oxide nanoparticles using grapefruit extract. ZnONPs had a size range of 12–72 nm (TEM) and functioned well as an antioxidant and photocatalyst in the methylene blue degradation process (>56% efficiency for photocatalysis). The information that is currently available on the size, characterization, and applications of some MONPs that are produced using different plant extracts in a green manner is reported in Table 1.

7. Mechanism of Nanoparticle Synthesis by Plant Extract

There are two methods commonly employed for the synthesis of metallic nanoparticles: intracellular and extracellular.
(a)
The extracellular technique, which utilizes plant extract, is often preferred for nanoparticle synthesis due to its simplicity in handling, cost-effectiveness, and the absence of an additional step required to disrupt cells in order to recover internal nanoparticles for subsequent processing. In the extracellular method, metal salts are converted into metal nanoparticles using a suspension of plant extract. Plant extracts are abundant in various bioactive compounds, which serve as both reducing agents and capping agents during nanoparticle production. These compounds play a crucial role in reducing metal ions and stabilizing the resulting nanoparticles [89]. Metal nanoparticles are formed through the reduction of metal salts using bioactive substances released by plant cells in an extract medium. The process involves two stages: nucleation and crystal growth. During the nucleation stage, precursor metal salts (M+) are transformed into metal nanoparticles (M0) through a reaction facilitated by a component present in the plant extract. This reaction initiates the formation of small clusters of metal atoms that serve as the building blocks for the nanoparticles. In the crystal growth stage, the metal nanoparticles (M0) undergo complexation with other metal salts (M+), resulting in the formation of M0-M+ complexes. This complexation process enables the nanoparticles to grow in size. Additionally, the nanoparticles undergo adsorption, where they bind together to form nanocrystals, often in the form of metal nanoparticle dimers. Overall, the combination of nucleation and crystal growth processes using bioactive substances from plant extract allows for the production of metal nanoparticles [90].
(b)
In intracellular synthesis, metal salts are reduced within the cells through the assistance of various enzymes such as nitrate reductase and NADH- and NADPH-dependent enzymes. These enzymes are involved in essential metabolic processes like photosynthesis, nitrogen fixation, and respiration. Additionally, plant pigments such as chlorophyll also contribute to the intracellular reduction process [90]. Both living and dry biomasses of plants are capable of performing intracellular synthesis of nanoparticles. This process occurs as a defense mechanism of plant cells against stress conditions induced by metals. When exposed to metal ions, the metabolic machinery of plant cells is redirected to mitigate the toxic effects of metals. As a result, the cells utilize their inherent enzymatic systems and pigments to reduce metal salts and produce nanoparticles intracellularly [91]. The process of nanoparticle formation inside cells involves the binding of positively charged metal ions to the surface of the cell wall or to negatively charged protein or enzyme groups within the cytoplasm. Once the metal ions are trapped, they undergo reduction to form small nuclei, which serve as the starting point for nanoparticle synthesis. This process can lead to the production of nanoparticles with various morphologies. In the case of algal cells, there is an intracellular enzyme present that facilitates the internal conversion of metal salts into nanoparticles. This enzyme plays a crucial role in the reduction process, enabling the transformation of metal ions into nanoparticles within the cellular environment [91].

8. Characterization of MONP Plant Synthesis

Characterization refers to the process of describing a nanoparticle’s properties after its synthesis, including its size, capping agents, shape, crystallinity, surface plasmon resonance, associated functional groups, polydispersity, and surface charge. These physical and chemical properties are essential in designing nanomaterials for specific practical applications. Table 2 displays the different methods used to characterize MONPs for these properties, and the following descriptions outline the techniques used to determine each feature [92].

9. Adsorption Method for the Elimination of Various Pollutants

Adsorption methods are applied most often to remediate contaminated liquids. Adsorption methods are used for removing organic and inorganic pollutants from aqueous solutions, with the aggregation of metal ions onto adsorbent pores and surfaces creating a layer that includes ions as other impurities [93].
For a number of reasons, adsorption is preferred over traditional metal removal methods. Adsorbents are inexpensive and synthesis of them is economical. They also have the following benefits: (1) no sludge is produced, (2) they are regenerative, i.e., the adsorbent can be reused after the pollutant ions have been removed, (3) they use less energy and chemicals, (4) there may be metal recovery, and (5) their performance is competitive, i.e., their achievements are comparable (Figure 3) [9]. Parameters that influence the sorption efficiency include the adsorbate–adsorbent interaction, adsorbent surface area, temperature, contact time, amount of adsorbent, initial pollutant concentration, particle size, pH, etc. [11,94].

9.1. Application of Green MONPs Using Plant Extracts in Water Treatment

9.1.1. Removal of Heavy Metals

Due to the simplicity of modifying their surface functionality and their high surface area-to-volume ratio, MONPs have enhanced adsorption capacity, efficiency, and reusability. In addition, several researchers have investigated the use of MONPs for heavy metal adsorption [95]. According to Srivastava et al. [96], green ZnONPs show a good removal efficacy for Cd(II). Within an hour of contact time, 92% of Cd(II) was removed when the adsorbent dosage was 200 mg L−1. Therefore, ZnONPs might be used to successfully remove cadmium metal from effluent wastewater.
Additionally, Ehrampoush et al. [97] investigated the fabrication of spherical Fe2O3-NPs with an average size of 50 nm by tangerine peel extract in order to eliminate Cd(II) from wastewater. At a pH of 4 and an adsorbent dosage of 0.4 g 100 mL−1 during a 90 min contact period, the peel extract of Fe2O3-NPs effectively removed cadmium ions by about 90%. Fazlzadeh et al. [98] studied the synthesis of ZnONPs by Peganum harmala seed extract. The nanoparticle offered by the authors was for the removal of Cr(VI) from effluent water, and it displayed 97.9% elimination efficiency at 1000 mg L−1 initial concentration of Cr(VI), pH 2 in 30 min, and 2 g L−1 of the MONPs. Table 3 presents the removal of heavy metals and dye pollutants using biosynthesized MONPs.
The study by Somu and Paul [100] investigated the synthesis of ZnONPs on a case-by-case basis to simultaneously purify wastewater of heavy metals, dyes, and harmful bacteria. The ZnONP, at pH 8.0, displayed the highest adsorption efficiency of 95.35% for Pb(II), 85.63% for Cd(II), and 71.23% for Co(II). The dye degradation was 93.72% for Congo red at pH 6.0 and 93.58% for MB at pH 7. Adsorption loss for loaded ZnONPs after five cycles was only 10–12% for both colors and metals, indicating that the ZnONPs were reusable when the adsorption–desorption methods were repeated five times. The green synthesis of cupric oxide nanoparticles (CuONPs) arbitrated by Calotropis procera latex was used for removing Cr(VI) from aqueous solutions. For the experiment, initial Cr(VI) concentrations and pH were 5 and 2 mg L−1, respectively [104].
While Srivastava et al. [96] revealed that ZnO NPs had a high removal efficiency for Cd(II) using 20 mg L−1 of adsorbent dose, the removal efficiency was ~55%; upon increasing the dose to 200 mg L−1, a 92% removal of Cd(II) within 1 h of contact time occurred. Thus, the ZnO nanoparticle could be successfully used for the removal of Cd from effluents. Debnath et al. [111] prepared ZnO nanoparticles by a green synthesis route from Hibiscus rosa-sinensis leaf extract for the removal of Congo red (CR) dye from an aqueous solution. The maximum adsorption capacity was 9.615 mg/g, with a high percentage of removal efficiency of 95.55%. Through a green approach, Nayak et al. [112] synthesized ZnONPs from Ocimum sanctum (Tulsi leaf) leaf extract as an adsorbent for the removal of Congo red (CR) dye. The removal capacity of ZnO-T was higher (97%) than commercially available ZnO (78%), and the maximum adsorption capacity of ZnO-CT was 74.07 mg/g. Furthermore, Mittal et al. [113] synthesized ZnO nanoparticles through an in situ process by using gum Arabic grafted polyacrylamide (GA-Cl-PAM). The GA-Cl-AAM-ZnO composite was used as an effective adsorbent to take up malachite green (MG) dye from an aqueous solution. The maximum adsorption capacity was 766.52 mg/g, with 99% high adsorption efficiency.

9.1.2. Fluoride Removal

One of the most severe health issues in the world is the fluoride pollution of drinking water. Excessive intake of fluoride leads to the depletion of calcium from the tooth matrix, leading to the formation of cavities and the eventual development of dental fluorosis. With prolonged exposure, skeletal fluorosis may also occur. Fluoride adsorption is greatly favored by metal oxide nanoparticles’ increased surface area [114]. Given their strong adsorption capacity, constrained water solubility, non-toxic composition, and favorable desorption potential, MONPs are regarded as promising materials.
Many different NP types, including Fe3O4@Al(OH)3 and Ce-Ti@Fe3O4, have been utilized to remove fluoride from aqueous solutions because of their large surface area, self-assembly, strong reactivity, specificity, and recyclable nature. The former has a high adsorption rate, whereas the latter has a maximum adsorption capacity (qmax = 91.04 mg g−1) at pH 7. Also, the author mentioned that, after five cycles of usage, there was no discernible loss in adsorption capacity. Composite Fe3O4@Al(OH)3 nanoparticles combine the advantages of magnetic separation and Al2O3, because aluminum oxide has a large affinity for fluoride [115]. Furthermore, the study by Chai et al. [116] developed separated Fe3O4/Al2O3 nanoparticles with sulfate doping for the removal of fluoride from drinking water. Nearly 90% of the adsorption could be accomplished in about 20 min, and only 10–15% of additional elimination happened in 8 h. By using the Langmuir model, the nano-adsorbent had a fluoride absorption capacity of 70.4 mg g−1 at pH 7.0.
The study displayed by Silveira et al. [117] discovered that fluoride ion adsorption was beneficial, with maximum adsorption at pH 7. They created FeONPs by employing Moringa oleifera leaf extract as a model for the removal of fluoride ions from wastewater. With an adsorption potential of 1.40 mg g−1 and an adsorption kinetics test, equilibrium was reached in 40 min. FeONPs can be reused three times according to the regeneration process.
In another study by Kumari and Khan [118], Simmondsia chinensis (jojoba) defatted meal was used to synthesize Fe3O4 for defluoridation applications in drinking water. However, in the pH scope of 5.0–9.0, the F removal percentage decreased. While F adsorption decreased at alkaline pH levels because of competition among F ions and hydroxyl ions for active surface sites, F removal decreased at acidic pH levels owing to the formation of hydrofluoric acid (HF). At an 80 min contact time, it was found that 93% of the fluoride had been removed. The Fe3O4 NPs’ polyurethane foam (PUF) exhibited a greater water defluoridation capacity of 34.48 mg g−1.

