Next Article in Journal
Critical Review on Two-Stage Anaerobic Digestion with H2 and CH4 Production from Various Wastes
Previous Article in Journal
Synergic Origin and Evolution of TDS, Mg and Fluoride in Groundwater as Relative to Chronic Kidney Disease of Unknown Etiology (CKDu) in Sri Lanka
Previous Article in Special Issue
Enhanced Photocatalytic Degradation of Tetracycline by Magnetically Separable g-C3N4-Doped Magnetite@Titanium Dioxide Heterostructured Photocatalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Use of Zero-Valent Iron Nanoparticles (nZVIs) from Environmentally Friendly Synthesis for the Removal of Dyes from Water—A Review

by
Cristina Rodríguez-Rasero
,
Vicente Montes-Jimenez
,
María F. Alexandre-Franco
*,
Carmen Fernández-González
,
Jesús Píriz-Tercero
and
Eduardo Manuel Cuerda-Correa
Departamento de Química Orgánica e Inorgánica, Instituto Universitario de Investigación del Agua Cambio Climático y Sostenibilidad (IACYS), Universidad de Extremadura (UEX), Avda. Elvas s/n, 06006 Badajoz, Spain
*
Author to whom correspondence should be addressed.
Water 2024, 16(11), 1607; https://doi.org/10.3390/w16111607
Submission received: 21 April 2024 / Revised: 9 May 2024 / Accepted: 28 May 2024 / Published: 4 June 2024

Abstract

:
This review article addresses the increasing environmental concerns posed by synthetic dyes in water, exploring innovative approaches for their removal with a focus on zero-valent iron nanoparticles (nZVIs) synthesized through environmentally friendly methods. The article begins by highlighting the persistent nature of synthetic dyes and the limitations of conventional degradation processes. The role of nanoparticles in environmental applications is then discussed, covering diverse methods for metallic nanoparticle production aligned with green chemistry principles. Various methods, including the incorporation of secondary metals, surface coating, emulsification, fixed support, encapsulation, and electrostatic stabilization, are detailed in relation to the stabilization of nZVIs. A novel aspect is introduced in the use of plant extract or biomimetic approaches for chemical reduction during nZVI synthesis. The review investigates the specific challenges posed by dye pollution in wastewater from industrial sources, particularly in the context of garment coloring. Current approaches for dye removal in aqueous environments are discussed, with an emphasis on the effectiveness of green-synthesized nZVIs. The article concludes by offering insights into future perspectives and challenges in the field. The intricate landscape of environmentally friendly nZVI synthesis has been presented, showcasing its potential as a sustainable solution for addressing dye pollution in water.

Graphical Abstract

1. Introduction

Water pollution stems from various sources, including industrial, agricultural, and domestic effluents, as well as geologic weathering, mining effluent, and atmospheric sources [1]. Water pollution, primarily from industrial and agricultural activities, poses a significant threat to water quality, rendering it unfit for consumption and harming aquatic life [2]. This widespread pollution affects drinking water from rivers, streams, lakes, and oceans globally, contributing to a high number of fatalities, particularly in developing countries [3]. Hazardous water pollutants pose severe health risks to humans and the environment. For instance, drinking polluted water can result in gastrointestinal, hepatobiliary, and neurological complications, even leading to mortality [4]. Inorganic contaminants such as arsenic, lead, and cadmium in high concentrations can cause cancer, diabetes mellitus, and cardiovascular diseases, further exacerbating the health impact [5]. The economic consequences of water pollution are significant, affecting sectors such as health, agriculture, tourism, real estate, and aquaculture/fisheries. Regions downstream of heavily polluted rivers experience reduced economic growth, reflecting underestimated costs of environmental degradation and inefficient pollution management [6]. Economic burdens associated with poor water quality include health-related expenses, high water treatment costs, and adverse impacts on economic activities like agriculture and manufacturing [7]. Regulatory measures to tackle water pollution include efforts like the European Union’s Water Framework Directive 2000/60/EC, aimed at addressing chemical pollution of surface waters. Such regulations play a crucial role in mitigating the adverse effects of water pollution and safeguarding human health and the environment.
Emerging contaminants, also known as pollutants of emerging concern, encompass a wide range of compounds found in surface water and groundwater. These pollutants originate from industrial, pharmaceutical, and municipal sources, as well as various human activities [8,9]. Examples of emerging contaminants include pharmaceuticals, personal care products, illicit drugs, nanomaterials, and endocrine-disrupting chemicals [10,11]. Due to their widespread distribution, ecological impact, and potential risks to human health and aquatic life, dyes can be considered as emerging pollutants [12]. These contaminants are synthetic or naturally occurring chemicals not commonly monitored or regulated in the environment but are suspected to have adverse health effects [13]. They pose serious threats to ecosystems, as even low concentrations can be mobile or persistent in air, water, soil, and sediments, directly or indirectly exposing aquatic species and populations to potential harm [14]. One of the significant challenges in addressing emerging contaminants is the lack of legislation to regulate their discharge or monitor these compounds, presenting obstacles for water sustainability and environmental protection [15].
Emerging pollutants, including dyes, are dumped in natural environments, posing serious health risks, and affecting life in general [16]. The application of multiple dyes and pigments in sectors derived from the textile, paper, pharmaceutical, and cosmetics industries have caused them to be present in the environment, being considered as persistent pollutants in ecosystems and, therefore, of great concern [17].
Studies indicate that during the dyeing process, a percentage of the dye used remains as a substance fixed to the fabric, but there is another quite high percentage (15–50%) that remains in the effluents of the textile industry [18]. The presence of dyes in the effluents of the industry has been corroborated. In aquatic environments, dyes represent a great threat to aquatic organisms, as they can participate in the reduction of dissolved oxygen, generating anoxic and anaerobic conditions, and impede the passage of sunlight and the diffusion of oxygen, thus interfering with the development of different forms of aquatic life [19]. Studies have revealed that dyes are the cause of illnesses and can even cause cancer [20]. In this sense, water pollution by dyes has become a great threat due to their high toxicity and their chemical stability, which makes them resistant to many degradation processes and difficult to eliminate in conventional treatment plants. This is why, over the years, studies have been carried out to evaluate the efficiency of non-conventional treatment processes for the elimination of this type of emerging pollutant [21].
Within the alternative treatments and, more specifically, considering the processes of dye removal with non-conventional materials, given their associated benefits, the implementation of metallic nanoparticles, specifically nZVI nanoparticles synthesized through green chemistry, is positioned as an alternative of great interest.
In the pursuit of well-being, modern society frequently requires resources and carries out practices for their transformation into useful products. Over time, an increase in environmental pollution has occurred due to these human activities, introducing external chemical or biological agents. This pollution has negative impacts not only on the environment but also on the organisms living within it. Consequently, the primary sources of water pollution predominantly come from human activities, including urban discharges of organic pollutants, as well as contributions from agriculture and livestock, like pesticides and fertilizers, and various industrial processes [22,23,24,25].
Urban pollution arises from residential, commercial, and service sectors, generating wastewater containing refuse and chemicals such as bleach, detergents, pharmaceuticals, and cosmetics [26,27,28,29]. These chemicals cannot be released indiscriminately into lakes or traditional waterways without proper treatment. Additionally, contaminants like metals and hydrocarbons originate from vehicle emissions. Agricultural pollution primarily results from the runoff of fertilizers [30] and pesticides [31] into water streams, often containing nitrogen, phosphorus, or sulfur, which subsequently leads to soil contamination and groundwater pollution [32]. However, the most significant and dangerous source of pollution comes from industrial activities, given the multitude of materials and substances present in industrial wastewater. These pollutants stem from various industrial processes, including manufacturing, transformation, utilization, cleaning, maintenance, and consumption, each associated with specific products and industry types. Notably, industries such as petrochemicals, textiles, energy, and paper manufacturing contribute significantly to the production of the most environmentally harmful substances. For instance, the textile industry [33], known for its extensive use of dyes and pigments [34], often releases these chemicals into wastewater without adequate treatment, resulting in substantial environmental waste.
Considering the rapid pace of industrialization, coupled with population growth and urbanization, the World Health Organization (WHO) reports that more than 2.1 billion individuals lack access to safe drinking water [35] (as per the Sustainable Development Goals by the World Health Organization). Additionally, the European Environment Agency (EEA) estimates that roughly three million locations within its member countries have been impacted by contaminants originating from local sources (as stated in the European Environment Agency’s report on Contamination from Local Sources).
In response to the mentioned pollution problem, the scientific community has directed significant attention towards exploring the potential of nanotechnology in creating nanomaterials endowed with exceptional characteristics and properties, ideally suited for remedying water pollution issues. It is worth noting that the term “nanotechnology” refers to a collection of procedures and techniques that enable the manipulation and operation of molecular structures and their constituent atoms on a nanoscale, typically spanning lengths from 1 to 100 nanometers [36]. In recent years, the studies and advancements in nanotechnology and nanoscience have rapidly translated into a wide array of products, including nanoparticles, nanowires, nanotubes, and nanolayers, that have been synthesized for application in diverse sectors of modern society [37]. Nanotechnology is not merely an area of promising research [38,39]; it has already begun to demonstrate its initial commercial applications across various industries, including electronics, automotive, sports equipment, cosmetics, and more. This expands research areas, both in advanced and emerging economies [40].
Simultaneously, nanotechnology has emerged as a captivating research subject, reflecting our fascination with the ability to manipulate the shape and composition of matter at the atomic and molecular scale through various techniques and approaches. The potential of nanotechnology is the outcome of scientific advancements throughout the 20th century, which have progressively deepened our understanding of matter’s properties at increasingly smaller scales, evolving from what can be termed “microtechnologies” to “nanotechnologies”. The remarkable development of characterization and detection techniques such as scanning and transmission electron microscopes (SEM, TEM) and atomic force microscopy (AFM) has contributed to our deliberate and active comprehension and control of matter’s properties at the atomic and molecular levels, allowing for the feasible modification of structures within this size range [41]. In addition, they can provide various types of physicochemical information on nanoparticles, but their usefulness will depend on the technology and the properties of the material, such as composition and scale.
The properties exhibited by nanomaterials [42] and their behavior differ from those of their counterpart bulks and expand when compared to the same materials outside the nanoscale. Furthermore, within this size range, all properties—physical, chemical, and biological—undergo alterations, both at the individual atom and molecule levels and as composite materials [37]. Many of these alterations depend on the behavior of electrons within the nanomaterials or how the atoms are arranged within the matter. It is important to note that in nanometer-sized materials, the movement of electrons is significantly constrained by the material’s dimensions. Additionally, the ratio of surface atoms to interior atoms is much higher than in larger materials, resulting in a heightened density of exposed sites per unit mass, and thus, might increase chemical reactivity. Notably, metallic materials like iron, silver, and gold, semiconductors such as cadmium chalcogenides, gallium arsenide, and indium phosphide, and insulators like iron oxide and titanium are among the most extensively researched materials [43].
By precisely controlling synthesis conditions and appropriate functionalization, it is possible to obtain a narrow range of sizes and shapes. Thus, the properties of nanomaterials can be tailored to some extent [44]. Consequently, nowadays there exists methodologies and technology to tune the dimension and shape of the material, and therefore, its properties, enabling the design of materials with specific characteristics [37].
In this review, the use of zero-valent iron nanoparticles (nZVIs) as an eco-friendly solution for dye removal from water is analyzed. The manuscript unfolds progressively, commencing with an examination of the role of nanoparticles in environmental applications, followed by an in-depth analysis of various synthesis strategies employed for nanoparticle production. Next, it delves into conventional and environmentally friendly methods for synthesizing zero-valent iron nanoparticles (nZVIs), highlighting their advantages and limitations. A comprehensive discussion ensues on the influence of synthetic methods on nZVI preparation, alongside strategies for stabilizing nZVIs to enhance their effectiveness in dye removal. The review also describes emerging green alternatives for nZVI synthesis, emphasizing sustainability and environmental compatibility. Furthermore, the prevalent issue of dye pollution in wastewater and current removal strategies is explored, underscoring the importance of efficient and environmentally benign approaches. The manuscript concludes with a forward-looking discussion on future perspectives and challenges in the field, aiming to provide insights for advancing sustainable water remediation practices.
This review also aims to fill a critical knowledge gap in the existing literature by consolidating and synthesizing dispersed information on the synthesis, stabilization, and application of zero-valent iron nanoparticles (nZVIs) for dye removal in water. While previous studies have explored various aspects of nZVI synthesis and its potential in water treatment, there has been a lack of comprehensive and integrated analysis that encompasses the diverse synthesis methods, stabilization techniques, and practical applications of nZVIs specifically for dye removal. By systematically reviewing and synthesizing findings from a wide range of research studies, reviews, and technical reports, this review provides a comprehensive resource that tries to bridge this gap.
Our approach involved meticulous analysis and synthesis of the existing literature, ensuring a thorough examination of each aspect of nZVI-based dye removal. By consolidating scattered knowledge and offering insights into the efficacy, limitations, and future directions of environmentally friendly nZVIs for sustainable water remediation, this review aims to inform and guide researchers, policymakers, and practitioners in the field toward more effective and environmentally benign water treatment solutions.

2. Nanoparticles in Environmental Applications

Nanotechnology continues to gain momentum as a powerful tool in addressing environmental challenges, establishing itself as one of the most significant applications in different fields, such as water depuration [44], gas cleaning [45], green synthesis [46], catalysis [43], etc. This can be attributed to the profound understanding of nanoparticle properties, enabling the development of novel synthesis methods, and enhancing environmental applications [47]. Nanoparticles find their place in environmental applications due to their remarkable capacity for removing pollutants, thanks to their high surface area-to-volume ratio. As particle size decreases, the proportion of atoms on the surface increases, leading to enhanced adsorption capacity, ease of interaction with other atoms and molecules, and improved reactivity on chemical or biological surfaces compared to bulk materials [48]. However, this heightened chemical reactivity can result in nanoparticle agglomeration, reducing their surface energy.
Zero-valent metals, exemplified by nanoscale iron (nZVI or nanozero-valent iron), possess highly reactive surfaces, and are commonly employed in the treatment of water, sediments, and soil. Noble metals, such as gold, platinum, and palladium, have been used for degradation of gas pollutants like volatile organic compounds (VOCs). Silver has a special application in the biomedical field due to its great antibacterial activity. Semiconductors like titanium, zinc, and cerium have been widely used as catalysts for water cleaning using light as a source of energy. On many occasions they are combined with other metals like copper or noble metals to increase their activity.
As a counterpoint, the nanoparticles used can reach the environment and act as a contaminant. Westerhoff et al. showed the removal of over 96% of titanium in influent sewage, with effluents typically containing <25 μg/L [49]. Batley et al. discussed the future of the bioaccumulation of nanoparticles in the environment, concluding that it is rising and potentially it will happen [50]. In recent years, some authors have reported the bioaccumulation of nanoparticles in some foods that we can find in supermarkets today. For example, Gallocchio reported how mussels exposed to TiO2 nanoparticles can adsorb them at concentrations of 209–1119 µg/kg [51]. Thus, great attention must be paid to releasing nanoparticles to the environment and, additionally, low toxicity must be a characteristic.

3. Nanoparticle Synthesis Strategies

Various techniques have been described for synthesizing nanoparticles, and the synthesis method significantly influences the size and shape of the resulting nanoparticles. Special attention is given to controlling factors such as size, shape, composition, crystallinity, and structure, as these parameters determine the intrinsic properties of metal nanostructures [52]. To achieve specific characteristics, process conditions must be meticulously controlled. These characteristics include uniform size across all particles (referred to as monodisperse samples), consistent shape or morphology, the chemical composition of both the core and the modified surface, crystalline structure, and the prevention of aggregation. To address the latter, nanoparticle syntheses typically employ stabilizing agents that attach to the surface, providing a stabilizing charge to keep the nanoparticles suspended and prevent aggregation [47].
Nanoparticle synthesis processes may seem conceptually straightforward but are challenging to execute according to the literature. Typically, authors describe two distinct approaches for the controlled synthesis of nanostructures, each with its own set of advantages and disadvantages [52]. The first approach, termed “top-down”, begins with larger materials and uses powerful techniques to break them down into particles of the desired size and is the most conventional method for nanoparticle synthesis. In contrast, the second and more recently developed approach, known as “bottom-up”, starts with individual atoms and molecules, assembling them into nanoparticles. Nowadays, some methods combine both approaches.
Top-down methods involve the subdivision of solid metals (bulk materials). This subdivision can be made by mechanical forces (ball milling, grinding, ultrasound, etc.) or chemically (degradation, consumption, etc.). A key drawback is the imperfect surface structure, along with a broad particle size distribution, typically exceeding 10 nm, and limited reproducibility. Within the “top-down” approach, various methods are utilized, with the most representative ones being thermal evaporation, high-energy milling, laser ablation, etching, etc.
Bottom-up methods focus on constructing nanomaterials atom by atom, molecule by molecule, or cluster by cluster. A common synthetic approach involves reducing metal ions to their corresponding atoms and subsequently controlling the aggregation of these atoms. Key methods within the “bottom-up” approach include the colloidal method, photochemical and radiochemical reduction, microwave irradiation, dendrimer utilization, solvothermal synthesis, and the sol–gel method [53]. Also, a bulky precursor can be fragmented into ions or molecules that are further combined to form nanoparticles. This could be called a mixed method (top–bottom method, Figure 1), and some examples are chemical vapor deposition, solvolysis, thermal evaporation, etc. Each method has inherent advantages and disadvantages in the synthesis process.
Figure 1 summarizes these processes. The main methods of preparation of nanoparticles are described below, together with wide variety of processes involving physical, chemical, or a combination of both methods in the synthesis of metal nanoparticles to develop environmentally friendly solutions for water treatment, detection of persistent pollutants, and remediation of soil and water.
Initially, any synthesis approach can be categorized as either environmentally harmful or environmentally friendly. In scientific discourse, techniques employing chemical reagents are typically deemed polluting, referred to as conventional or gray methods. Conversely, methods emphasizing the utilization of renewable resources like biomass or waste from other processes are considered environmentally friendly or green. To accurately classify them, various parameters need to be assessed (Figure 1), including CO2 levels, energy efficiency, waste generation, and overall performance. However, essential data for these evaluations, such as the yield of the synthetic process, are often lacking in the literature. Consequently, there is a pressing need for studies that rigorously assess the potential minor contaminations associated with purportedly green methods.
In the following paragraphs, the most reported methods are briefly described, giving some examples of each one.

3.1. Sol–Gel

The sol–gel process is a bottom-up nanoparticle production method. The process involves two steps, first the obtaining of a sol, which consists of dissolving the precursors, normally in an acidic solvent, obtaining a colloidal solution. This solution evolves into a network, or gel, of the aforesaid monomer. Once the gel is formed, the excess of solvent must be removed, and finally, a drying or calcination step is required [55,56,57]. The efficiency of the process is usually very high, above 70%, and with moderate energy consumption, most of which is consumed in the drying/calcination process.

3.2. Hydrothemal

The hydrothermal synthesis method is one of the most used (Figure 1). In summary, it consists of introducing the precursor salts into a suitable solvent and subjecting them to a certain temperature and pressure. The solubility and stability of the salts in the solvent help make the growth of the nanoparticles controllable. Furthermore, this technique allows the incorporation of stabilizing, emulsifying, or reducing agents. As a rule, it has a high performance, and the energy consumption is not excessive [58,59], although in many cases it must be completed with a calcination or thermal decomposition process.

3.3. Chemical Vapor Condensation of a Metal

In the chemical vapor condensation approach for nanoparticle production, the metal is evaporated on a non-wetting surface in the presence of a solvent. This results in the deposition of metal in the form of islands on the surface [60]. Subsequent condensation of the solvent covers these particles, preventing the formation of a continuous metal film. Typically, high temperatures are required. However, a modified method has been developed recently, involving the thermal decomposition of pentacarbonyl iron (Fe(CO)5) within a surfactant/solvent system [61]. This compound has a lower enthalpy of formation (−185 kcal/mol), making its decomposition easier to achieve. This approach can yield small nZVI particles (10–20 nm) at lower temperatures (140–160 °C).

3.4. Chemical Reduction Methods

This method is commonly used for reduction in solutions at atmospheric pressure. The solution of metal precursors is usually in water, and therefore, the most common precursors are inorganic salt, e.g., sulfates, nitrates, and chlorides. The chemical reduction agents also are soluble chemical reagents, such as NaHB4 or similar. The specific case of iron nanoparticles will be discussed in more detail in Section 4. However, some other reduction methods are described below.
In some cases, the reducing agent can be a gas flow, so that some authors called this method the gas-phase reduction technique. For nZVI production, the process involves reducing iron compounds like goethite (FeO(OH)) or hematite (Fe2O3) at elevated temperatures using H2 as the reducing agent [62]. Typically, nanoparticles generated through this method are larger, around 100 nanometers, and exhibit more irregular shapes with a crystalline core structure. The composition of the oxide shell is generally magnetite (Fe3O4). This synthesis process is easily scalable and has found application in the production of commercial nZVIs.

3.5. Co-Precipitation

The co-precipitation method is probably the simplest and most efficient chemical method for the synthesis of magnetic nanoparticles of the M-FeXOY-type, where M can be Co, Ni, Cu, etc. It is based on the precipitation of salt precursors of the metals forming the nanoparticles in a basic medium. The main advantage of this method is the large amount of material that can be synthesized, but the main disadvantage is the lack of control over the particle size generated, since the only control on the reaction is the kinetics. The process can be described in two stages: in the first one, rapid nucleation occurs until the concentration of the species reaches supersaturation; in the second one, the nuclei formed grow uniformly by diffusion of the species towards the surface of the particle until the final size is reached. To obtain monodisperse nanoparticles, these two stages must be separated, and nucleation must be avoided during the growth process. In general, it is difficult to maintain a homogeneous level of supersaturation throughout the reaction volume. The shape and size of the nanoparticles can be modified relatively easily by adjusting the pH, the ionic strength, the synthesis temperature, the nature of the reagents, or the proportion of Fe and metal (Co, Cu, Ni) salts [63,64].