9.2. Photocatalytic Activity of MONPs

The term “photocatalysis” derives from combining “photo” and “catalysis,” where “photo” refers to light and “catalysis” refers to modifying a chemical reaction using a catalyst [119]. Photocatalysts are catalysts that become active when exposed to light radiation and can convert pollutants in the air or water to less toxic compounds, such as water and CO2. When exposed to visible sunlight, photocatalysts can degrade organic molecules similarly to semiconductors. A hole is created in the valence band because of excitably moving an electron from the valence band to the conduction band [120,121,122].
In the presence of oxygen and water, the holes and electrons created during this process can generate anionic superoxide and hydroxyl radicals, which are capable of decomposing organic molecules [123]. To be effective, a photocatalyst must have a bandgap energy that is less than 3 eV so that it can absorb light in the visible spectrum. Either homogeneous or heterogeneous photocatalytic development is possible, based on the physical characteristics of the catalyst and the interacting species [124].

9.2.1. Advantages of Nano-Photocatalysts

The reduction or elimination of harmful organic compounds is mostly aided by nano-photocatalysts. The main benefit of nano-photocatalysts is related to the quantum size effect, which widens the energy bandgap (Eg) and reduces the size of the nanoparticles. Therefore, there are many benefits to using the photodegradation approach, including generally full deterioration, cheap cost, and being reusable. Despite all these developments, toxicity and catalyst recovery remain issues for nano-photocatalysts. The main advantages and disadvantages of the several physicochemical treatments discussed in this work are reported in Table 4.

9.2.2. Factors Affecting the Photocatalytic Degradation

(a) 
Amount of Catalyst
In general, as a catalyst dosage is increased, there is an increase in the quantity of photomineralization of water pollutants. This is because the photocatalyst has more active sites exposed to light, which absorb more photons and generate more OH radicals and positive holes when exposed to light. The rate of pollutant breakdown accelerates as a result of these positive holes and hydroxyl radicals taking part in the photocatalytic activity. The majority of the solar light is passed directly from the solution when utilizing a lower dose catalyst, and just a small fraction is utilized by the catalyst to function [126]. A larger catalyst dosage, after a certain threshold, lowers the rate of pollutant photodegradation. As a result, fewer photoactive species are active, which lowers the percentage of deterioration. Another explanation could be that there are fewer active surface sites available due to the aggregation of nanoparticles at high concentrations in the solution [126].
(b) 
Amount of Pollutant
The quantity of organic pollutants adsorbed onto the surface of the photocatalyst is one of the key parameters affecting the degradation percentage in a photocatalyst reaction. The quantity of contamination in the bulk aqueous media has no bearing on this. [127]. The synthesis and interaction of OH radicals with the contaminants affects how quickly the pollutants degrade. The photocatalyst’s capacity for adsorption is affected by the initial concentration of pollutants in the medium, which normally rises as the pollutant concentration increases. Photodegradation efficiency is enhanced when there is a higher initial concentration of pollutants, as it is easier for more contaminant molecules to be stimulated by light irradiation [128].
(c) 
Influence of pH
The pH of the aqueous solution affects a variety of processes, including surface charge, agglomeration, ionization, valence band potential of the photocatalyst, and contaminant adsorption. The three main factors that contribute to photodegradation are: (a) oxidation of pollutants owing to the high oxidation potential of positive holes, (b) attack by hydroxyl radicals (OH), and (c) reduction of contaminants by conduction electrons. As a result, determining the influence of pH on the effectiveness of catalysts for photodegradation is a difficult task [129].
(d) 
Temperature, Time, and Morphology
Temperature, time, and shape all have a massive effect on photocatalytic performance. With the use of environmentally friendly ZnONPs produced in the temperature range of 25 to 40 °C, Hassan et al. [130] showed how temperature might affect the rate of anthracene degradation. According to the degradation efficacy values, a minor rise in reaction rate is caused by a temperature increase. Raising the temperature may have little detrimental effect on photocatalytic degradation, but it may also induce the molecules to collide more frequently.

9.2.3. General Reaction Mechanism of Photocatalysis

Numerous groups have considered the fundamental mechanics of photocatalytic pollution degradation. In the study by Theerthagiri et al. [131], they stated that the primary goal of a photocatalyst in breaking down pollutants is to expedite the oxidation and reduction methods when light is present. The process of photocatalytic degradation of pollutants includes three general stages: (1) the separation of electron and hole under light irradiation, (2) the dispersion of charge carriers on the photocatalyst’s surface, and (3) the light-driven catalytic oxidation and reduction that occur on the catalyst’s active sites. Equation (1) through (5) identify the significant steps in the photocatalytic degradation pathway. Firstly, the photocatalyst generates holes (h+) and electrons (e) in the valence band (VB) and conduction band (CB), respectively, once exposed to light. These photogenerated charges transfer to the photocatalyst’s surface to initiate the required oxidation and reduction reactions. The interactions of h+ and e with H2O and O2 create the highly reactive radical species O2 and OH, respectively, as depicted in step (2). As shown in step (3), the breakdown of these pollutants occurs as a result of the interaction of these radical species with organic molecules.
C a t a l y s t + h v e C B + h V B +
e C B + O 2 O 2
O 2 + e C B + 2 H O H + O H
h V B + + H 2 O H + + O H
D y e + O H D e g r a d a t i o n   p r o d u c t s   ( C O 2 , H 2 O , e t c . )

9.2.4. Application of Green MONPs’ Organic Pollutant Treatment by Photocatalytic Activity

According to Jayaseelan et al. [132], excellent photodegradation effectiveness (>80%, 0.24 gL−1 of ZnONPs, 1 h) was achieved by ZnO nanoparticles produced from Jackfruit (Artocarpus heterophyllus) leaves for the removal of Rose Bengal dye emitted by the textile industry’s activities in the water. Cassia fistula leaves were utilized to produce ZnO nanoparticles, which were then used to photodegrade methylene blue dye in an aqueous solution.
In contrast to the same nanoparticle, Ref. [133] manufactured ZnONPs by using C. citriodora leaf extract as a bioreductant for the removal of MB dye. Under visible light irradiation for 90 min, the bioinspired ZnONPs (20 mg) showed an MB degradation efficiency of 83.45%, while the conventionally synthesized ZnONPs only had a degradation efficiency of 59.47%. Consistent with the study by Fulekar et al. [134], TiO2 nanoparticles for the removal of methyl orange dye under UV irradiation were created utilizing sorghum root extract. The results displayed that 99.7% of the methyl orange dye solution was degraded by TiO2 nanoparticles. Concerning Prasad [34], who synthesized ZnO NP from lemon juice, under UV light irradiation, 100 mg nanogranules of ZnO was evaluated to cure three different dye solutions: Congo red, CR (67.16 mg L−1), MB (31.98% mg L−1), and Rhodamine, Rh-B (47.90 mg L−1). Photocatalytic degradation efficiency was obtained for 91.17% of the MB solution after 70 min, 90% for the CR solution after 35 min, and 98% for the Rh-B solution after 50 min.
According to Muthukumar and Matheswaran (2015), the green synthesis of FeO NPs was mediated by an extract from the leaves of Amaranthus spinosus. When employing as-synthesized FeO-NPs under solar radiation, the photocatalytic color removal effectiveness for MO (75%) was greater than that of MB (69%) [135]. More recently, Rather and Sundarapandian [136] synthesized rod-shaped magnetic iron oxide nanostructures utilizing Wedelia urticifolia leaf extract in an environmentally friendly manner. These nanostructures had a strong ability to breakdown methylene blue dye. Another study by Prasad et al. [137] used an aqueous extract of Pisum sativum peels to determine an environmentally friendly approach for the fabrication of Fe3O4 nanoparticles. The outcomes verified that Fe3O4 NPs were ferromagnetic. UV-visible spectroscopy was used to examine the catalytic capabilities of Fe3O4 NPs for the breakdown of methyl orange (MO) dye in an aqueous solution. The biosynthesis of iron oxide nanoparticles utilizing peel extracts from Citrus paradisi was recently introduced by Kumar et al. [138] when 1, 1-diphenyl-2-picrylhydrazyl was present; the prepared iron nanoparticles showed antioxidant activity and eliminated the dyes MR (96.65%, 50 mg L−1), MB (80.76%, 10 mg L−1), and MO (89.64%, 10 mg/L).
ZnO NPs were synthesized using a green method (Corymbia citriodora leaf extract), and their photocatalytic activity was demonstrated by Yuhong et al. [133] by degrading MB under UV light, and the photocatalytic activity of hydrothermal and green-synthesized ZnO NPs was examined. Because of their smaller size, the ZnO NPs made using the green synthetic method had higher photocatalytic activity (83.45%) after 90 min of irradiation than ZnO NPs made using the hydrothermal method (59.47%). CuO NPs were synthesized by Nethravathi et al. [139] utilizing leaf extract from Tinospora cordifolia. When exposed to UV and sunlight, the produced particles in the size range effectively degraded MB, exhibiting a structure akin to a sponge. The findings showed that, when the catalytic load was raised, the effectiveness of MB photodegradation rose. At pH 2, the photodegradation efficiency was 91.23% and, at pH 4, the dye degradation climbed to 96.93%.
Amaranthus spinosus leaf extract was used in the green production of FeO NPs, as reported by Muthukumar and Matheswaran [135]. Under solar radiation, the percentage of color reduction achieved with leaf-mediated FeO NPs was 75 ± 2% for MO and 69 ± 2% for MB. The green synthesis of TiO2 NPs employing the Rhizoma coptidis root was reported by Sreekanth et al. The photocatalytic activity of MB and MG dyes in both dark and UV light settings was assessed using the produced NPs. After 30 min in the dark, about 39% (MB) and 4% (MG) had been absorbed. The TiO2 NPs showed good photocatalytic activity under UV irradiation; after 60 min, 71% of the MB and 78% of the MG dyes were removed [140]. In addition, NiO NPs were synthesized by Ezhilarasi et al. [141] utilizing leaf extract from Aegle marmelos. According to this study, 4-chlorophenol mineralization was less than 1% at 90 min, but it swiftly rose to 87% at 180 min, and total mineralization was seen at 210 min. Yadav et al. [142] used watermelon juice and a straightforward solution combustion approach to synthesize CeO2 NPs. Under UV radiation, about 98% of the MB dye degradation was seen after 180 min, whereas 93% of the dye degradation occurred in the presence of sunshine.
Other studies on the photocatalytic activity of metal oxide nanoparticles for the elimination of dye pollutants are recorded in Table 5.

9.2.5. Water Disinfection from Microbes

Numerous engineering and natural nanomaterials have also been found to possess potent antibacterial capabilities [160]. ZnONPs may be utilized to eliminate total coliform bacteria from municipal wastewater effluent, according to a study conducted by Mostafaii et al. [161]. The effectiveness of nanoparticles was 96% at 0.3 gL−1 in 20–90 min, 75–98.66% at 0.5 gL−1 in 20–90 min, and 65% at 0.7 gL−1 in 20 min, reaching 99% in 90 min. Finally, 100% of the total coliform (TC) bacteria was completely eliminated from municipal sewages in 90 min at a concentration of 1.1 g L−1 of ZnO.
Aeromonas hydrophila was employed to synthesize new ZnO nanoparticles, which were then evaluated against pathogenic microbes. E. coli, A. hydrophila, S. aureus, A. flavus, S. pyogenes, E. faecalis, A. niger, P. aeruginosa, and C. albicans were the objectives of the good diffusion tests. It has been reported that using Bauhinia tomentosa leaf extract as the bioreducing agent produced ZnONPs with an average size of 22–29 nm. The bioinspired NP’s antibacterial efficacy was estimated against P. aeruginosa, S. aureus, B. subtilis, and E. coli. ZnONPs were discovered to have a larger bactericidal effect on Gram-negative bacteria than Gram-positive bacteria owing to the structural differences between the two types of bacteria [162]. Bacterial growth was significantly halted after bactericidal doses of ZnONPs were administered [163]. The utilization of an aqueous leaf extract of Origanum vulgare to create CuONPs by the environmentally conscious generation led to a reduction in the cyanobacterium Microcystis aeruginosa population. This bacterium is commonly found in eutrophic waterways and produces secondary metabolites that are harmful to both humans and animals. CuONPs suppressed Microcystis aeruginosa colony growth by about 89.7% at a load of 50 mg L−1 CuONPs and by 62.6% at a low dose of 25 mg L−1 [164].

10. Conclusions and Recommendations

The innovative features of nanoparticles produced by their synthesis processes lead to new synthesis directions in a variety of areas. Chemical, physical, and biological mechanisms can be used to create metal oxide nanoparticles. The biosynthetic approach provides a number of advantages over conventional chemical-based techniques. It has been confirmed that several biological sources, including bacteria, fungi, and plants, are capable of producing nanoparticles. The green approach to the manufacture of MONPs provides many advantages over conventional approaches, including simplicity in scaling up, affordability, and ease of synthesis. With this context in mind, this review highlights the many types of synthesized MONPs and nanocomposites, as well as their essential mechanisms in environmental pollution remediation. An innovative, inexpensive, and non-toxic technique for producing metal oxide nanoparticles on a wide scale could be offered by plant-mediated synthesis, given the expected rise in demand for nanoparticles. Heavy metals and some organic contaminants have recently been found in the environment in a variety of harmful forms that are challenging to eliminate using conventional methods, necessitating the urgent need for new material-removal solutions. Owing to their high nanoscale size effect properties, which make them excellent candidates to provide a solution for environmental remediation, metal oxide nanomaterials have recently garnered much attention. In addition to them, inexpensive green synthetic MONPs made from various plant and algal extracts may also be utilized, with massive benefits and results in the remediation field. Despite promising results, it has also been noted that the use of algal metal oxide nanoparticles for the recovery of hydrocarbon pollutants has been comparatively underexplored. Plant-mediated synthesis offers a promising solution for the large-scale synthesis of metallic nanoparticles at a low cost and without toxicity, which will be increasingly important as nanoparticle demand grows in the future. The recommendation from the current study can be pointed out as follows: Additional research is required to investigate the fate and transport of metal oxide nanoparticles as well as their toxicity when introduced into the marine environment. More research will also be needed to ascertain the effectiveness of green MONPs in large-scale remediation by various plant species (fixed-bed or fluidized-bed studies), in addition to their usage for industrial wastewater treatment.

Author Contributions

Conceptualization, A.E.A. and M.A.; validation, A.E.A. and M.A.; project administration, M.A. and A.E.A.; visualization, A.E.A., M.A. and A.T.M.; funding acquisition, A.E.A., M.A., A.T.M., A.M.A. and A.T.M.; writing—original draft preparation, M.A., A.E.A., A.M.A. and A.T.M.; writing—review and editing, M.A., A.E.A., A.M.A. and A.T.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Deanship of Scientific Research, Vice Presidency for Graduate Studies and Scientific Research, King Faisal University, Saudi Arabia (GRANT5,060).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

BETSurface area analysis
ICP-MSInductively coupled plasma-mass spectrometry
DLSDynamic light scattering
EDSEnergy-dispersive X-ray spectroscopy
EDXEnergy-dispersive X-ray
FTIRFourier transform infrared spectroscopy
NMRNuclear magnetic resonance spectroscopy
PLPhotoluminescence
SEMScanning electron microscopy
TEMTransmission electron microscope
TGAThermogravimetric analysis
XPSX-ray Photoelectron spectroscopy
X-rayX-ray diffraction technique