3.6. Electrochemical Methods

The method of electrolysis synthesis involves the creation of metal nanoparticles from a solution containing Mn+/Mm+ salts by applying an electric current between the anode and cathode. The metal particles produced are deposited on the cathode, usually within a size range of between 20 and 30 nm, and are spherical and uniform shape nanoparticles [65]. This method is straightforward, cost-effective, and requires minimal time. However, a drawback is that the nanoparticles tend to aggregate and form clusters on the cathode, often necessitating the use of surfactants to disperse them.

3.7. Precision Milling

In precision milling synthesis, micro-sized metal is milled in a high-speed chamber using stainless steel balls [66]. This method has several advantages, including the absence of toxic reagents, short production times, and scalability [67]. The resulting nZVIs are characterized by small size (~20 nm) and a large surface area (~40 m2/g) [68]. However, the nanoparticles exhibit irregular shapes due to deformation during milling and have a propensity to agglomerate.
Methods ranging from hydrothermal and solvothermal synthesis, pyrolysis, ultrasonication, sol–gel methods, precipitation and co-precipitation, chemical reduction methods using chemical agents such as NaBH4, citrates, ascorbic acid, sodium hydroxide, hydrogen, among others (see Table 1), are described in the literature. These strategies aim to explore new material properties at both the atomic and molecular level, but they often involve complex processes, form precipitates that are mostly amorphous, have high energy requirements with difficulties controlling the growth rate of metal nanoparticles, use hazardous solvents and expensive reagents, and pose a potential biohazard. Due to these drawbacks, alternative techniques and sources for the synthesis of nanoparticles are being promoted more decisively such as the use of plant extracts as reactive agents for the synthesis of nanoparticles from green chemistry.
Table 1. Use of various metals in nanoparticle and nanocomposite synthesis and their subsequent use in environmental pollutant removal.
Table 1. Use of various metals in nanoparticle and nanocomposite synthesis and their subsequent use in environmental pollutant removal.
NanoparticlesSupportSynthesis TechniquesEnvironmental
Applications
Reference
Removal of Dyes
ZnO-Sol–gel co-precipitation in KOHMethylene blue[55]
Zr-doped Fe2O3CoOxHydrothermalOrange II[59]
Ca peroxideStarchPrecipitationMethylene blue[63]
TiO2g-C3N4(g-CN)
nanolayers
Co-precipitation;
thermal polymerization
Methylene blue[64]
AuAu/VO2/CeO2PhotodepositionMethylene blue[69]
Magnetite (MN)Fe3O4@CMicrowave-inducedMethylene blue[70]
Ag Na0.5Bi0.5TiO3
nanospheres
Chemical solution,
hydrothermal
Rhodamine B [71]
HRC-Ag/AgarAgarose hydrogelReduction with KBH4Methylene blue,
rhodamine B
[72]
Cationic
polyethyleneimine (PBVPR218)
--Dyes[73]
EC-AgElettaria
cardamomum
Reduction with NaBH4Methylene blue,
rhodamine B
[74]
nZVIJanus particlesReduction with NaBH4Methyl orange[75]
CdAl2O4FeCo-precipitationBrilliant blue,
brilliant green
[76]
CuNiFe2O4g-C3N4Gel auto combustionMethylene blue[77]
Black phosphorus (BPQDs)-IO TiO2-Sonication assisted;
liquid exfoliation
Rhodamine B,
methylene orange
[78]
g-C3N4-Fe3O4@KFKapok fiberOne stepRhodamine B[79]
CeO2-Co-precipitation,
calcination
Methylene blue[80]
PLA/CMC/GO
_f-COOH@Ag
Polymeric matrixReduction with
ascorbic acid
Methylene blue[81]
(B-GNP)GraphenePolyvinyl alcohol filmAtrazine[82]
Magnetite nanoparticles (MNs)β-cyclodextrin (β-CD) and quaternary
ammonium salts
Co-precipitationMethylene blue,
Orange G
[83]
C3N5-LDH-AgC3N5-LDHPrecipitation in NaOHTartrazine[84]
Removal of Organic Compounds
Ferrihydrite
(Fh)
Non-ionic surfactant Brij L4Co-precipitationCooking oils[44]
TiO2-Sol–gelPhenol[56]
Zr-doped Fe2O3CoOxHydrothermalBisphenol A[59]
AuNPsAu/VO2/CeO2PhotodepositionAromatic alcohols,
p-nitrophenol
[69]
nZVIJanus particlesReduction with NaBH4Trichloroethylene[75]
Fe3O4/FeSBiocharPollen pyrolysisphenol[85]
AA@Fe°Amino acidsReduction with KBH4Tributyl phosphate,
n-dodecane
[86]
Biochar-PyrolysisBisphenol A[87]
Fe3O4rGOCo-precipitation4-Aminophenol[88]
Aptamer-MrGO@Au and ssDNA-AuNP@MBsBPA aptamerReduction with
trisodium citrate
Bisphenol A[89]
ZnO@SCF SppCobalt FerriteSpinel cobalt ferritePrecipitationPhenantrene[90]
Fe⁰@C/surfactantCarbonHydrothermal;
carbothermic;
reduction with NaBH4
Nitrobenzene[91]
Zn/FeOSurfactant foamsReduction with NaBH4Diesel[92]
Removal of
Pharmaceutical
Compounds
CuO and nZVI-Sol–gel;
precipitation in NaOH;
reduction with NaBH4
Levofloxacin[57]
CuO-Co-precipitation;
calcination
Metronidazol
[93]
Ag-Cu-LiMultimetal
nanorods
Precipitation in NaOH
Antibacterial activity[94]
WS2-Hydrothermal
thiourea
Tetracycline[95]
ZnOAlginate nanofibersHydrothermalTetracycline[96]
BiFeO3-CombustionOfloxacin[97]
TGC/NiCr2O4Tubular g-C3N4 (TGC)CalcinationTetracycline[98]
ZnOCarbon cloth (CC)HydrothermalHydroxychloroquine[99]
Ag/Ag3PO4-VAgAg nanocluster with vacancies of Ag (Ag/Ag3PO4-Vag)Reduction with NaBH4Sulfamethoxazole[100]
Au-Reduction with NaBH4/320 °CPenicillin G,
sulfamethazine,
tetracycline,
enrofloxacin,
skewered fish
[101]
Bi-Ag nanoalloys-SolvothermalBacterial infections[102]
TiO2NW-CuO-
cellulose
CelluloseSolvothermalEscherichia coli[103]
Removal of
Pesticides
Ag (SNP)-Reduction with NaBH4Miticidal activity[104]
AuOvalbuminReduction with
trisodium citrate
Carbaryl[105]
Au@ZIF-67ZIF-67HydrothermalThiram,
Carbendazim
[106]
Chitosan (CS)-Ionic gelationPlant diseases[107]
Cu (ChNC@Cu)Chitin nanocrystals (ChNC)TEMPO oxidationPesticides[108]
SiO2BiopolymersSol–gelPlastic films[109]
ZnO and TiO2-One-potPolycyclic aromatic hydrocarbons[110]
Pd-nZVI and
S-nZVI
-Reduction with NaBH4Trichloroethylene[111]
Toxicity abatement
PbO2Humic acidChlorinationToxicity in medaka
fish
[112]
Metal, metal oxide, carbon, plastic- Phytotoxicity[113]
nFe@Fe3O4, nFe3O4 nFe2O3 -Toxic effects on
zebra fish
[114]
ZnO and Ag-Sol–gel+
reduction with NaBH4
Acute toxicity in zebra fish[115]
Chemical processes
Pt@S-1β zeolite-Synthesis of naphtha[116]
ZnOMolasses and ureaPrecipitationSynthesis of fertilizer[117]
OMGR Oximagnesite/green rust-HydrothermalPhosphate recovery[118]
CrMnFeCoNi-HEO; MgCoNiCuZn-HEOSpinel-structure sHEOSol–gelDesalination techniques of seawater[119]
RuNi-rGO@
MNCs
N-doped C nanosheetsReduction with H2Hydrogen evolution reaction (HER)[120]
Others
TiO2Recycled rubber tilesSol–gelAirborne pollution[121]

3.8. Commercial Nanoparticles

Commercial nanoparticles (CNPs) are also widely used in various fields such as medicine, pharmaceuticals, and consumer products. The main advantage of using CNPs lies in the fact that they offer well-established synthesis methods and scalability for mass production, making them readily available for industrial applications. CNPs are extensively characterized using various spectroscopic techniques like TEM, SEM, FTIR, and UV–visible absorption spectroscopy, ensuring consistent quality and properties. A wide variety of uses of CNPs have been detailed in the literature and are illustrated in Table 2.
Table 2. Commercial nanoparticles for environmental applications.
Table 2. Commercial nanoparticles for environmental applications.
NanoparticlesSupportEnvironmental ApplicationsReference
SiO2@CA MSsMicrospheresRemoval of dyes[122]
UHMWPE/TiO2PolyethyleneRemoval of methyl orange, methylene blue, Congo red, and tetracycline[123]
nZVI, Fe2O3, Fe3O4-Florfenicol containing cow manure[124]
ZnO-Escherichia coli[125]
nZVI-Tetracyclines[126]
ZnO and TiO2-Tetranychus urticae and Neoseiulus californicus[127]
NiO-Neurotoxicity
in zebra fish
[128]
ZnOFilm of chitosanLoaf bread shelf life[129]
nZVI-Polycyclic aromatic hydrocarbons[130]
ZnO-Gut microbiome alterations in rats[131]
Ag-Improved quality of Capsicum annum crops[132]
SiO2, Al2O3, and
O-CNT
-Clofibric acid, acetaminophen,
sulfamerazine
[133]
nZVI-Hydrophobic
organic
compounds,
sediments
[134]
ZnO-Biofortification of
Dracocephalum moldavica
[135]
ZnOPannonibacter-
phragmitetus
Inhibition of microbial Cr(VI) reduction[136]
Se-As (III), soybean roots[137]
CNPs exhibit consistent sizes and morphologies, which surpass those synthesized in the laboratory, resulting in less variability in their chemical properties. Furthermore, they exhibit superior adherence to various supports like microspheres and films, as outlined also in the table, facilitating the attainment of more comparable results.
CNPs are produced using traditional chemical and physical methods, which may involve the use of hazardous compounds and harsh reaction conditions. In contrast, green-synthesized nanoparticles are characterized by being less toxic, cost-effective, and eco-friendly, as they utilize biogenic materials such as plants, algae, waste biomass, and microorganisms as reducing agents. Nonetheless, despite the benefits of green synthesis, the commercial production of green-synthesized nanoparticles has not scaled up, highlighting limitations in the scaling of green synthesis for commercial production. In other words, commercial nanoparticles offer established synthesis methods, scalability, and extensive characterization. The differences in production processes highlight the challenges in scaling-up green synthesis for commercial production.