References

  1. Abualnaja, K.M.; Alprol, A.E.; Abu-Saied, M.A.; Ashour, M.; Mansour, A.T. Removing of Anionic Dye from Aqueous Solutions by Adsorption Using of Multiwalled Carbon Nanotubes and Poly (Acrylonitrile-styrene) Impregnated with Activated Carbon. Sustainability 2021, 13, 7077. [Google Scholar] [CrossRef]
  2. Abualnaja, K.M.; Alprol, A.E.; Abu-Saied, M.A.; Mansour, A.T.; Ashour, M. Studying the Adsorptive Behavior of Poly(Acrylonaitrile-co-Styrene) and Carbon Nanotubes (Nanocomposites) Impregnated with Adsorbent Materials towards Methyl Orange Dye. Nanomaterials 2021, 11, 1144. [Google Scholar] [CrossRef] [PubMed]
  3. Alprol, A.E.; Mansour, A.T.; El-Beltagi, H.S.; Ashour, M. Algal Extracts for Green Synthesis of Zinc Oxide Nanoparticles: Promising Approach for Algae Bioremediation. Materials 2023, 16, 2819. [Google Scholar] [CrossRef] [PubMed]
  4. Adeniran, A.A.; Ayesu-Koranteng, E.; Shakantu, W. A Review of the Literature on the Environmental and Health Impact of Plastic Waste Pollutants in Sub-Saharan Africa. Pollutants 2022, 2, 531–545. [Google Scholar] [CrossRef]
  5. Saif, S.; Tahir, A.; Chen, Y. Green synthesis of iron nanoparticles and their environmental applications and implications. Nanomaterials 2016, 6, 209. [Google Scholar] [CrossRef] [PubMed]
  6. Xie, K.; Wang, J.; Yu, S.; Wang, P.; Sun, C. Tunable electronic properties of free-standing Fe-doped GaN nanowires as high-capacity anode of lithium-ion batteries. Arab. J. Chem. 2021, 14, 103161. [Google Scholar] [CrossRef]
  7. Mabrouk, M.M.; Ashour, M.; Labena, A.; Zaki, M.A.A.; Abdelhamid, A.F.; Gewaily, M.S.; Dawood, M.A.O.; Abualnaja, K.M.; Ayoub, H.F. Nanoparticles of Arthrospira platensis improves growth, antioxidative and immunological responses of Nile tilapia (Oreochromis niloticus) and its resistance to Aeromonas hydrophila. Aquacult. Res. 2022, 53, 125–135. [Google Scholar] [CrossRef]
  8. Sharawy, Z.Z.; Ashour, M.; Labena, A.; Alsaqufi, A.S.; Mansour, A.T.; Abbas, E.M. Effects of dietary Arthrospira platensis nanoparticles on growth performance, feed utilization, and growth-related gene expression of Pacific white shrimp, Litopenaeus vannamei. Aquaculture 2022, 551, 737905. [Google Scholar] [CrossRef]
  9. Mansour, A.T.; Alprol, A.E.; Khedawy, M.; Abualnaja, K.M.; Shalaby, T.A.; Rayan, G.; Ramadan, K.M.; Ashour, M. Green Synthesis of Zinc Oxide Nanoparticles Using Red Seaweed for the Elimination of Organic Toxic Dye from an Aqueous Solution. Materials 2022, 15, 5169. [Google Scholar] [CrossRef]
  10. Singh, J.; Dutta, T.; Kim, K.-H.; Rawat, M.; Samddar, P.; Kumar, P. ‘Green’synthesis of metals and their oxide nanoparticles: Applications for environmental remediation. J. Nanobiotechnol. 2018, 16, 84. [Google Scholar] [CrossRef]
  11. Alprol, A.E.; Mansour, A.T.; Abdelwahab, A.M.; Ashour, M. Advances in Green Synthesis of Metal Oxide Nanoparticles by Marine Algae for Wastewater Treatment by Adsorption and Photocatalysis Techniques. Catalysts 2023, 13, 888. [Google Scholar] [CrossRef]
  12. Kumar, B.; Smita, K.; Cumbal, L.; Debut, A.; Angulo, Y. Biofabrication of copper oxide nanoparticles using Andean blackberry (Rubus glaucus Benth.) fruit and leaf. J. Saudi Chem. Soc. 2017, 21, S475–S480. [Google Scholar] [CrossRef]
  13. Shah, M.; Fawcett, D.; Sharma, S.; Tripathy, S.K.; Poinern, G.E.J. Green synthesis of metallic nanoparticles via biological entities. Materials 2015, 8, 7278–7308. [Google Scholar] [CrossRef] [PubMed]
  14. Mansour, A.T.; Alprol, A.E.; Abualnaja, K.M.; El-Beltagi, H.S.; Ramadan, K.M.A.; Ashour, M. The Using of Nanoparticles of Microalgae in Remediation of Toxic Dye from Industrial Wastewater: Kinetic and Isotherm Studies. Materials 2022, 15, 3922. [Google Scholar] [CrossRef] [PubMed]
  15. Mansour, A.T.; Alprol, A.E.; Ashour, M.; Ramadan, K.M.; Alhajji, A.H.; Abualnaja, K.M. Do Red Seaweed Nanoparticles Enhance Bioremediation Capacity of Toxic Dyes from Aqueous Solution? Gels 2022, 8, 310. [Google Scholar] [CrossRef]
  16. Akhtar, M.S.; Panwar, J.; Yun, Y.-S. Biogenic synthesis of metallic nanoparticles by plant extracts. ACS Sustain. Chem. Eng. 2013, 1, 591–602. [Google Scholar] [CrossRef]
  17. Darwesh, O.; Shalapy, M.; Abo-Zeid, A.; Mahmoud, Y. Nano-bioremediation of municipal wastewater using myco-synthesized iron nanoparticles. Egypt. J. Chem. 2021, 64, 2499–2507. [Google Scholar] [CrossRef]
  18. Jeevanandam, J.; Chan, Y.S.; Danquah, M.K. Biosynthesis of metal and metal oxide nanoparticles. ChemBioEng Rev. 2016, 3, 55–67. [Google Scholar] [CrossRef]
  19. El-Sheshtawy, H.; Khalil, N.; Ahmed, W.; Amin, N. Enhancement the bioremediation of crude oil by nanoparticle and biosurfactants. Egypt. J. Chem. 2017, 60, 835–848. [Google Scholar] [CrossRef]
  20. Sundrarajan, M.; Bama, K.; Bhavani, M.; Jegatheeswaran, S.; Ambika, S.; Sangili, A.; Nithya, P.; Sumathi, R. Obtaining titanium dioxide nanoparticles with spherical shape and antimicrobial properties using M. citrifolia leaves extract by hydrothermal method. J. Photochem. Photobiol. B Biol. 2017, 171, 117–124. [Google Scholar]
  21. Ruddaraju, L.K.; Pammi, S.; Pallela, P.V.K.; Padavala, V.S.; Kolapalli, V.R.M. Antibiotic potentiation and anti-cancer competence through bio-mediated ZnO nanoparticles. Mater. Sci. Eng. C 2019, 103, 109756. [Google Scholar] [CrossRef] [PubMed]
  22. Madan, H.; Sharma, S.; Suresh, D.; Vidya, Y.; Nagabhushana, H.; Rajanaik, H.; Anantharaju, K.; Prashantha, S.; Maiya, P.S. Facile green fabrication of nanostructure ZnO plates, bullets, flower, prismatic tip, closed pine cone: Their antibacterial, antioxidant, photoluminescent and photocatalytic properties. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2016, 152, 404–416. [Google Scholar] [CrossRef] [PubMed]
  23. Momeni, S.S.; Nasrollahzadeh, M.; Rustaiyan, A. Green synthesis of the Cu/ZnO nanoparticles mediated by Euphorbia prolifera leaf extract and investigation of their catalytic activity. J. Coll. Interf. Sci. 2016, 472, 173–179. [Google Scholar] [CrossRef] [PubMed]
  24. Ali, K.; Dwivedi, S.; Azam, A.; Saquib, Q.; Al-Said, M.S.; Alkhedhairy, A.A.; Musarrat, J. Aloe vera extract functionalized zinc oxide nanoparticles as nanoantibiotics against multi-drug resistant clinical bacterial isolates. J. Coll. Interf. Sci. 2016, 472, 145–156. [Google Scholar] [CrossRef]
  25. Mansour, A.T.; Alprol, A.E.; Abualnaja, K.M.; El-Beltagi, H.S.; Ramadan, K.M.A.; Ashour, M. Dried Brown Seaweed’s Phytoremediation Potential for Methylene Blue Dye Removal from Aquatic Environments. Polymers 2022, 14, 1375. [Google Scholar] [CrossRef] [PubMed]
  26. Ashour, M.; Alprol, A.E.; Khedawy, M.; Abualnaja, K.M.; Mansour, A.T. Equilibrium and Kinetic Modeling of Crystal Violet Dye Adsorption by a Marine Diatom, Skeletonema costatum. Materials 2022, 15, 6375. [Google Scholar] [CrossRef] [PubMed]
  27. Ashour, M.; Alprol, A.E.; Heneash, A.M.M.; Saleh, H.; Abualnaja, K.M.; Alhashmialameer, D.; Mansour, A.T. Ammonia Bioremediation from Aquaculture Wastewater Effluents Using Arthrospira platensis NIOF17/003: Impact of Biodiesel Residue and Potential of Ammonia-Loaded Biomass as Rotifer Feed. Materials 2021, 14, 5460. [Google Scholar] [CrossRef] [PubMed]
  28. Miettinen, M.; Khan, S.A. Pharmaceutical pollution: A weakly regulated global environmental risk. Rev. Eur. Comp. Int. Environ. Law 2022, 31, 75–88. [Google Scholar] [CrossRef]
  29. Al-Qahtani, K.M. Cadmium removal from aqueous solution by green synthesis zero valent silver nanoparticles with Benjamina leaves extract. Egypt. J. Aquat. Res. 2017, 43, 269–274. [Google Scholar] [CrossRef]
  30. Al-Senani, G.M.; Al-Fawzan, F.F. Adsorption study of heavy metal ions from aqueous solution by nanoparticle of wild herbs. Egypt. J. Aquat. Res. 2018, 44, 187–194. [Google Scholar] [CrossRef]
  31. Seabra, A.B.; Durán, N. Nanotoxicology of metal oxide nanoparticles. Metals 2015, 5, 934–975. [Google Scholar] [CrossRef]
  32. Vrček, I.V.; Žuntar, I.; Petlevski, R.; Pavičić, I.; Dutour Sikirić, M.; Ćurlin, M.; Goessler, W. Comparison of in vitro toxicity of silver ions and silver nanoparticles on human hepatoma cells. Environ. Toxicol. 2016, 31, 679–692. [Google Scholar] [CrossRef] [PubMed]
  33. Gour, A.; Jain, N.K. Advances in green synthesis of nanoparticles. Artif. Cells Nanomed. Biotechnol. 2019, 47, 844–851. [Google Scholar] [CrossRef] [PubMed]
  34. Prasad, R. Microbial Nanobionics; Springer: Berlin/Heidelberg, Germany, 2019. [Google Scholar]
  35. Pal, G.; Rai, P.; Pandey, A. Green synthesis of nanoparticles: A greener approach for a cleaner future. In Green Synthesis, Characterization and Applications of Nanoparticles; Elsevier: Amsterdam, The Netherlands, 2019; pp. 1–26. [Google Scholar]
  36. Yu, S.; Liu, J.; Yin, Y.; Shen, M. Interactions between engineered nanoparticles and dissolved organic matter: A review on mechanisms and environmental effects. J. Environ. Sci. 2018, 63, 198–217. [Google Scholar] [CrossRef] [PubMed]
  37. Srivastava, M.; Singh, J.; Mishra, R.K.; Ojha, A.K. Electro-optical and magnetic properties of monodispersed colloidal Cu2O nanoparticles. J. Alloys Compd. 2013, 555, 123–130. [Google Scholar] [CrossRef]
  38. De Montferrand, C.; Hu, L.; Milosevic, I.; Russier, V.; Bonnin, D.; Motte, L.; Brioude, A.; Lalatonne, Y. Iron oxide nanoparticles with sizes, shapes and compositions resulting in different magnetization signatures as potential labels for multiparametric detection. Acta Biomater. 2013, 9, 6150–6157. [Google Scholar] [CrossRef]
  39. Uzair, B.; Liaqat, A.; Iqbal, H.; Menaa, B.; Razzaq, A.; Thiripuranathar, G.; Fatima Rana, N.; Menaa, F. Green and cost-effective synthesis of metallic nanoparticles by algae: Safe methods for translational medicine. Bioengineering 2020, 7, 129. [Google Scholar] [CrossRef]
  40. Ahmad, P.; Tripathi, D.K.; Sharma, S.; Chauhan, D.K.; Dubey, N. Nanomaterials in Plants, Algae and Microorganisms; Elsevier: Amsterdam, The Netherlands, 2017. [Google Scholar]
  41. Patra, J.K.; Baek, K.-H. Green nanobiotechnology: Factors affecting synthesis and characterization techniques. J. Nanomat. 2014, 2014, 417305. [Google Scholar] [CrossRef]
  42. Kim, D.-Y.; Saratale, R.G.; Shinde, S.; Syed, A.; Ameen, F.; Ghodake, G. Green synthesis of silver nanoparticles using Laminaria japonica extract: Characterization and seedling growth assessment. J. Clean. Prod. 2018, 172, 2910–2918. [Google Scholar] [CrossRef]
  43. Bhattacharya, P.; Swarnakar, S.; Ghosh, S.; Majumdar, S.; Banerjee, S. Disinfection of drinking water via algae mediated green synthesized copper oxide nanoparticles and its toxicity evaluation. J. Environ. Chem. Eng. 2019, 7, 102867. [Google Scholar] [CrossRef]
  44. Singh, M.; Sharma, N.K.; Prasad, S.B.; Yadav, S.S.; Narayan, G.; Rai, A.K. The freshwater cyanobacterium Anabaena doliolum transformed with ApGSMT-DMT exhibited enhanced salt tolerance and protection to nitrogenase activity, but became halophilic. Microbiology 2013, 159, 641–648. [Google Scholar] [CrossRef] [PubMed]
  45. Kaur, P.; Thakur, R.; Duhan, J.S.; Chaudhury, A. Management of wilt disease of chickpea in vivo by silver nanoparticles biosynthesized by rhizospheric microflora of chickpea (Cicer arietinum). J. Chem. Technol. Biotechnol. 2018, 93, 3233–3243. [Google Scholar] [CrossRef]
  46. Salam, H.A.; Rajiv, P.; Kamaraj, M.; Jagadeeswaran, P.; Gunalan, S.; Sivaraj, R. Plants: Green route for nanoparticle synthesis. Int. Res. J. Biol. Sci. 2012, 1, 85–90. [Google Scholar]
  47. Cauerhff, A.; Castro, G.R. Bionanoparticles, a green nanochemistry approach. Electr. J. Biotechnol. 2013, 16, 11. [Google Scholar]
  48. Ruddaraju, L.K.; Pammi, S.V.N.; sankar Guntuku, G.; Padavala, V.S.; Kolapalli, V.R.M. A review on anti-bacterials to combat resistance: From ancient era of plants and metals to present and future perspectives of green nano technological combinations. Asian J. Pharm. Sci. 2020, 15, 42–59. [Google Scholar] [CrossRef] [PubMed]
  49. Kumar, B.; Smita, K.; Cumbal, L.; Debut, A. Green synthesis of silver nanoparticles using Andean blackberry fruit extract. Saudi J. Biol. Sci. 2017, 24, 45–50. [Google Scholar] [CrossRef]
  50. Yadav, M.; Kaur, P. A review on exploring phytosynthesis of silver and gold nanoparticles using genus Brassica. Int. J. Nanopart. 2018, 10, 165–177. [Google Scholar] [CrossRef]
  51. Predescu, A.M.; Matei, E.; Berbecaru, A.C.; Pantilimon, C.; Drăgan, C.; Vidu, R.; Predescu, C.; Kuncser, V. Synthesis and characterization of dextran-coated iron oxide nanoparticles. R. Soc. Open Sci. 2018, 5, 171525. [Google Scholar] [CrossRef]
  52. Hussain, I.; Singh, N.; Singh, A.; Singh, H.; Singh, S. Green synthesis of nanoparticles and its potential application. Biotechnol. Lett. 2016, 38, 545–560. [Google Scholar] [CrossRef]
  53. Stankic, S.; Suman, S.; Haque, F.; Vidic, J. Pure and multi metal oxide nanoparticles: Synthesis, antibacterial and cytotoxic properties. J. Nanobiotechnol. 2016, 14, 73. [Google Scholar] [CrossRef]
  54. Kaur, P.; Thakur, R.; Chaudhury, A. Biogenesis of copper nanoparticles using peel extract of Punica granatum and their antimicrobial activity against opportunistic pathogens. Green Chem. Lett. Rev. 2016, 9, 33–38. [Google Scholar] [CrossRef]
  55. Ahmed, B.; Hashmi, A.; Khan, M.S.; Musarrat, J. ROS mediated destruction of cell membrane, growth and biofilms of human bacterial pathogens by stable metallic AgNPs functionalized from bell pepper extract and quercetin. Adv. Powder Technol. 2018, 29, 1601–1616. [Google Scholar] [CrossRef]
  56. Sharma, D.; Sabela, M.I.; Kanchi, S.; Mdluli, P.S.; Singh, G.; Stenström, T.A.; Bisetty, K. Biosynthesis of ZnO nanoparticles using Jacaranda mimosifolia flowers extract: Synergistic antibacterial activity and molecular simulated facet specific adsorption studies. J. Photochem. Photobiol. B Biol. 2016, 162, 199–207. [Google Scholar] [CrossRef]
  57. Al-Ruqeishi, M.S.; Mohiuddin, T.; Al-Saadi, L.K. Green synthesis of iron oxide nanorods from deciduous Omani mango tree leaves for heavy oil viscosity treatment. Arab. J. Chem. 2019, 12, 4084–4090. [Google Scholar] [CrossRef]
  58. Kumar, B.; Smita, K.; Cumbal, L.; Debut, A. Green approach for fabrication and applications of zinc oxide nanoparticles. Bioinorg. Chem. Applic. 2014, 2014, 523869. [Google Scholar] [CrossRef] [PubMed]
  59. Panhwar, S.; Buledi, J.A.; Mal, D.; Solangi, A.R.; Balouch, A.; Hyder, A. Importance and analytical perspective of green synthetic strategies of copper, zinc, and titanium oxide nanoparticles and their applications in pathogens and environmental remediation. Curr. Anal. Chem. 2021, 17, 1169–1181. [Google Scholar] [CrossRef]
  60. Sharma, J.K.; Srivastava, P.; Ameen, S.; Akhtar, M.S.; Singh, G.; Yadava, S. Azadirachta indica plant-assisted green synthesis of Mn3O4 nanoparticles: Excellent thermal catalytic performance and chemical sensing behavior. J. Coll. Interf. Sci. 2016, 472, 220–228. [Google Scholar] [CrossRef] [PubMed]
  61. Bhuyan, T.; Mishra, K.; Khanuja, M.; Prasad, R.; Varma, A. Biosynthesis of zinc oxide nanoparticles from Azadirachta indica for antibacterial and photocatalytic applications. Mater. Sci. Semicond. Process. 2015, 32, 55–61. [Google Scholar] [CrossRef]
  62. Matinise, N.; Fuku, X.; Kaviyarasu, K.; Mayedwa, N.; Maaza, M. ZnO nanoparticles via Moringa oleifera green synthesis: Physical properties & mechanism of formation. Appl. Surf. Sci. 2017, 406, 339–347. [Google Scholar]
  63. Surendra, T.; Roopan, S.M. Photocatalytic and antibacterial properties of phytosynthesized CeO2 NPs using Moringa oleifera peel extract. J. Photochem. Photobiol. B Biol. 2016, 161, 122–128. [Google Scholar] [CrossRef]
  64. Moon, S.A.; Salunke, B.K.; Alkotaini, B.; Sathiyamoorthi, E.; Kim, B.S. Biological synthesis of manganese dioxide nanoparticles by Kalopanax pictus plant extract. IET Nanobiotechnol. 2015, 9, 220–225. [Google Scholar] [CrossRef]
  65. Vidya, C.; Prabha, M.C.; Raj, M.A. Green mediated synthesis of zinc oxide nanoparticles for the photocatalytic degradation of Rose Bengal dye. Environ. Nanotechnol. Monit. Manag. 2016, 6, 134–138. [Google Scholar]
  66. Stan, M.; Popa, A.; Toloman, D.; Dehelean, A.; Lung, I.; Katona, G. Enhanced photocatalytic degradation properties of zinc oxide nanoparticles synthesized by using plant extracts. Mater. Sci. Semicond. Process. 2015, 39, 23–29. [Google Scholar] [CrossRef]
  67. Suresh, D.; Shobharani, R.; Nethravathi, P.; Kumar, M.P.; Nagabhushana, H.; Sharma, S. Artocarpus gomezianus aided green synthesis of ZnO nanoparticles: Luminescence, photocatalytic and antioxidant properties. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 141, 128–134. [Google Scholar] [CrossRef]
  68. Kumar, P.V.; Pammi, S.; Shameem, U. A Green approach for the synthesis of iron oxide nanoparticles by using roots of A. racemosus and its deg-radation of dye methyl orange. Int. J. Pharm. Drug Anal. 2018, 6, 22–28. [Google Scholar]
  69. Ardakani, L.S.; Alimardani, V.; Tamaddon, A.M.; Amani, A.M.; Taghizadeh, S. Green synthesis of iron-based nanoparticles using Chlorophytum comosum leaf extract: Methyl orange dye degradation and antimicrobial properties. Heliyon 2021, 7, 1. [Google Scholar]
  70. Karpagavinayagam, P.; Vedhi, C. Green synthesis of iron oxide nanoparticles using Avicennia marina flower extract. Vacuum 2019, 160, 286–292. [Google Scholar] [CrossRef]
  71. Lohrasbi, S.; Kouhbanani, M.A.J.; Beheshtkhoo, N.; Ghasemi, Y.; Amani, A.M.; Taghizadeh, S. Green synthesis of iron nanoparticles using Plantago major leaf extract and their application as a catalyst for the decolorization of azo dye. BioNanoScience 2019, 9, 317–322. [Google Scholar] [CrossRef]
  72. Rostamizadeh, E.; Iranbakhsh, A.; Majd, A. Green synthesis of Fe2O3 nanoparticles using fruit extract of Cornus mas L. and its growth-promoting roles in barley. J. Nanostruct. Chem. 2020, 10, 125–130. [Google Scholar] [CrossRef]
  73. Bibi, I.; Nazar, N.; Ata, S.; Sultan, M.; Ali, A.; Abbas, A.; Jilani, K.; Kamal, S.; Sarim, F.M.; Khan, M.I. Green synthesis of iron oxide nanoparticles using pomegranate seeds extract and photocatalytic activity evaluation for the degradation of textile dye. J. Mater. Res. Technol. 2019, 8, 6115–6124. [Google Scholar] [CrossRef]
  74. Jamzad, M.; Kamari Bidkorpeh, M. Green synthesis of iron oxide nanoparticles by the aqueous extract of Laurus nobilis L. leaves and evaluation of the antimicrobial activity. J. Nanostructure Chem. 2020, 10, 193–201. [Google Scholar] [CrossRef]
  75. Belaiche, Y.; Khelef, A.; Laouini, S.E.; Bouafia, A.; Tedjani, M.L.; Barhoum, A. Green synthesis and characterization of silver/silver oxide nanoparticles using aqueous leaves extract of Artemisia herba-alba as reducing and capping agents. Rev. Romana Mater. 2021, 51, 342–352. [Google Scholar]
  76. Roopan, S.M.; Bharathi, A.; Prabhakarn, A.; Rahuman, A.A.; Velayutham, K.; Rajakumar, G.; Padmaja, R.; Lekshmi, M.; Madhumitha, G. Efficient phyto-synthesis and structural characterization of rutile TiO2 nanoparticles using Annona squamosa peel extract. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2012, 98, 86–90. [Google Scholar] [CrossRef]
  77. Fu, L.; Fu, Z. Plectranthus amboinicus leaf extract–assisted biosynthesis of ZnO nanoparticles and their photocatalytic activity. Ceram. Int. 2015, 41, 2492–2496. [Google Scholar] [CrossRef]
  78. Venkateswarlu, S.; Rao, Y.S.; Balaji, T.; Prathima, B.; Jyothi, N. Biogenic synthesis of Fe3O4 magnetic nanoparticles using plantain peel extract. Mater. Lett. 2013, 100, 241–244. [Google Scholar] [CrossRef]