3.9. Emerging Green Alternatives

The use of plant extracts in the green synthesis of nanoparticles has demonstrated significant advantages over conventional synthesis methods. The synthesis of nanoparticles by chemical reduction from plant extracts represents a sustainable and environmentally friendly approach with numerous environmental applications, including water treatment, for the removal of a wide variety of pollutants, such as dyes, pharmaceutical compounds, pesticides, inorganic compounds, and heavy metals. The synthesis of green nanoparticles takes advantage of the reducing and stabilizing properties of phytochemicals present in various plant extracts to convert metal ions into nanoparticles. As shown in Table 3, a comprehensive review of this method and its environmental applications has been performed. Plant extracts rich in bioactive compounds such as flavonoids, phenolics, terpenoids, and alkaloids serve as reducing and stabilizing agents for metal salts, usually salts of silver, gold, copper, iron, magnesium, titanium, or zinc, dissolved in solution. The plant extract is added to the metal salt solution, initiating the reduction process. The phytochemicals in the extract reduce the metal ions to form nanoparticles. The nanoparticles are stabilized by the phytochemicals in the extract, which prevents their aggregation and ensures their colloidal stability.
Table 3. Synthesis of nanoparticles by chemical reduction using plant extracts (green chemistry) for environmental applications.
Table 3. Synthesis of nanoparticles by chemical reduction using plant extracts (green chemistry) for environmental applications.
NanoparticlesSupportReactive AgentEnvironmental ApplicationsReference
Removal of Dyes
Ag/Ti-Aloe vera L.e.Rhodamine B[46]
MoO3 and WO3
nanorods
-Leidenfrost (H2O)
with NaOH
Methylene blue[138]
Ag-Antidesma acidum L.e.Congo red,
methylene blue
[139]
B-doped g-C3N4/TiO2g-C3N4Spinacia oleraceaMethylene blue[140]
Au-Wedelia urticifoliaRhodamine B[141]
ZnS/Fe3O4Carboxymethyl
cellulose
Co-precipitation of Fe(II)
and Fe(III)
Methylene blue,
methyl orange
Congo red,
and rhodamine B
[142]
ZnO-Citrus x lemonReactive green 19[143]
IRCFA-PDA@AgIron-rich coal fly ash (IRFA)-Polydopamine (PDA)FloralMethylene blue[144]
ZnO, CuO,
MnO2 and MgO
-Leucaena leucocephalaGolden yellow-145,
Direct red-31
[145]
nZVI-Cow and goat milkMethyl orange[146]
Au-Cell-free filtrate
of Penicillium rubens
Methylene blue,
phenol red,
bromothymol blue,
methyl orange
[147]
ZnOMesoporous
carbon
Firecracker wasteMethyl orange[148]
ZnO-Aloe veraMalachite green,
Basic violet 3
[149]
ZnO and SiO2-Cyperus alternifoliusMethylene blue[150]
Organic Compounds
Pt-SnO2rGO-CHAmaranthus spinosusMethanol[151]
nZVI-Tung,
aspen, and holly L.e.
Tetrabromobisphenol A[152]
CeO2-Dillenia indica2,2-difphenyl-1-
-picrylhydrazyl
[153]
GM-Ag-Gnetum montanum3-nitrophenols and
4-nitrophenols
[154]
MgO-Jatropha oilCompounds in air, HC,
CO, CO2
[155]
Fe-C, Co-C, and Ni-C-Tea residueCystine[156]
Removal of
Microorganisms
ZnO Brassica oleraceaGram-negative Bc
Escherichia coli
[157]
Ag Pseudomonas canadensis bacterial isolatePseudomonas tolaasii Pt18[158]
CuO Alpinia officinarumColletotrichum
gloeosporioides
[159]
nZVI Common mantlePathogenic fungi,
wheat plants
[160]
Removal of
Pharmaceutical
Compounds
ZnO-Aloe veraAmoxicillin[149]
Ag-Passiflora foetidaAntibacterial[161]
Ag-Calotropis proceraFungal and
bacterial pathogens
[162]
Ag, Cu, and Fe Catharanthus roseus L.e.Several anti-inflammatory
drugs
[163]
LigninNanocellulose
cryogels
Kraft ligninDiclofenac,
metropolol,
tramadol,
carbamazepine
[164]
Removal of
Inorganic Compounds
MgO-Jatropha oilCO, CO2 in air[155]
Fe-C, Co-C, and Ni-C-Tea residueCrO42¯[156]
Chitosan-Salix subserrata
bark extract
As in rats[165]
Magnetite-Amla (tree bark)U(VI)[166]
ZnO-Acacia catechu L.e.As[167]
Note: L.e.: leaf extract.
There are many environmental applications, such as water treatment. Silver nanoparticles are used for the removal of dyes, pharmaceutical compounds, and pesticides and can also be used to disinfect water because of their strong antimicrobial properties. Zinc and magnesium oxide nanoparticles synthesized using green chemistry methods can be used to remove heavy metals and organic pollutants from contaminated soil and water. Gold nanoparticles can catalyze the degradation of organic pollutants such as dyes (rhodamine B, methylene blue, phenol red, bromothymol blue, methyl orange) in wastewater through different processes, contributing to the remediation of contaminated water bodies. Furthermore, nanoparticles obtained by green chemistry methods can be used for the elimination of conventional pesticides, reducing the environmental impact of agricultural activities. They can also be incorporated into air filtration systems to capture and remove pollutants and particulates, helping to improve indoor and outdoor air quality.
The advantages of green synthesis using plant extracts are that it reduces reliance on hazardous chemicals and minimizes the generation of toxic by-products, making it environmentally sustainable. Green synthesis methods are often cost-effective compared to traditional chemical synthesis routes, as they use readily available plant materials and simple reaction conditions.
In conclusion, the literature review underscores the viability of applying nanotechnology to combat emerging contaminants, notwithstanding inherent challenges. Through the synthesis of nanoparticles utilizing diverse metals, a spectrum of methodologies, ranging from traditional techniques to eco-friendly approaches and readily available commercial routes, has emerged as a potent strategy in environmental remediation efforts. Traditional synthetic procedures have long been employed, taking advantage of the inherent properties of metals to engineer nanoparticles for specific applications. Green chemistry principles have further enriched this landscape, advocating for environmentally benign methodologies that minimize waste and energy consumption. Concurrently, the accessibility of commercial nanoparticles has streamlined research endeavors, offering standardized materials with consistent properties that enhance reproducibility and comparability of results.
Beyond these established methods, recent innovations continue to expand the arsenal of nanotechnology in environmental stewardship. Emerging approaches such as functionalized nanoparticles and nanocomposite materials exhibit promising capabilities for targeted contaminant removal, paving the way for more efficient and selective remediation strategies. Furthermore, the interdisciplinary nature of nanotechnology facilitates holistic solutions to complex environmental challenges by integrating principles from materials science, chemistry, and environmental engineering.

4. Zero-Valent Iron Nanoparticles: Conventional and Eco-Friendly Synthesis

4.1. Zero-Valent Iron Nanoparticles

Given the need to develop sustainable methods and solve environmental problems, iron is used as a biocompatible material with a relatively non-toxic nature and high catalytic efficiency among the transition elements. It has been widely used in the reduction of environmental pollutants in an environmentally friendly way [168,169,170], in biomedical applications [171,172,173], in batteries [174,175,176], as a catalyst in a large variety of processes [177,178,179], etc. Iron has the advantage of being suitable for use in sustainable technologies as it avoids the toxic load generated by other materials. It normally substitutes Ni or Co in a reduction catalyst [180,181], Pd in hydrogenation reactions [182,183], and metal oxides like TiO2 or CuO in pollutant degradation [184,185,186,187]. Iron has two oxidation states, Fe2+ and Fe3+, which give rise to oxide minerals such as hematite (Fe2O3), magnetite (Fe3O4), goethite (FeO(OH)), and siderite (FeCO3); in water, the predominant species are Fe2+ and Fe3+ and ferrous and ferric organic complexes. In the absence of air and at pH ≈ 2, the inorganic Fe3+ salts are the most stable species. Under normal conditions the most abundant species is Fe2+, while at basic pH the insoluble hydroxide predominates [188]. Reduction of Fe ions into Fe(0) obviously requires the presence of Fe(II) or Fe(III) in solution. Fe(III) starts to precipitate at pH above 7.5, and Fe(III) at pH above 3–3.5, thus giving rise to the presence of solid iron hydroxides and oxohydroxides. Hence, when attempting to prepare Fe(0) nanoparticles from dissolved iron species, it is of the utmost importance to operate within the pH interval at which Fe(II) and/or Fe(III) exist in solution. This can be easily understood from the Pourbaix diagrams of iron that can be found in Ref. [189]. The Pourbaix diagram indicates that the stability of iron in solution is related to both the electrode potential and the pH of the aqueous solution. It predicts the direction of spontaneous iron reduction, analyses the trend in metal oxidation from thermodynamics and estimates the composition of the reduction products [190]. The strong reducing character of the metal core of nZVI nanoparticles acts synergistically in the removal of pollutants from groundwater, soil, and wastewater [189,191].
Because of their exceptional characteristics, effectiveness in removing contaminants, low toxicity, and ability to move effectively through porous media, nZVIs can serve as cost-effective alternatives to in situ contaminant removal methods. These nZVIs contain donor electrons, which are responsible for their high reactivity in aqueous environments. This reactivity is a result of both iron (Fe0) corrosion reactions and non-selective reactions with dissolved oxygen. In the particular case of dyes, understanding factors influencing the oxidation process, such as nZVI modifications, concentrations, pH, temperature, and soil constituents, is crucial for optimizing treatment efficiency [192]. Dyes are primarily removed by destroying their chromogenic groups through reduction or Fenton-like reactions with zero-valent iron nanoparticles [193]. Nonetheless, the presence of active oxygen species in solution also plays a key role in the degradation process. For instance, persulfate activation, facilitated by nZVIs, yields sulfate radicals with potent oxidizing power, offering high selectivity in pollutant degradation [194,195,196]. Additionally, aerobic degradation pathways, enabled by nZVIs, involve the reductive activation of molecular oxygen, generating reactive oxygen species for the oxidation or mineralization of organic pollutants [197]. The core–shell structure-dependent pathways (see below) and the molecular oxygen activation mechanism provide valuable insights into aerobic degradation processes mediated by nZVIs.
It is commonly accepted that nZVI nanoparticles exhibit a core–shell structure, where a metallic iron core is enveloped by a thin layer of iron oxides or oxyhydroxides [198,199]. The thickness of the passivation layer, which varies depending on synthesis conditions, significantly impacts the electron transfer process; thicker passivation layers hinder effective electron transfer from the metallic core to the outer surface [200]. In environmental applications, the core–shell nZVI model satisfactorily explains the efficacy in wastewater treatment and catalytic processes, as the outer shell prevents iron nanoparticle aggregation while offering protective coverage [201]. Characterization techniques such as spherical-aberration-corrected scanning transmission electron microscopy (Cs-STEM) and energy-dispersive X-ray spectroscopy (EDS) provide confirmation of the core–shell structure and enable analysis of the passivation layer composition [199]. The core’s structural attributes hold significant importance for the chemistry of nZVIs. While the core serves as the source of electrons for reactions involving nZVIs, the shell is the focal point for intricate chemical reactions and electrostatic interactions [202,203,204,205]. The shell oxide is flawed and its disordered nature renders it more reactive when compared to a straightforward passive oxide.
The typical size of Fe nanoparticles falls in the range of ten to several hundred nanometers, depending on the technique used. The surface area of the nanoparticle depends on the particle size, being theoretically between 150 and 15 m2/g for the usual range (10 to 100 nm; see Figure 2a). Most of the articles that have reported their surface area obtained a range of 30 to 50 m2/g for nanoparticles of 20–40 nm.
In cases where iron nanoparticles aggregate, they feature a continuous oxide shell, with the metal cores separated by a thinner interfacial oxide layer. Nemati et al. synthesized Fe nanoparticles by thermal decomposition of organometallic compounds [206]. A transmission electron microscopy (TEM) image reveals a typical cluster of nanoscale zero-valent iron (nZVI) particles. The core shell ranging from 3 nm to 7 nm is visible. Montferand et al. synthesized iron nanoparticles by reducing a salt precursor in liquid media at temperatures above 200 °C using stabilizing and surfactant agents [207], resulting in several shapes depending on the stabilizing agent and concentration.
Ansari et al. [208] prepared Fe nanoparticles using ferric iron and sodium borohydride as the reducing agents under ambient conditions. These nanoparticles visibly clustered together, forming interconnected chain-like structures, primarily due to magnetic and electrostatic interactions. Each particle consisted of a dense core surrounded by a thin shell, which exhibited noticeably lower contrast than the interior core. When examining the core region through selected area electron diffraction (SAED), diffused ring patterns are observed, characteristic of nanocrystalline body-centered cubic (bcc) iron metal. Notably, the chains of interconnected nanoparticles exhibit a continuous oxide shell, but each metallic core is separated from its neighbors by a thinner interfacial oxide layer, approximately 1 nanometer in thickness. The observed shell thickness under TEM aligns reasonably well with measurements from previous X-ray photoelectron spectroscopy (XPS) studies. These earlier investigations estimated an average oxide thickness of approximately 2–3 nanometers based on the relative intensities of oxidized and metallic iron signals. A phase-contrast TEM image shell further illustrated that the oxide displays a speckled contrast and lacks periodic lattice fringes, indicating an amorphous character in the oxide layer’s structure. This disordered oxide layer can be partly attributed to the exceedingly small radio of the nanoparticles and the curvature of the oxide shell, both of which introduce significant strain, hindering the formation of crystalline structures. Additionally, the presence of a small amount of boron in the form of iron borides [209] in the oxide film, stemming from the borohydride precursor used during the synthesis step, may contribute to defective sites and alter the oxide structure [203].