  79. Vidhu, V.; Philip, D. Biogenic synthesis of SnO2 nanoparticles: Evaluation of antibacterial and antioxidant activities. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 134, 372–379. [Google Scholar] [CrossRef]
  80. Anbuvannan, M.; Ramesh, M.; Viruthagiri, G.; Shanmugam, N.; Kannadasan, N. Synthesis, characterization and photocatalytic activity of ZnO nanoparticles prepared by biological method. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 143, 304–308. [Google Scholar] [CrossRef]
  81. Sankar, R.; Manikandan, P.; Malarvizhi, V.; Fathima, T.; Shivashangari, K.S.; Ravikumar, V. Green synthesis of colloidal copper oxide nanoparticles using Carica papaya and its application in photocatalytic dye degradation. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 121, 746–750. [Google Scholar] [CrossRef]
  82. Diallo, A.; Manikandan, E.; Rajendran, V.; Maaza, M. Physical & enhanced photocatalytic properties of green synthesized SnO2 nanoparticles via Aspalathus linearis. J. Alloys Compd. 2016, 681, 561–570. [Google Scholar]
  83. Suresh, D.; Nethravathi, P.; Rajanaika, H.; Nagabhushana, H.; Sharma, S. Green synthesis of multifunctional zinc oxide (ZnO) nanoparticles using Cassia fistula plant extract and their photodegradative, antioxidant and antibacterial activities. Mater. Sci. Semicond. Proces. 2015, 31, 446–454. [Google Scholar] [CrossRef]
  84. Thovhogi, N.; Park, E.; Manikandan, E.; Maaza, M.; Gurib-Fakim, A. Physical properties of CdO nanoparticles synthesized by green chemistry via Hibiscus Sabdariffa flower extract. J. Alloys Compd. 2016, 655, 314–320. [Google Scholar] [CrossRef]
  85. Ijaz, F.; Shahid, S.; Khan, S.A.; Ahmad, W.; Zaman, S. Green synthesis of copper oxide nanoparticles using Abutilon indicum leaf extract: Antimicrobial, antioxidant and photocatalytic dye degradation activitie. Trop. J. Pharm. Res. 2017, 16, 743–753. [Google Scholar] [CrossRef]
  86. Malleshappa, J.; Nagabhushana, H.; Sharma, S.; Vidya, Y.; Anantharaju, K.; Prashantha, S.; Prasad, B.D.; Naika, H.R.; Lingaraju, K.; Surendra, B. Leucas aspera mediated multifunctional CeO2 nanoparticles: Structural, photoluminescent, photocatalytic and antibacterial properties. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 149, 452–462. [Google Scholar] [CrossRef]
  87. Osuntokun, J.; Onwudiwe, D.C.; Ebenso, E.E. Biosynthesis and photocatalytic properties of SnO2 nanoparticles prepared using aqueous extract of cauliflower. J. Clust. Sci. 2017, 28, 1883–1896. [Google Scholar] [CrossRef]
  88. Kumar, P.N.; Sakthivel, K.; Balasubramanian, V. Microwave assisted biosynthesis of rice shaped ZnO nanoparticles using Amorphophallus konjac tuber extract and its application in dye sensitized solar cells. Mater. Sci. Pol. 2017, 35, 111–119. [Google Scholar] [CrossRef]
  89. Khanna, P.; Kaur, A.; Goyal, D. Algae-based metallic nanoparticles: Synthesis, characterization and applications. J. Microbiol. Methods 2019, 163, 105656. [Google Scholar] [CrossRef]
  90. Ramanan, V.; Thiyagarajan, S.K.; Raji, K.; Suresh, R.; Sekar, R.; Ramamurthy, P. Outright green synthesis of fluorescent carbon dots from eutrophic algal blooms for in vitro imaging. ACS Sustain. Chem. Eng. 2016, 4, 4724–4731. [Google Scholar] [CrossRef]
  91. Patil, M.P.; Kim, G.-D. Marine microorganisms for synthesis of metallic nanoparticles and their biomedical applications. Colloids Surf. B Biointerfaces 2018, 172, 487–495. [Google Scholar] [CrossRef]
  92. Mourdikoudis, S.; Pallares, R.M.; Thanh, N.T. Characterization techniques for nanoparticles: Comparison and complementarity upon studying nanoparticle properties. Nanoscale 2018, 10, 12871–12934. [Google Scholar] [CrossRef]
  93. Ghoneim, M.M.; El-Desoky, H.S.; El-Moselhy, K.M.; Amer, A.; Abou El-Naga, E.H.; Mohamedein, L.I.; Al-Prol, A.E. Removal of cadmium from aqueous solution using marine green algae, Ulva lactuca. Egypt. J. Aquat. Res. 2014, 40, 235–242. [Google Scholar] [CrossRef]
  94. Yasir, M.; Šopík, T.; Ali, H.; Kimmer, D.; Sedlařík, V. Green synthesis of titanium and zinc oxide nanoparticles for simultaneous photocatalytic removal of estrogens in wastewater. In Proceedings of the NANOCON 2021—International Conference on Nanomaterials, Brno, Czech Republic, 20–22 October 2021. [Google Scholar]
  95. Ibrahim, R.K.; Hayyan, M.; AlSaadi, M.A.; Hayyan, A.; Ibrahim, S. Environmental application of nanotechnology: Air, soil, and water. Environ. Sci. Pollut. Res. 2016, 23, 13754–13788. [Google Scholar] [CrossRef]
  96. Srivastava, V.; Gusain, D.; Sharma, Y.C. Synthesis, characterization and application of zinc oxide nanoparticles (n-ZnO). Ceram. Int. 2013, 39, 9803–9808. [Google Scholar] [CrossRef]
  97. Ehrampoush, M.H.; Miria, M.; Salmani, M.H.; Mahvi, A.H. Cadmium removal from aqueous solution by green synthesis iron oxide nanoparticles with tangerine peel extract. J. Environ. Health Sci. Eng. 2015, 13, 84. [Google Scholar] [CrossRef]
  98. Fazlzadeh, M.; Khosravi, R.; Zarei, A. Green synthesis of zinc oxide nanoparticles using Peganum harmala seed extract, and loaded on Peganum harmala seed powdered activated carbon as new adsorbent for removal of Cr (VI) from aqueous solution. Ecol. Eng. 2017, 103, 180–190. [Google Scholar] [CrossRef]
  99. Mukherjee, D.; Ghosh, S.; Majumdar, S.; Annapurna, K. Green synthesis of α-Fe2O3 nanoparticles for arsenic (V) remediation with a novel aspect for sludge management. J. Environ. Chem. Eng. 2016, 4, 639–650. [Google Scholar] [CrossRef]
  100. Somu, P.; Paul, S. Casein based biogenic-synthesized zinc oxide nanoparticles simultaneously decontaminate heavy metals, dyes, and pathogenic microbes: A rational strategy for wastewater treatment. J. Chem. Technol. Biotechnol. 2018, 93, 2962–2976. [Google Scholar] [CrossRef]
  101. Azizi, S.; Mahdavi Shahri, M.; Mohamad, R. Green synthesis of zinc oxide nanoparticles for enhanced adsorption of lead ions from aqueous solutions: Equilibrium, kinetic and thermodynamic studies. Molecules 2017, 22, 831. [Google Scholar] [CrossRef]
  102. Venkateswarlu, S.; Yoon, M. Rapid removal of cadmium ions using green-synthesized Fe3O4 nanoparticles capped with diethyl-4-(4 amino-5-mercapto-4 H-1, 2, 4-triazol-3-yl) phenyl phosphonate. RSC Adv. 2015, 5, 65444–65453. [Google Scholar] [CrossRef]
  103. Gebre, S.H.; Sendeku, M.G. New frontiers in the biosynthesis of metal oxide nanoparticles and their environmental applications: An overview. SN Appl. Sci. 2019, 1, 928. [Google Scholar] [CrossRef]
  104. Dubey, S.; Sharma, Y.C. Calotropis procera mediated one pot green synthesis of Cupric oxide nanoparticles (CuO-NPs) for adsorptive removal of Cr (VI) from aqueous solutions. Appl. Organomet. Chem. 2017, 31, e3849. [Google Scholar] [CrossRef]
  105. Goutam, S.P.; Saxena, G.; Singh, V.; Yadav, A.K.; Bharagava, R.N.; Thapa, K.B. Green synthesis of TiO2 nanoparticles using leaf extract of Jatropha curcas L. for photocatalytic degradation of tannery wastewater. Chem. Eng. J. 2018, 336, 386–396. [Google Scholar] [CrossRef]
  106. Singh, J.; Kumar, V.; Kim, K.-H.; Rawat, M. Biogenic synthesis of copper oxide nanoparticles using plant extract and its prodigious potential for photocatalytic degradation of dyes. Environ. Res. 2019, 177, 108569. [Google Scholar] [CrossRef]
  107. Khani, R.; Roostaei, B.; Bagherzade, G.; Moudi, M. Green synthesis of copper nanoparticles by fruit extract of Ziziphus spina-christi (L.) Willd.: Application for adsorption of triphenylmethane dye and antibacterial assay. J. Mol. Liq. 2018, 255, 541–549. [Google Scholar] [CrossRef]
  108. Sathyavathi, S.; Manjula, A.; Rajendhran, J.; Gunasekaran, P. Extracellular synthesis and characterization of nickel oxide nanoparticles from Microbacterium sp. MRS-1 towards bioremediation of nickel electroplating industrial effluent. Bioresour. Technol. 2014, 165, 270–273. [Google Scholar] [CrossRef]
  109. Mahmoud, A.E.D.; Al-Qahtani, K.M.; Alflaij, S.O.; Al-Qahtani, S.F.; Alsamhan, F.A. Green copper oxide nanoparticles for lead, nickel, and cadmium removal from contaminated water. Sci. Rep. 2021, 11, 12547. [Google Scholar] [CrossRef]
  110. Vidovix, T.B.; Quesada, H.B.; Bergamasco, R.; Vieira, M.F.; Vieira, A.M.S. Adsorption of Safranin-O dye by copper oxide nanoparticles synthesized from Punica granatum leaf extract. Environ. Technol. 2022, 43, 3047–3063. [Google Scholar] [CrossRef]
  111. Debnath, P.; Mondal, N.K. Effective removal of congo red dye from aqueous solution using biosynthesized zinc oxide nanoparticles. Environ. Nanotechnol. Monit. Manag. 2020, 14, 100320. [Google Scholar] [CrossRef]
  112. Nayak, A.; Sahoo, J.K.; Sahoo, S.K.; Sahu, D. Removal of congo red dye from aqueous solution using zinc oxide nanoparticles synthesised from Ocimum sanctum (Tulsi leaf): A green approach. Int. J. Environ. Anal. Chem. 2022, 102, 7889–7910. [Google Scholar] [CrossRef]
  113. Mittal, H.; Morajkar, P.P.; Al Alili, A.; Alhassan, S.M. In-situ synthesis of ZnO nanoparticles using gum arabic based hydrogels as a self-template for effective malachite green dye adsorption. J. Polym. Environ. 2020, 28, 1637–1653. [Google Scholar] [CrossRef]
  114. Suriyaraj, S.; Selvakumar, R. Advances in nanomaterial based approaches for enhanced fluoride and nitrate removal from contaminated water. RSC Adv. 2016, 6, 10565–10583. [Google Scholar] [CrossRef]
  115. Zeng, Y.; Xue, Y.; Liang, S.; Zhang, J. Removal of fluoride from aqueous solution by TiO2 and TiO2–SiO2 nanocomposite. Chem. Speciat. Bioavailab. 2017, 29, 25–32. [Google Scholar] [CrossRef]
  116. Chai, L.; Wang, Y.; Zhao, N.; Yang, W.; You, X. Sulfate-doped Fe3O4/Al2O3 nanoparticles as a novel adsorbent for fluoride removal from drinking water. Water Res. 2013, 47, 4040–4049. [Google Scholar] [CrossRef] [PubMed]
  117. Silveira, C.; Shimabuku, Q.L.; Fernandes Silva, M.; Bergamasco, R. Iron-oxide nanoparticles by the green synthesis method using Moringa oleifera leaf extract for fluoride removal. Environ. Technol. 2018, 39, 2926–2936. [Google Scholar] [CrossRef] [PubMed]
  118. Kumari, S.; Khan, S. Defluoridation technology for drinking water and tea by green synthesized Fe3O4/Al2O3 nanoparticles coated polyurethane foams for rural communities. Sci. Rep. 2017, 7, 8070. [Google Scholar] [CrossRef] [PubMed]
  119. Zandsalimi, Y.; Taymori, P.; Darvishi Cheshmeh Soltani, R.; Rezaee, R.; Abdullahi, N.; Safari, M. Photocatalytic removal of Acid Red 88 dye using zinc oxide nanoparticles fixed on glass plates. J. Adv. Environ. Health Res. 2015, 3, 102–110. [Google Scholar]
  120. Yasir, M.; Ali, H.; Masar, M.; Ngwabebhoh, F.A.; Zubair, M.; Sopik, T.; Machovsky, M.; Kuritka, I.; Sedlarik, V. Design and fabrication of TiO2/Nd polyurethane nanofibers based photoreactor: A continuous flow kinetics study for Estriol degradation and mechanism. J. Water Process Eng. 2023, 56, 104271. [Google Scholar] [CrossRef]
  121. Wei, Y.; Hao, J.-G.; Zhang, J.-L.; Huang, W.-Y.; Ouyang, S.-b.; Yang, K.; Lu, K.-Q. Integrating Co (OH) 2 nanosheet arrays on graphene for efficient noble-metal-free EY-sensitized photocatalytic H2 evolution. Dalton Trans. 2023, 52, 13923–13929. [Google Scholar] [CrossRef]
  122. Chen, H.-Q.; Hao, J.-G.; Wei, Y.; Huang, W.-Y.; Zhang, J.-L.; Deng, T.; Yang, K.; Lu, K.-Q. Recent Developments and Perspectives of Cobalt Sulfide-Based Composite Materials in Photocatalysis. Catalysts 2023, 13, 544. [Google Scholar] [CrossRef]
  123. Subramanian, H.; Krishnan, M.; Mahalingam, A. Photocatalytic dye degradation and photoexcited anti-microbial activities of green zinc oxide nanoparticles synthesized via Sargassum muticum extracts. RSC Adv. 2022, 12, 985–997. [Google Scholar] [CrossRef]
  124. Pugazhendhi, A.; Prabhu, R.; Muruganantham, K.; Shanmuganathan, R.; Natarajan, S. Anticancer, antimicrobial and photocatalytic activities of green synthesized magnesium oxide nanoparticles (MgONPs) using aqueous extract of Sargassum wightii. J. Photochem. Photobiol. B Biol. 2019, 190, 86–97. [Google Scholar] [CrossRef]
  125. Barakat, M. New trends in removing heavy metals from industrial wastewater. Arab. J. Chem. 2011, 4, 361–377. [Google Scholar] [CrossRef]
  126. Fosso-Kankeu, E.; Pandey, S.; Ray, S.S. Photocatalysts in Advanced Oxidation Processes for Wastewater Treatment; John Wiley & Sons: Hoboken, NJ, USA, 2020. [Google Scholar]
  127. Singh, N.; Chakraborty, R.; Gupta, R.K. Mutton bone derived hydroxyapatite supported TiO2 nanoparticles for sustainable photocatalytic applications. J. Environ. Chem. Eng. 2018, 6, 459–467. [Google Scholar] [CrossRef]
  128. Buazar, F.; Alipouryan, S.; Kroushawi, F.; Hossieni, S. Photodegradation of odorous 2-mercaptobenzoxazole through zinc oxide/hydroxyapatite nanocomposite. Appl. Nanosci. 2015, 5, 719–729. [Google Scholar] [CrossRef]
  129. Rezaee, A.; Rangkooy, H.; Khavanin, A.; Jafari, A.J. High photocatalytic decomposition of the air pollutant formaldehyde using nano-ZnO on bone char. Environ. Chem. Lett. 2014, 12, 353–357. [Google Scholar] [CrossRef]
  130. Hassan, S.S.; El Azab, W.I.; Ali, H.R.; Mansour, M.S. Green synthesis and characterization of ZnO nanoparticles for photocatalytic degradation of anthracene. Adv. Nat. Sci. Nanosci. Nanotechnol. 2015, 6, 045012. [Google Scholar] [CrossRef]
  131. Theerthagiri, J.; Chandrasekaran, S.; Salla, S.; Elakkiya, V.; Senthil, R.; Nithyadharseni, P.; Maiyalagan, T.; Micheal, K.; Ayeshamariam, A.; Arasu, M.V. Recent developments of metal oxide based heterostructures for photocatalytic applications towards environmental remediation. J. Solid State Chem. 2018, 267, 35–52. [Google Scholar] [CrossRef]
  132. Jayaseelan, C.; Rahuman, A.A.; Kirthi, A.V.; Marimuthu, S.; Santhoshkumar, T.; Bagavan, A.; Gaurav, K.; Karthik, L.; Rao, K.B. Novel microbial route to synthesize ZnO nanoparticles using Aeromonas hydrophila and their activity against pathogenic bacteria and fungi. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2012, 90, 78–84. [Google Scholar] [CrossRef]
  133. Zheng, Y.; Fu, L.; Han, F.; Wang, A.; Cai, W.; Yu, J.; Yang, J.; Peng, F. Green biosynthesis and characterization of zinc oxide nanoparticles using Corymbia citriodora leaf extract and their photocatalytic activity. Green Chem. Lett. Rev. 2015, 8, 59–63. [Google Scholar] [CrossRef]
  134. Fulekar, J.; Dutta, D.P.; Pathak, B.; Fulekar, M.H. Novel microbial and root mediated green synthesis of TiO2 nanoparticles and its application in wastewater remediation. J. Chem. Technol. Biotechnol. 2018, 93, 736–743. [Google Scholar] [CrossRef]
  135. Muthukumar, H.; Matheswaran, M. Amaranthus spinosus leaf extract mediated FeO nanoparticles: Physicochemical traits, photocatalytic and antioxidant activity. ACS Sustain. Chem. Eng. 2015, 3, 3149–3156. [Google Scholar] [CrossRef]
  136. Rather, M.Y.; Sundarapandian, S. Magnetic iron oxide nanorod synthesis by Wedelia urticifolia (Blume) DC. leaf extract for methylene blue dye degradation. Appl. Nanosci. 2020, 10, 2219–2227. [Google Scholar] [CrossRef]
  137. Prasad, C.; Yuvaraja, G.; Venkateswarlu, P. Biogenic synthesis of Fe3O4 magnetic nanoparticles using Pisum sativum peels extract and its effect on magnetic and Methyl orange dye degradation studies. J. Magn. Magn. Mater. 2017, 424, 376–381. [Google Scholar] [CrossRef]
  138. Kumar, B.; Smita, K.; Galeas, S.; Sharma, V.; Guerrero, V.H.; Debut, A.; Cumbal, L. Characterization and application of biosynthesized iron oxide nanoparticles using Citrus paradisi peel: A sustainable approach. Inorg. Chem. Commun. 2020, 119, 108116. [Google Scholar] [CrossRef]
  139. Nethravathi, P.; Kumar, M.P.; Suresh, D.; Lingaraju, K.; Rajanaika, H.; Nagabhushana, H.; Sharma, S. Tinospora cordifolia mediated facile green synthesis of cupric oxide nanoparticles and their photocatalytic, antioxidant and antibacterial properties. Mater. Sci. Semicond. Proces. 2015, 33, 81–88. [Google Scholar]
  140. Sreekanth, T.; Shim, J.-J.; Lee, Y.R. Degradation of organic pollutants by bio-inspired rectangular and hexagonal titanium dioxide nanostructures. J. Photochem. Photobiol. B Biol. 2017, 169, 90–95. [Google Scholar] [CrossRef]
  141. Ezhilarasi, A.A.; Vijaya, J.J.; Kaviyarasu, K.; Kennedy, L.J.; Ramalingam, R.J.; Al-Lohedan, H.A. Green synthesis of NiO nanoparticles using Aegle marmelos leaf extract for the evaluation of in-vitro cytotoxicity, antibacterial and photocatalytic properties. J. Photochem. Photobiol. B Biol. 2018, 180, 39–50. [Google Scholar] [CrossRef]
  142. Reddy Yadav, L.; Manjunath, K.; Archana, B.; Madhu, C.; Raja Naika, H.; Nagabhushana, H.; Kavitha, C.; Nagaraju, G. Fruit juice extract mediated synthesis of CeO2 nanoparticles for antibacterial and photocatalytic activities. Eur. Phys. J. Plus 2016, 131, 154. [Google Scholar] [CrossRef]
  143. Aminuzzaman, M.; Ying, L.P.; Goh, W.-S.; Watanabe, A. Green synthesis of zinc oxide nanoparticles using aqueous extract of Garcinia mangostana fruit pericarp and their photocatalytic activity. Bull. Mater. Sci. 2018, 41, 50. [Google Scholar] [CrossRef]
  144. Li, S.-W.; Gao, R.-M.; Zhao, J.-s. Deep oxidative desulfurization of fuel catalyzed by modified heteropolyacid: The comparison performance of three kinds of ionic liquids. ACS Sustain. Chem. Eng. 2018, 6, 15858–15866. [Google Scholar] [CrossRef]
  145. Bishnoi, S.; Kumar, A.; Selvaraj, R. Facile synthesis of magnetic iron oxide nanoparticles using inedible Cynometra ramiflora fruit extract waste and their photocatalytic degradation of methylene blue dye. Mater. Res. Bull. 2018, 97, 121–127. [Google Scholar] [CrossRef]
  146. Alshehri, A.; Malik, M.A.; Khan, Z.; Al-Thabaiti, S.A.; Hasan, N. Biofabrication of Fe nanoparticles in aqueous extract of Hibiscus sabdariffa with enhanced photocatalytic activities. RSC Adv. 2017, 7, 25149–25159. [Google Scholar] [CrossRef]
  147. Vasantharaj, S.; Sathiyavimal, S.; Senthilkumar, P.; LewisOscar, F.; Pugazhendhi, A. Biosynthesis of iron oxide nanoparticles using leaf extract of Ruellia tuberosa: Antimicrobial properties and their applications in photocatalytic degradation. J. Photochem. Photobiol. B Biol. 2019, 192, 74–82. [Google Scholar] [CrossRef] [PubMed]
  148. Shah, Y.; Maharana, M.; Sen, S. Peltophorum pterocarpum leaf extract mediated green synthesis of novel iron oxide particles for application in photocatalytic and catalytic removal of organic pollutants. In Biomass Conversion and Biorefinery; Springer: Berlin/Heidelberg, Germany, 2022; pp. 1–14. [Google Scholar]
  149. Bhattacharjee, A.; Ahmaruzzaman, M. Photocatalytic-degradation and reduction of organic compounds using SnO2 quantum dots (via a green route) under direct sunlight. RSC Adv. 2015, 5, 66122–66133. [Google Scholar] [CrossRef]
  150. Thandapani, K.; Kathiravan, M.; Namasivayam, E.; Padiksan, I.A.; Natesan, G.; Tiwari, M.; Giovanni, B.; Perumal, V. Enhanced larvicidal, antibacterial, and photocatalytic efficacy of TiO2 nanohybrids green synthesized using the aqueous leaf extract of Parthenium hysterophorus. Environ. Sci. Pollut. Res. 2018, 25, 10328–10339. [Google Scholar] [CrossRef]
  151. Reddy Yadav, L.; Lingaraju, K.; Daruka Prasad, B.; Kavitha, C.; Banuprakash, G.; Nagaraju, G. Synthesis of CeO2 nanoparticles: Photocatalytic and antibacterial activities. Eur. Phys. J. Plus 2017, 132, 239. [Google Scholar] [CrossRef]
  152. Lingaraju, K.; Naika, H.R.; Manjunath, K.; Nagaraju, G.; Suresh, D.; Nagabhushana, H. Rauvolfia serpentina-mediated green synthesis of CuO nanoparticles and its multidisciplinary studies. Acta Metall. Sin. (Engl. Lett.) 2015, 28, 1134–1140. [Google Scholar] [CrossRef]