4.2. Influence of Synthetic Methods for the Preparation of nZVIs

The selected synthetic method might allow rational control of the size and shape of the nanoparticles to tailor as much as possible the properties of the formed nanomaterials to a specific application [193]. The most common are illustrated in Figure 3, with the typical particle size range and the yield obtained.
The most used methods for the synthesis of iron nanoparticles are electrochemical, chemical reduction using chemical agents or plant extracts (sometimes call biomimetic), hydrothermal, and sol–gel. The range of particle sizes is quite wide within each technique and varies depending on the preparation conditions. However, to finely control the range some techniques are better, especially the ones that allow the use of additives, such as chemical reduction, hydrothermal, sol–gel, or microemulsion. Techniques such as chemical vapor deposition lead to smaller particle sizes, but the size range is very narrow, which can be an advantage or disadvantage depending on the application. The yield is not commonly reported, and it is one of the most influential factors in being able to declare the technique as “green”. The reported yields in Figure 3 are extracted from several publications and are very approximate. Reporting the yield of nZVIs obtained should be encouraged. The methods with the largest yield are chemical reduction, hydrothermal, sol–gel, microemulsion, and ball milling. Regarding morphology, the most common shape is spheres; however, it varies with the size and the technique. Normally, small nanoparticles tend to be spherical, while large nanoparticles can be rods, spheres, or cubes. The introduction of additives or other metals during the synthesis helps in tuning the morphology.
Chemical reduction methods require a metal precursor, a reducing agent, and sometimes the use of a stabilizing agent. Regarding the metal precursors, the most common are inorganic salts; nitrates, chlorides, or sulfates. The reducing agent can be divided into two large groups: common chemical reagents and plant extracts. In the former case, dissolved iron salts are reduced, using NaBH4 or KBH4 as a reducing agent, effectively converting Fe (II) or Fe (III) salts into Fe nanoparticles [210], typically at room temperature and with a reaction time of one hour or less. Next, iron nanoparticles must be separated and washed thoroughly to remove all impurities of the reducing agent. Normally, this methodology leads to a narrow range of particle size distribution and the method is quite reproducible. However, it is worth noting that NaBH4 is highly toxic, corrosive, and flammable, posing environmental risks. Moreover, nanoparticles synthesized with NaBH4 tend to agglomerate easily, leading to reduced reactivity and stability. During reduction with NaBH4, it is oxidized to various species, such as B2O3, B(OH)3, and some intermediates. The reuse of these species could go a long way towards the sustainability of the method, giving a very reproducible and sustainable method. There are currently projects that consider boron as a possible agent for the storage of H2 and boron oxides are already being used to obtain NaBH4. The Kotai Hydrogen Project is a collaborative project with experts from Curtin University in Perth. The research project is investigating the feasibility of using sodium borohydride (NaBH4) as a safe hydrogen ‘carrier’ that can be used on demand where needed.
In chemical reduction using plant extracts, dissolved iron salts are reduced using plant extracts as natural reducing agents, which usually is claimed as green synthesis by the scientific community. However, this is not supported by evidence such as life cycle assessment studies, calculation of CO2 emissions, etc. This point will be discussed in detail in Section 5.

4.3. Stabilization of nZVIs

Fe(0) nanoparticles tend to become unstable under atmospheric conditions and readily form oxides and hydroxides, including Fe2O3, Fe3O4, and FeO(OH), under specific conditions. In addition, like other nanomaterials, they have a naturally strong inclination to agglomerate into larger particles due to van der Waals interactions and subsequently settle, as graphically shown in Figure 4. In efforts to prevent the aggregation of nZVIs and enhance their environmental remediation capabilities, various modifications to its structure and synthesis methods have been developed. These modifications aim to boost the stability, reactivity, and surface area of nZVIs [211]. Figure 4 also schematizes such procedures. Techniques for modifying nZVIs include (among others):
  • Incorporating a second metal into nZVIs.
  • Coating the surface of nZVIs with organic substances.
  • Synthesizing emulsified nZVIs.
  • Fixed support.
  • Encapsulation.
  • Electrostatic stabilization.
  • Steric stabilization.
Figure 4. Scheme of different stabilization methods.
Figure 4. Scheme of different stabilization methods.
Water 16 01607 g004

4.3.1. Incorporating a Second Metal into nZVIs

To enhance the reaction rate of nZVIs in pollutant degradation, small amounts of transition metals like Pt, Ag, Cu, and Ni can be added to the nanoparticle surface, serving as catalysts [43]. These metals are believed to accelerate degradation through mechanisms such as facilitating electron transfer via galvanic coupling and generating reactive atomic hydrogen. Additionally, bimetallic particles reduce the accumulation of corrosion products on the particle surface and protect against passivation. Consequently, bimetallic nanoparticles find applications in the decolorization of aromatic compounds and polychlorinated biphenyls, which typically degrade slowly. Bimetallic nanoparticles can be synthesized through two methods [212]. The first involves mixing the as-synthesized nZVIs with a solution containing the second metal salt to be added. As the reduction potential of transition metals (Pt, Ag, Cu, or Ni) is more positive than that of Fe [213], a metal replacement reaction occurs, and the zero-valent metal is deposited on the nZVI surface. This is known as the solution deposition method. The second method, called the co-reduction method, entails the simultaneous reduction of iron and the second metal salt using sodium borohydride [214]. Regardless of the method used, the transition metal forms a thin, discontinuous layer on the nZVI surface.
Among the studied transition metals, Pd has proven to be the most effective due to its optimal structure and chemical properties for generating active hydrogen species and breaking carbon–halogen bonds in dehalogenation reactions [215]. For instance, Fe/Pd nanoparticles have demonstrated reaction rates over one order of magnitude higher than common nZVIs for trichloroethylene degradation. However, a key drawback of bimetallic systems is their acceleration of iron corrosion via galvanic effects, which can rapidly decrease system reactivity. Further research is required to enhance the properties of bimetallic systems and extend their reactivity over prolonged periods.

4.3.2. Coating of nZVI Surfaces

Various compounds, including synthetic and natural polymers, surfactants, carboxylic acids, and polysaccharides [216], have been employed to coat the surfaces of nZVIs. Generally, molecules with high molecular weights and a high density of functional groups are more effective stabilizers [217]. The size of the resulting nanoparticles can be controlled by adjusting the concentration and chain length of the surfactant. Notably, biopolymers have garnered significant attention due to their cost-effectiveness and environmentally friendly nature. Examples include food-grade polysaccharides like chitosan [218,219], carboxymethyl cellulose (CMC) [218], starch [63,220], and guar gum [221], which have demonstrated their effectiveness as coatings for nZVIs in multiple studies. Among these, CMC and polyacrylic acid (PAA) have emerged as particularly promising stabilizers. CMC has proven to enhance nZVI reactivity and in situ derivability for various applications, while PAA/nZVI has shown improved mobility compared to CMC/nZVI under specific conditions, such as carbonate porous media and concentrated Ca2+ groundwater [222].

4.3.3. Emulsified nZVIs

Nanoparticle emulsification represents a cutting-edge technique that combines the properties of nanoparticles with the versatility of emulsions, opening a range of possibilities in diverse fields such as pharmaceuticals, cosmetics, food, and materials science, as well as contaminant removal. This process involves the dispersion of nanoscale particles within a continuous phase to form a stable emulsion. Several preparation methods are used to emulsify nanoparticles, including high-energy methods such as ultrasonication, high-pressure homogenization, and microfluidics. These techniques help break down larger nanoparticle aggregates into smaller, uniform particles dispersed throughout the emulsion. Stabilizing agents such as surfactants, polymers, and proteins are often used to prevent agglomeration or settling of the nanoparticles within the emulsion. These agents form a protective layer around the nanoparticles, preventing them from coalescing and maintaining the stability of the emulsion.
Recent studies indicate that the type of emulsion used in nZVIs is a water/oil emulsion, which offers the advantage of protecting the surface, preventing aggregation, and improving the degradation efficiency of non-polar substances. Two types of emulsions are presented: (i) oil-in-water, where nZVIs are placed in a non-polar substance that is dispersed in an aqueous solution [223]; and (ii) water-in-oil, nZVIs contained in water droplets of 10–20 µm in size are surrounded by an oil film [224]. Different surfactants are used to ensure the stability of the emulsions.

4.3.4. Fixed Support

The literature has also reported the utilization of Fe(0) nanoparticles supported on diverse materials such as clays, activated carbon, sawdust, and natural and synthetic zeolites to improve their effectiveness in removing contaminants from aqueous environments and prevent nanoparticle aggregation. Hydrophobic supports also extend the reactivity of nanoparticles and enable the adsorption of hydrophobic contaminants like PCBs onto the nZVI surface. Materials such as phyllosilicate minerals [225], granitic residual [226], silica [227], activated carbon [228], zeolites [229], biochar [230] or polymer membranes have been utilized as supports for nZVIs. Immobilization of nZVIs on these supports can be achieved through carboxyl, hydroxyl, or amine groups serving as chelating sites. Depending on the carrier and its characteristics, the adsorption properties of nZVIs can be improved. Numerous studies have demonstrated that immobilizing nZVI particles on solid supports simplifies operation while preserving the excellent reducing capacity of nZVIs [231]. However, a notable drawback of this modification method is that nanoparticles are exposed on the support surface, making them susceptible to rapid oxidation by air. Encapsulation of nZVIs is thus preferred, to enhance stability without compromising their contaminant remediation efficiency.

4.3.5. Encapsulation of nZVIs

Another method for modifying nZVIs involves their capture or encapsulation within a matrix, such as chitosan [232], arabic gum, calcium carbonate [233], calcium alginate, or carbon materials [91]. Among these, encapsulation within carbon nanospheres or microspheres has shown significant promise. Such encapsulation results in unique core–shell nanomaterials that exhibit both the adsorption properties of carbon and the reducing agent properties of nZVIs. Methods for synthesizing these hybrid materials include the confined plasma arc method and hydrothermal carbonization.

4.3.6. Electrostatic Stabilization

Electrostatic stabilization consists of repulsion between nanoparticles by forces related to charges or surface charge densities. These forces appear in nanoparticles, especially in those that have an interface between two crystallinity states, have doping agents, impurities of the precursor salts, etc. This effect can also be enhanced if components are adsorbed on the surface of the charged nanoparticles [234].

4.3.7. Steric Stabilization

Steric stabilization of nanoparticles is a technique used to prevent their aggregation and favor their dispersion in a solvent or medium. This technique involves the use of long-chain polymers or surfactants adsorbed on the surface of nanoparticles to create a steric barrier that hinders interactions between nanoparticles, thus improving their stability in dispersion. The advantage is that they result in stable dispersions that can be stored for extended periods without significant changes in the size or properties of the nanoparticles. Polyethylene glycol (PEG) or polyvinyl alcohol (PVA) are long-chain polymers that have a flexible structure that extends away from the nanoparticle surface, creating a physical barrier between the nanoparticles [235]. Surfactants, such as the Tween or Triton series, can also provide steric stabilization. The surfactant molecules adsorb on the surface of the nanoparticles, forming a layer that prevents the particles from coming into close contact with each other. Steric stabilization of the nanoparticles is usually achieved during their synthesis or post-synthesis.
Table 4 summarizes the most widely used conventional methods for nZVI synthesis and stabilization.
Table 4. Conventional synthesis of zero-valent iron nanoparticles.
Table 4. Conventional synthesis of zero-valent iron nanoparticles.
FeNPsPrecursorSupportSize/nmMorphologyCharacterizationReference
MCMZVIFeCl3·6H2O+
glucose+CO(NH2)2
Mesoporous carbon20–100Uneven sizeXRD, SEM, XPS, FTIR, BET[194]
CS-nZVI
(core–shell)
FeCl3·6H2O+NaBH4-15.42–97.57SphericalUV–vis, FTIR, TEM, SEM EDX, XRF[198]
nZVIFeCl3·6H2O+NaBH4-34–110SphericalUV, XRD, SEM, EDX, TEM, DSL[208]
nZVI@FSGFeSO4·7H2O+NaBH4Flaxseed
gum
73–87SphericalDLS, FESEM, EDX, FTIR, DR5000[216]
CMC-S-nZVIFeSO4·7H2O+NaBH4Carboxymethyl cellulose (CMC)90SphericalTEM-EDS,
UV–vis, PSD,
ζ-potential
[218]
nZVI/SS/BCFeSO4·7H2O+NaBH4Biochar with stable starch45.7–37-SEM, EDS,
BET, FTIR, XRD
[220]
nZVIFeSO4·7H2O+KBH4Guar gum87.4CubicTEM, XRD, SEM, XPS[221]
nZVIFeSO4·7H2O+KBH4Attapulgite--FTIR, SEM-EDS, XPS[225]
Fe@SiO2FeSO4·7H2O+KBH4Silica40–50NanospheresXRS, FTIR, TEM, EDS[227]
nZVI/ACFeSO4·7H2O+NaBH4Activated
carbon
40–100RoughSEM, XRD, BET
XPS
[228]
CS@nZVIFeSO4·7H2O+NaBH4Chitosan13.12-XRD, FESEM, EDS, FTIR[232]
CaCO3-nZVIFeSO4·7H2O+NaBH4CaCO375–89SphericalSEM-EDX, XRD, TEM, FTIR, XPS, BET[233]
nZVI-A and nZVI-SFeCl3·6H2O+NaBH4Rhamnolipids60 and 42-SEM, XRD, FTIR, TG, DLS,
ζ-potential
[236]
FG-nZVIFeSO4·7H2O+NaBH4Flaxseed gum extract<100SphericalDLS, FESEM, RDX, FTIR[237]
nZVIFeCl3·6H2O+NaBH4-20–60Uniform morphologyXRD, SEM, TEM[238]
nZVIFeSO4·7H2O+NaBH4-14.3-XRD, SEM, TEM, FTIR[239]
nZVI/GOFe (NO3)3·9H2O+NaBH4Graphene
oxide
10–20DispersedSEM, TEM-EDS, FTIR[240]
nZVI-kaol/PESFeCl3·6H2O+NaBH4Kaolin and
poly-ethersulfone (PES)
42Ridge and valleyXRD, FESEM,
FTIR
[241]
NC-nZVIFeCl3·6H2O+NaBH4Nanocelluloses (NC)116–200SphericalSEM, TEM-EDX,
FTIR, XRD, XPS
[242]
Fe@BCBlack liquor lignin
and Fenton sludge
in one step
Biochar20–50Notable
aggregation
XRD, FTIR, BET, BJH, FESEM, TEM[243]
MFO@nZVIFeCl3·6H2O+NaBH4MnFe2O4 (MFO)
hydrogel
-Mass of spheroidal particlesXRD, SEM, FTIR, UV–vis DRS[244]
CDLA@nZVI and
CDCA@nZVI
FeCl3·6H2O+NaBH4β-cyclodextrin (CD): CDLA and CDCA25 and 30AmorphousNMR, FTIR, HRTEM, DLS,
ζ-potential, FESEM, EDAX, VSM, XRD,
XPS, TGA
[245]
Fe3O4@nZVI-PEIFe3O4+FeSO4·7H2O+NaBH4Poly
ethylenimine
--SEM, TEM, FTIR, XRD, XPS[246]
S-nZVIFeSO4·7H2O+KBH4 -[247]
nZVI-LBCFeCl3·6H2O+NaBH4Biochar--FTIR, XRD, TEM, XPS, VSM, BET,
ζ-potential
[248]
nZVI/GACFeSO4·7H2O+NaBH4Granular
activated
carbon (GAC)
40SphericalSEM, BET, XRD[249]
nZVI/GOFeCl3·6H2O+NaBH4Graphene oxide (GO)4.97-SEM-EDS, XRD, FTIR[250]
nZVI/Sch-AP and nZVI/
Sch-CO
FeSO4·7H2O+NaBH4Schwermannite (Sch)50SphericalSEM, XRD, BET, XPS[251]
A400-nZVIFeSO4·7H2O+NaBH4Polystyrenic gel
(Purolite A400)
75–150-FTIR, SEM, EDAX, XRD, TGA[252]
nZVI@
Zr(OH)4
FeCl3·6H2O+NaBH4Zirconium
hydroxide
--TEM-EDS, XRD, BET, FTIR, XPS[253]
pyGA-nZVIFeCl3·6H2O+NaBH4Pyrogallic acid (pyGA)40–90SphericalSEM, TEM-EDS, BET,
ζ-potential, XRD, FTIR, XPS
[254]
Ox-nZVIFeCl3·6H2O+NaBH4Oxalate30–40SphericalBET, SEM-EDS, FTIR, XRD, XPS[255]
nZVI 1 and nZVI 2FeSO4·7H2O+NaBH4
FeCl3·6H2O+NaBH4
-72 and 38SphericalSEM, TEM-EDS, XRD, BET, XPS, Raman
spectroscopy
[256]
nZVI/
copper slag
FeCl3·6H2O+NaBH4Copper slag30-FE-SEM, EDX, XRD, FTIR, BET, VSM, ζ-potential[257]
WPANF/
nZVI
FeCl3·6H2O+NaBH4Polyacrylonitrile fiber (WPANF)15–50-XRD, FTIR, BET, XPS, SEM-EDS, TEM[258]
nZVI, ds-coated nZVI, and ds-FeSFeSO4·7H2O+NaBH4-60.12 and 110SphericalSEM, XRD, FTIR, ζ-potential, DLS[259]
nZVISelf-combustion of Fe2O3 and NaBH4-2–8AmorphousXRD, HRTEM, EDS[260]
nZVILaser fragmentation in
liquids (LFL) ethylene
glycol and polyethylene glycol 400
-10.5 and below 3-DLS, LDE, TEM, XPS[261]
CS@BC/
S-nZVI
FeSO4·7H2O+NaBH4Chitosan and
Biochar
--SEM, BET, FTIR, XRD, XPS[262]
3D-RGO@nZVI/
Al2O3
FeSO4·7H2O+NaBH4Reduced
graphene oxide
<100SphericalSEM,
BET,
Raman
spectroscopy,
XRD, XPS
[263]
nZVI/n-ligninFeCl3·6H2O+NaBH4Lignin--TEM, XPS, XRD[264]
BP-S-nZVIFeCl3+NaBH4+sulfate-
reducing bacteria (SRB)
---FESEM, TEM, XRD,
BET
[265]
nZVI-DEFeCl2·4H2O+NaBH4Diatomaceous earth (DE)20–40SphericalXRD, SEM, EDX, TEM, BET[266]
CnZVIFeSO4·7H2O+NaBH4-80–99SphericalUV, FTIR, XRD, TEM[267]
G-nZVI-BC and
C-nZVI-BC
FeSO4·7H2O+NaBH4Biochar--SEM, XRD, FTIR, XRF[268]
Iron oxideFeCl3·6H2O+FeCl2·4H2O+
NaOH
-10 ± 4Regular
crystalline
TEM[269]
CS/nZVIFeCl3+NaBH4Chitosan25SphericalSEM, FTIR, XRD, EDX, VSM, BET, TGA, DSC[270]
BC@nFe-CAFeSO4·7H2O+NaBH4Biochar--SEM, EDS, FTIR, Raman spectroscopy, XRD, XPS[271]
PDA@Fe/rGOFeSO4·7H2O+NaBH4Reduced graphene oxide (rGO) and
polydopamine (PDA)
51-TEM, XRD, FTIR, XPS, VSM[272]
PDCA@nZVIFeCl3+NaBH42,6-pyridinedicarboxylic acid, (PDCA)115NanospheresSEM, EDS, EDX, XRD, FTIR[273]
nZVIFeSO4·7H2O+NaBH4--SphericalSEM, TEM, BET[274]
nZVI/BCPyrolysis (Fe2O3+biochar)-200-XRD, SEM[275]
CMC-nZVI,
bare-nZVI,
PAA-nZVI,
PSM-nZVI,
PVP-nZVI
FeCl3+NaBH4CMC, PAA, PSM, PVP9.53, 65.4, 106.4, 106.6, and 109SphericalSEM-EDX, XRD, FTIR, TEM[276]
nZVI-HPBFeCl3·6H2O+NaBH4Hydrophilic
biochar
52–243-SEM, TEM, XRD, FTIR, XPS[277]
nZVIFeSO4·7H2O+NaBH4-20–60SphericalBET, SEM EDX, BET, XRPD, XPS[278]
nZVI/SBA-15FeCl3·6H2O+NaBH4Mesoporous silica (Santa
Bárbara-15)
50–80SphericalSEM EDS, TEM, BET[279]
nZVIFeCl3·6H2O+NaBH4-36Regular and irregularXRD, SEM, EDX, UV–vis[280]
nZVI-chitosanFeCl3·6H2O+NaBH4Chitosan15–20-SEM, XRD, TEM, FTIR[281]
nZVIBulk iron disks+
solvents+
laser
-9.4 and 3.5SphericalTEM EDS, XPS[282]
nZVI/RSFeSO4·7H2O+NaBH4Biochar--FTIR, XRD, Raman spectra, BET[283]
AHG@nZVIFeSO4·7H2O+KBH4Aluminum
hydroxide gel
-Irregular and roughSEM, FTIR, XPS, XRD[284]
S-nZVI/GAFeSO4·7H2O+NaBH4Graphene
aerogel (GA)
-Flake-like shellTEM, SEM, BET, XRD, XPS, FTIR, Raman spectrum[285]
nZVI and SnZVIFeCl2·4H2O+
FeSO4·7H2O+NaBH4
--Cross-linked sphericalTEM EDX, XPS, XRD, XAS[286]
Fe0@p-SiO2FeCl3·6H2O+KBH4SiO230–40-TEM, XRD, XPS,
ζ-potential
[287]
Fe/TRGOCarbothermal GO+
Fe(NO3)3·9H2O
Graphene oxide--HRTEM, XRD, XPS, Fe Mössbauer spectroscopy[288]
nZVIFeCl2·4H2O+NaBH4--Chain-likeFTIR, SEM, TEM, XRD, XPS[289]
Fe@CQDs MNCsFeCl3+NaBH4--SmoothFTIR, XRD, SEM, TEM[290]
nZVIFeCl3·6H2O+NaBH4--Irregular and noncircularXRD, TEM EDS, FESEM[291]
Nanostructured nZVIFeSO4·7H2O+NaBH4--RoughXRD, SEM, TEM, DLS[292]