  153. Siripireddy, B.; Mandal, B.K. Facile green synthesis of zinc oxide nanoparticles by Eucalyptus globulus and their photocatalytic and antioxidant activity. Adv. Powder Technol. 2017, 28, 785–797. [Google Scholar] [CrossRef]
  154. Prasad, A.R.; Ammal, P.R.; Joseph, A. Effective photocatalytic removal of different dye stuffs using green synthesized zinc oxide nanogranules. Mater. Res. Bull. 2018, 102, 116–121. [Google Scholar] [CrossRef]
  155. Nava, O.; Soto-Robles, C.; Gómez-Gutiérrez, C.; Vilchis-Nestor, A.; Castro-Beltrán, A.; Olivas, A.; Luque, P. Fruit peel extract mediated green synthesis of zinc oxide nanoparticles. J. Mol. Struct. 2017, 1147, 1–6. [Google Scholar] [CrossRef]
  156. Abdelmigid, H.M.; Hussien, N.A.; Alyamani, A.A.; Morsi, M.M.; AlSufyani, N.M.; Kadi, H.A. Green synthesis of zinc oxide nanoparticles using pomegranate fruit peel and solid coffee grounds vs. chemical method of synthesis, with their biocompatibility and antibacterial properties investigation. Molecules 2022, 27, 1236. [Google Scholar] [CrossRef]
  157. Kumar, B.; Smita, K.; Cumbal, L.; Debut, A. One pot synthesis and characterization of gold nanocatalyst using Sacha inchi (Plukenetia volubilis) oil: Green approach. J. Photochem. Photobiol. B Biol. 2016, 158, 55–60. [Google Scholar] [CrossRef] [PubMed]
  158. Ullah, H.; Ullah, Z.; Fazal, A.; Irfan, M. Use of vegetable waste extracts for controlling microstructure of CuO nanoparticles: Green synthesis, characterization, and photocatalytic applications. J. Chem. 2017, 2017, 2721798. [Google Scholar] [CrossRef]
  159. Samuel, M.S.; Selvarajan, E.; Mathimani, T.; Santhanam, N.; Phuong, T.N.; Brindhadevi, K.; Pugazhendhi, A. Green synthesis of cobalt-oxide nanoparticle using jumbo Muscadine (Vitis rotundifolia): Characterization and photo-catalytic activity of acid Blue-74. J. Photochem. Photobiol. B Biol. 2020, 211, 112011. [Google Scholar] [CrossRef] [PubMed]
  160. El-Aassar, A.H.; Said, M.; Abdel-Gawad, A.; Shawky, H. Using silver nanoparticles coated on activated carbon granules in columns for microbiological pollutants water disinfection in Abu Rawash area, Great Cairo, Egypt. Aust. J. Basic Appl. Sci. 2013, 7, 422–432. [Google Scholar]
  161. Mostafaii, G.; Chimehi, E.; Gilasi, H.; Iranshahi, L. Investigation of zinc oxide nanoparticles effects on removal of total coliform bacteria in activated sludge process effluent of municipal wastewater. J. Environ. Sci. Technol. 2017, 10, 49–55. [Google Scholar] [CrossRef]
  162. Sharmila, G.; Muthukumaran, C.; Sandiya, K.; Santhiya, S.; Pradeep, R.S.; Kumar, N.M.; Suriyanarayanan, N.; Thirumarimurugan, M. Biosynthesis, characterization, and antibacterial activity of zinc oxide nanoparticles derived from Bauhinia tomentosa leaf extract. J. Nanostruct. Chem. 2018, 8, 293–299. [Google Scholar] [CrossRef]
  163. Kaur, P.; Thakur, R.; Kumar, S.; Dilbaghi, N. Interaction of ZnO nanoparticles with food borne pathogens Escherichia coli DH5α and Staphylococcus aureus 5021 & their bactericidal efficacy. AIP Conf. Proc. 2011, 1393, 153–154. [Google Scholar]
  164. Sankar, R.; Prasath, B.B.; Nandakumar, R.; Santhanam, P.; Shivashangari, K.S.; Ravikumar, V. Growth inhibition of bloom forming cyanobacterium Microcystis aeruginosa by green route fabricated copper oxide nanoparticles. Environ. Sci. Pollut. Res. 2014, 21, 14232–14240. [Google Scholar] [CrossRef]
Figure 1. Various green synthesis approaches for the preparation of MONPs.
Figure 1. Various green synthesis approaches for the preparation of MONPs.
Processes 11 03356 g001
Figure 2. Synthesis of MONPs from plant extracts.
Figure 2. Synthesis of MONPs from plant extracts.
Processes 11 03356 g002
Figure 3. Advantages of green MONPs as adsorbents.
Figure 3. Advantages of green MONPs as adsorbents.
Processes 11 03356 g003
Table 1. The synthesis of MONPs by different plants.
Table 1. The synthesis of MONPs by different plants.
SpeciesPlant PortionMONPsDimensions (nm)FormRefs.
Cynodon dactylonLeafTiO213–34Hexagonal[59]
Azadirachta indicaLeafMn3O418.2Spherical[60]
LeafZnO9.6–25.5Hexagonal wurtzite[61]
Moringa oleiferaLeafZnO12.27–30.51-[62]
PeelCeO245Spherical[63]
Kalopanax pictusLeafMnO219.2Spherical[64]
Artocarpus hetero-LeafZnO15–25Hexagonal wurtzite[65]
Allium sativumBulbZnO14–70Hexagonal wurtzite[66]
Petroselinum crispumLeafZnO14–70Hexagonal wurtzite[66]
Artocarpus gomezi- anusFruitsZnO11.53Spherical, hexagonal wurtzite[67]
Asparagus racemosusRootFe2O430–40Spherical[68]
Chlorophytum comosumLeafFe3O4100-[69]
Punica granatumSeedFe3O425–55Semi-spherical[70]
Avicennia marineFlowerFe3O430–100Honeycomb [71]
Cornelian cherryFruit Fe3O429–37Spherical[72]
BorassusflabelliferSeedFe3O410–40Hexagonal[73]
Laurus nobilisLeafFe3O48.03 ± 8.99Spherical and hexagonal[74]
Punica granatum L.LeafFe3O421–23 nm-[75]
Annona squamosalPeelTiO223 ± 2Spherical[76]
Plectranthus amboinicusLeafZnO50–180 Rod shape[77]
Plantain peelPeelFe3O4<50 Spherical[78]
Saraca indicaFlowerSnO22–4Circular[79]
Phyllanthus niruriLeafZnO25.61Hexagonal wurtzite[80]
Carica papayaLeafCuO140Rod shape[81]
Aspalathus linearisLeafSnO28.23-[82]
Cassia fistulaLeafZnO5–15Hexagonal wurtzite[83]
Hibiscus SabdariffaFlowerCdO113Cuboid shape[84]
Abutilon indicumLeafCuO16.78Hexagonal wurtzite[85]
Leucas asperaLeafCeO24–13Uniform microspheres[86]
Brassica oleraceaLeafSnO23.62–6.34Quasi-spherical[87]
AmorphophallusRootZnO19.6Rice[88]
Table 2. Techniques for MONPs’ characterization.
Table 2. Techniques for MONPs’ characterization.
No.Studied Properties (Information)Characterization Techniques
1Elemental–chemical compositionNMR, SEM-EDX, ICP-MS, XRD, XPS,
2ShapeTEM, AFM
3Chemical state–oxidation stateXPS
4SizeTEM, XRD, DLS, SEM, ICP-MS, UV-Vis
5Crystal structureXRD
6Size distributionDLS, ICP-MS, SEM
7Growth kineticsNMR, TEM
8Surface composition, arrangement, density, ligand bindingNMR, FTIR, TGA, XPS
9Surface areaNMR, BET, Zeta potential
10Agglomeration stateTEM, UV-Vis, Zeta potential, SEM, DLS
113D visualizationAFM, SEM
12Dispersion of MONPs in matricesAFM, TEM, SEM
13Detection of NPsTEM, SEM
14Optical propertiesUV-Vis, PL
Table 3. Removal of different pollutants using biosynthesized MONPs.
Table 3. Removal of different pollutants using biosynthesized MONPs.
Plant
Extract
Green MONPsPollutant EliminationInitial Conc.
(mg L−1)
Contact Time (min)pHAmount NPs
(g L−1)
Removal Efficiency (%)Sorption Capacity
(mg g−1)
Ref.
Peganum harmalaZnO-NPsCr(VI)1000302297.674.7[98]
Aloe veraα-Fe2O3-NPsAs(V)2–30-6–80.5-38.4[99]
CaseinZnO-NPsCd(II)500307285.6156.7[100]
Tangerine peelFe2O3-NPsCd(II)159040.488.7-[97]
Zingiber zerumbetZnO-NPsPb(II)256050.19319.6[101]
Ananas comosusFe3O4-NPsCd(II)601060.196.149.1[102]
Emblicao fficinallisZnO-NPsAs(III)0.0026052.596.7 [103]
Calotropis proceraCuO-NPsCr(VI)5120218.899.9 [104]
Jatropha curcasTiO2-NPsCr(VI)--8.2-76.5-[105]
Psidium guajavaCuO-NPsNile blue 93 [106]
Ziziphus spina-christiCuO NPsCV dye 7.5 min 95 [107]
Microbac-terium spNiONi(II) 12072.17295 [108]
Mint leaf extractCuOPb(II)
Ni(II)
Cd(II)
2060 min60.5 88.8
54.9
15.6
[109]
Punica granatum leaf extractCuOSO dye-180 min6.11 189.5[110]
Table 4. The main advantages and disadvantages of nano-photocatalysts [125].
Table 4. The main advantages and disadvantages of nano-photocatalysts [125].
AdvantagesDisadvantages
Lower toxicityLong duration time
Excellent photoactive properties
Availability
High chemical stabilityLimited applications
Low cost
Eco-friendliness
Table 5. Photocatalytic activities of metal and MONPs.
Table 5. Photocatalytic activities of metal and MONPs.
PlantsPhotocatalystTypes of DyesDegradation (%)Refs.
Banana crustCuONPsCongo Red (CR)88[143]
Amaranthus spinosus, leafFeONPsMethyl Orange (MO)75[144]
FeONPsMethylene Blue (MB)69
Cynometra ramifloraFeONPsMBHigh[145]
Asparagus racemosusFeONPsMOGood[68]
Hibiscus sabdariffa, flowerFeONPsCongo Red96.1%[146]
Ruellia tuberose, leafFeONPsCrystal Violet80%[147]
Peltophorum Pterocarpum, leafFe3O4RhodamineHigh[148]
Sugar cane, juiceSnO2NPsRose Bengal99.3[149]
Sugar cane, juiceSnO2NPsMB96.8
Rhizoma CoptidisTiO2NPsMB71[140]
Parthenium hysterophorusTiO2NPsMB92.5[150]
TiO2NPsMO81.5
TiO2NPsCrystal Violet (CV)79.7
TiO2NPsAlizarin Red77.3
StarchTiO2NPsMB100196
Aspalathus linearisNiONPsMBGood[151]
Eucalyptus globulus
Eucalyptus globulus
ZnONPsMB98.3[152]
ZnONPsMO96.66
Lagerstroemia speciosaZnONPsMO93.5[153]
Cassia fistula, leafZnONPsMB98.71[83]
Artocarpus hetero phyllus, leafZnONPsRose Bengal85[65]
Lemon juiceZnONPsMB91.17[154]
ZnONPsCR90
ZnONPsRhodamine B98
Citrus aurantifolia (lemon), Citrus paradisiZnONPsMB 97[155]
Punica granatum, peelZnOAntibacterial propertyGood[156]
Eucalyptus globulus MB98.3[157]
Moringa oleifera, peelCeO2 CV97.5[63]
CauliflowerSnO2MB 91.89[87]
Cauliflower waste, potato peelCuOMB87.37–96[158]
Vitis rotundifolia, fruitCoOAcid blueHigh[159]
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

Ashour, M.; Mansour, A.T.; Abdelwahab, A.M.; Alprol, A.E. Metal Oxide Nanoparticles’ Green Synthesis by Plants: Prospects in Phyto- and Bioremediation and Photocatalytic Degradation of Organic Pollutants. Processes 2023, 11, 3356. https://doi.org/10.3390/pr11123356

AMA Style

Ashour M, Mansour AT, Abdelwahab AM, Alprol AE. Metal Oxide Nanoparticles’ Green Synthesis by Plants: Prospects in Phyto- and Bioremediation and Photocatalytic Degradation of Organic Pollutants. Processes. 2023; 11(12):3356. https://doi.org/10.3390/pr11123356

Chicago/Turabian Style

Ashour, Mohamed, Abdallah Tageldein Mansour, Abdelwahab M. Abdelwahab, and Ahmed E. Alprol. 2023. "Metal Oxide Nanoparticles’ Green Synthesis by Plants: Prospects in Phyto- and Bioremediation and Photocatalytic Degradation of Organic Pollutants" Processes 11, no. 12: 3356. https://doi.org/10.3390/pr11123356

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