5. Environmentally Friendly Methods for the Synthesis of nZVI

As an eco-friendly alternative to the use of chemical reductants in nanoparticle synthesis, the utilization of natural materials, such as plants or plant extracts, has emerged as one of the most effective strategies [293]. The so-called green synthesis method, or biomimetics, makes use of natural extracts from environmentally friendly substances, such as agricultural or food industry waste materials (fruit residues, vines, or eucalyptus leaves) as well as natural plant extracts typically used for human consumption, like coffee, tea, or grape must, water hyacinth, barley, lemon balm (Melissa officinalis), parsley (Crispum spp.), and sorghum bran (Sorghum spp.), known for their rich polyphenolic content. Very rarely is the concentration of polyphenols reported, and therefore, it is not possible to establish a relationship between their concentrations and the particle size, yield, or morphology of nanoparticles. It should be noted that there are some limitations with certain plant sources; for example, tea can be relatively expensive, while barley, although chemically stable, may not always be available for harvesting. Even so, these extracts, which are rich in phenolic compounds (polyphenols), reducing sugars, ascorbic acid, flavonoids, and amino acids, play a crucial role in the generation of iron nanoparticles. On the one hand, like borohydride, they make the reduction of iron ions possible; on the other hand, and this is a noticeable advantage compared to NaBH4, they contribute to the formation of stable iron complexes, rather than borohydrides.
The advantages of this green chemical reduction method commonly reported are:
  • Enhanced removal capacity and longevity: The presence of polyphenols in these extracts enhances their removal capabilities because bioactive compounds such as polyphenols have several benzene groups substituted by hydroxyl functional groups in their structure, so they can reduce metal ions to their elemental state, obtaining nanoparticles. They are also low cost or economically viable, reducing the consumption of organic compounds and the generation of toxic products.
  • Environmentally friendly: This method is hailed as a potential environmentally friendly process, characterized by lower toxicity levels, the absence of aggressive reagents, and the generation of non-harmful by-products.
  • Cost-effective: Many authors claim it might provide a low-cost alternative for nanoparticle synthesis.
  • Inherent stabilization: The extract matrix often acts as a stabilizer, reducing nanoparticle agglomeration without requiring the addition of dispersants.
  • Biomass valorization: The method offers potential for biomass valorization, further contributing to its sustainability.
Despite these advantages, widespread acceptance of green synthesis remains limited due to a lack of comprehensive understanding regarding the reactivity, physicochemical properties, and agglomeration tendencies of the resulting nanoparticles. Studies suggest that the choice of plant extract used leads to variations in nanoparticle size and specific surface area (refer to Table 5). Additionally, the reduction of iron to nZVIs by plant extracts can sometimes result in the formation of different iron compounds, including iron oxides or hydroxides, throughout the process.
The process of green synthesis typically involves the extraction of biomolecules from plants through heating in water, usually near the boiling point. Subsequently, the extract is separated from the residual plant material and combined with a solution containing Fe2+ or Fe3+. However, a key limitation of this method lies in the inability of wet plant extracts to efficiently reduce Fe2+/Fe3+ to nZVIs, potentially leading to the formation of iron (hydr)oxides rather than metallic iron [294,295]. The typical pH for extraction is around 5, which may affect the reduction potential. Additionally, the characteristics of the resulting nZVIs, including size and surface area, can vary significantly depending on the specific plant extract utilized.
Considering that polyphenols are potent antioxidants capable of trapping free radicals, their reaction with iron salts, such as FeCl3, results in the formation of ortho-diphenol compounds. These compounds exhibit a green coloration, while those with three adjacent hydroxyl groups appear dark gray. The presence of isolated phenolic hydroxyls is indicated by a yellow color. This color change occurs due to chloride ions attacking the hydrogen in the hydroxyl group, leading to bond breakage and the formation of a complex with iron (complex formation). A transition from the initial yellow to dark yellow signifies the presence of phenol and the reduction of Fe3+ to Fe0. Phenols interact with the empty orbitals of FeCl3, while the unreacted polyphenols continue to reduce the iron complex, forming an aromatic organic complex and further reducing Fe, resulting in the release of Fe nanoparticles in the nZVI state [296].
The nanoparticles are synthesized by the formation of phenoxy groups, dimerized by hydrogen bonds which bind to the Fe0. If the pH is increased (OH concentration is increased), more phenoxy groups will be generated, which can act as dye reductants and adsorbents. If the pH of the zero-charge point is acidic, this means that most of its surface is negatively charged, so better cation bonds will be formed.
It has been reported that the existence of hydroxyl and carbonyl groups in the structure of biomolecules makes them powerful reductants, giving rise to unstable zero-valent iron nanoparticles. However, if they have double bonds in their structure, these biomolecules act as a coating agent, which translates into an increase in the chemical stability of the iron nanoparticles [297].
Table 5. Synthesis of iron nanoparticles through chemical reduction using plant extract.
Table 5. Synthesis of iron nanoparticles through chemical reduction using plant extract.
FeNPsPrecursorSupportSize/nmMorphologyCharacterizationReference
BC-nZVI-BCOak+FeCl3Biochar68–521Twister and
serpentine
SEM, EDS, XRD, DLS[230]
nZVI,
ds-coated-nZVI+ds-FeS
Phoenix dactylifera
+FeSO4·7H2O
- SphericalSEM, XRD, FTIR,
ζ-potential,
DLS
[259]
GnZVIAmaranthus dubius
leaf extract+
FeSO4·7H2O
-1–3SphericalUV, FTIR,
XRD, TEM
[267]
G-nZVI-BC and C-nZVI-BCGreen tea residues+FeSO4·7H2O---SEM,
XRD,
FTIR, XRF
[268]
Iron oxideCymbopogon citratus
extract+
FeCl3·6H2O+Na2CO3
-9 ± 4Regular crystallineTEM[269]
nZVIAzadirachta indica (neem)
Mentha longifolia (mint) L.e.+FeCl3·6H2O
- SphericalSEM, TEM, BET[274]
GT-nZVIBlack tea+FeCl3·6H2O-80Regular and
irregular
XRD, SEM, EDX,
UV–vis
[280]
Fe-NP-GVMansoa alliacea
+FeSO4·7H2O
-18.22SphericalXRD, UV–vis,
AAS FTIR, TGA
[295]
nZVIShirazi thyme L.e.+FeSO4·7H2O/
Pistachio green hulls
pomegranate/banana/
mango+FeCl3
black tea+
FeCl3·6H2O
-40–70
114,
76,
95
--[298]
nZVIGreen tea+
Fe (NO3)3·9H2O
-5–45Amorphous sphericalTEM, SEM/EDS, XRD, BET[299]
AC/nZVIPomegranate
peel extract+FeCl2
Activated
carbon
CrystallineFTIR, XRD, BET, FESEM[300]
nZVICleistocalyx
operculatus L.s+FeCl3
-100SphericalSEM, XRD,
FTIR
[301]
nZVI
@gBC
Carbothermal
(sawdust+FeCl3·6H2O)
Graphene---[302]
nZVICleistocalyx operculatus L.s+FeCl3-100SphericalSEM,
XRD,
FTIR
[303]
TP-nZVI/PETea polyphenolsPolyethylene-RoughSEM, TEM,
ζ-potential,
XRD, FTIR, XPS
[304]
Fe@CRice powder+
Fe (NO3)3·9H2O
Biomass-
derived carbon
30–150IrregularXRD, FTIR, SEM, TEM, EDS[305]
nZVI@GNPsCleistocalyx
operculatus
Leaf extract+
graphene NPs
Graphene
nanoplatelets (GNPs)
30–100SphericalSEM, XRD, EDS, FTIR[306]
G-nZVIRipe mango peels+FeCl3·6H2O---XRD, FTIR, TEM, BET, SEM-EDX[307]
EGnZVIEucalyptus grandis+FeSO4·7H2O-50–500SphericalXRD, FTIR,
Raman,
SEM, TEM/EDS
[308]
G-nZVI/BGreen tea+
FeSO4·7H2O
Calcined
bentonite
8 30IrregularBET, SEM, TEM, FTIR, XRD, XPS[309]
nZVICorn+
FeCl3·6H2O
-150–300SphericalSEM, TEM-EDS, XRD, XPS[310]
EL-nZVIEucalyptus L.e.+FeSO4·7H2O-87SphericalSEM, FTIR[311]
SnZVI@HPACGreen tea+FeSO4·7H2O---SEM-EDS, TGA, FTIR, XRD[312]
SC-nZVIGreen tea waste
+FeCl3·6H2O
Silty clay--FTIR, SEM, XRD, BET, ζ-potential[313]
nZVI/CFCarbothermal
(cotton fiber
+Fe (NO3)3·9H2O)
Cotton
carbon fiber
--XRD, SEM,
BET
[314]
nZVI
@TP-Mont
Green tea+FeSO4·7H2OMontmorillonite15–30SphericalTEM, XRD, FTIR, SBET, XPS,
ζ-potential
[315]
RCL-nZVIRicinus communis L.e.+FeCl3·6H2O-4.84
25.6
IrregularSEM, TEM, FTIR, EDS, XRD, XPS,
ζ-potential
[316]
B-BT-nZVIBlack tea+ FeCl3Bentonite<50-AFM, SEM,
ζ-potential, BET
[317]
GT-nZVI@VCGreen tea+
Vit C+
FeSO4·7H2O
-100-XRD,
TEM,
SEM, FTIR
[318]
nZVIPomegranate
peel extract+
FeCl3·6H2O
-40–60SphericalUV–visible,
FTIR,
SEM
[319]
k-nZVIRuellia tuberosa+FeCl3·6H2OKaolin20–40SphericalXRD, SEM,
TEM, EDS
[320]
R-FeNPsGreen tea+FeCl3·6H2OResin20–40SphericalSEM, TEM, EDS[321]
nZVI
@Fe3O4
@HMIMPF6
Camellia
sinensis+
FeCl3·6H2O+
FeCl2·4H2O
Magnetite+
1-hexyl-3-methylimidazolium
hexafluorophosphate
30SphericalFTIR, XRD,
VSM, BET,
SEM, TEM
[322]
FeNPsJFJackfruit peel (JFP)
extract+FeCl2
-33Irregular sphericalFTIR, TEM, XRD, SEM, EDX[323]
G-nZVIPomegranate fruit peel+FeCl3·6H2O-60–75Spherical to cubicalUV–vis, XRD, TEM, SEM, DLS,
ζ-potential
[324]
nZVI
coupling with MR-1
Green tea+
FeSO4·7H2O
Shewanella
oneidensis
MR-1
--SEM-EDS, XPS, FTIR, Raman,
EEM
[325]
RC-nZVIRicinus communis seed
extract+Fe3+
-20SphericalSEM, TEM, FTIR, XRS, EDS, XRD, XPS, ζ-potential[326]
FeNPsDenitrifying bacteria+
+FeCl3
---UV–vis, XPS,
FTIR,
TEM
[327]
AC/nZVIPomegranate peel extract+FeSO4·7H2OActivated
carbon
--XRD, FTIR, FESEM[328]
Zeolite/
nZVI
Pomegranate peel extract+FeCl2·4H2OZeolite30-FTIR, FESEM,
BET,
XRF
[329]
BGT-nZVIGreen tea+FeCl3·6H2OBentonite-- [330]
PPAC-nZVIPomegranate peel extracts+FeCl2·4H2OActivated
carbon
19–24-FESEM, BET, FTIR[331]
nZVI-RBCRice husk+FeSO4·7H2O Rice husk-derived
biochar (RBC)
Rice husk-derived biochar (RBC)100-SEM EDS, XRD, FTIR, BET, XPS[332]
S-nZVI/ACUlva. prolifera+
FeCl2·4H2O
Algal carbon-Flower-likeSEM EDS, TEM, XRD, BET, XPS[333]
Fe/N-OBCarbothermal
(hematite oak
wood biochar)
-- XRD, XPS, SEM EDS, FTIR, BET[334]
GT-nZVIBlack tea+FeCl3·6H2O-83IrregularXRD, SEM, EDAX[335]
n-ZVI-NPsMentha piperita+FeCl3-5–10SphericalUV–vis, SEM-EDX, DLS[336]
Fe0+
Fe1.91C0.09+
Fe3O4
Neurospora crassa
+urea+FeSO4
-50-SEM EDX, XRD, XPS, BET[337]
TP-ZVI-OBGreen tea+FeCl3·6H2OOak wood
biochar
--BET, FTIR, SEM, EDS, XPS, XRD[338]
F–Fe0 adsFicus sycomorus dry L.e.+FeCl3·6H2OWheat bran (B), rice bran (RB), activated charcoal (Ach), and bentonite
(Bent)
2.46–11.49CircularUV–vis, HRTEM[339]
NA-FeNPsNephrolepis auriculata+ FeCl3-40–70SpheroidalTEM, XRD, EDS,
XPS, FTIR
[340]
GMP-nZVIMango peel extract+FeCl3·6H2O-1–10-UV–vis, EDX, XRD, XPS, FTIR[341]
DOX@GTCs-FeNPsGreen tea catechin
powder+FeCl3·6H2O
159 TEM, UV–vis, AAS,[342]
Micro/FeNPsMango, rose, Neem L.e.
carom seeds, and clove buds+FeCl2·4H2O
Polyvinyl alcohol (PVP)75–6500Spherical or irregularSEM, XRD, EDX, FTIR, UV–vis[343]
Iron oxideEucalyptus L.e.
+FeCl3·6H2O
Cetyltrimethylammonium bromide (CTA)80–90SphericalXRD, EDS, FTIR, TGA[344]
FeNPsEichhornia crassipes L.s
+FeSO4·7H2O
--RodUV–vis, SEM, TEMXRD, FTIR[345]
Iron oxideCocos nucifera+FeCl3-10–100ClusteredUV–vis,
TEM, XRD, XPS
[346]
FeNPsEucalyptus globulus,
Mangifera indica,
Syzygium cumini,
Psidium guajava+FeCl3·6H2O
-38–47IrregularUV–vis, FTIR, FESEM EDS, XRD[347]
VIOak L.e.+FeCl3·6H2O-20–100IrregularTEM, EDS, XRD[348]
Fe3O4NPsCoriandrum sativum L.e.
+FeCl3
-20–90SphericalUV–vis, FTIR, XRD SEM EDX[349]
Iron oxideLantana camara fruit+FeSO4·7H2O+
FeCl3·6H2O
-28SphericalFTIR, TGA, PSA, SEM EDAX,
ζ-potential
[350]
Iron oxideLantana camara L.e.+FeSO4·7H2O-10–20NanorodsXRD, FTIR, SEMEDX, UV–vis[351]
LGFeNPsEucalyptus L.e.+
laterite
-20–70SphericalFESEM EDX, XRD, FTIR, BET[352]
FeNPsEucalyptus L.e.+FeSO4·7H2O-70 ± 20SphericalSEM EDS, FTIRE, XRD, TEM, XPS, XRD, BET[353]
Fe2O3@SiO2Zanthoxylum rhetsa
+FeCl3·6H2O
SiO212.2 ± 0.8Cluster-likeFTIR, XRD, SEM EDX, HRTEM[354]
SJA-FeNPsSyzgium jambos
+FeCl3
-13.7 ± 5SphericalUV–vis, TEM, XRD, XPS[355]
nZVIVaccinium corymbosum
+FeCl3·6H2O
-52.4IrregularTEM, SEM, BET, XRD[356]
Fe3O4@ZnOAzadirachta
indica(neem)+
FeSO4·7H2O+
Fe (NO3)3·9H2O
ZnO38Brick-likeXRD, FTIR, SEM EDX, TEM, TGA[357]
Ec-Fe-NPsEichhornia crassipes+FeCl3-20–80AmorphousSEM, EDS, TEM, XPS, FTIR, DLS,
ζ-potential
[358]
FeNPSMoringa oleifera+FeCl3-2.6–6.2 and 3.4–7.4SphericalUV–vis, XRD, FTIR, TEM[359]
Iron oxideBlack tea+FeSO4·7H2O-5–50AmorphousXRD, FTIR, SEM, TEM, EDS[360]
nZVIGreen tea L.e.+FeCl3-116-FTIR, SEM[361]
FeNPsRosa damascene (RD),
Thymus vulgaris (TV),
and Urtica dioica (UD)
+FeCl3·4H2O
-100NonuniformFTIR, SEM, TEM, XRD[362]
FeNPsEuphorbia cochinchensis
+FeCl3
-100SphericalGMS, TEM, XPS, XRD, BET[363]
FeNPsEucalyptus L.e.
+FeSO4·7H2O
-20–80PolydisperseSEM, XRD, XPS, FTIR[364]
Iron oxideSapindus mukorossi
+Fe(NO3)3·9H2O)+FeCl3
-<50NanorodsXRD, FESEM, TEM[365]
Iron oxide nanorods (IONRs)Mangifera indica L.e.
+FeSO4·7H2O
3.0 ± 0.2NanorodsFESEM, EDX,
XRD, TEM
[366]
FeNPSMangifera indica,
Murraya koenigii,
Azadiracta indica,
Magnolia champaca
+FeSO4·7H2O
AI
(96–110), MC
(99–129), MIAND MK
(100–150)
SphericalUV–vis, SEM-EDS-FTIR[367]
GT-Fe NPs and EL-Fe NPsEucalyptus L.e.+green tea L.e.+FeSO4·7H2O 20–80Quasi-sphericalSEM, XRD, FTIR[368]
nZVIsWaste from citrus juice (orange, lime, lemon
and mandarin)
+FeSO4·7H2O
3–300Spherical,
cylindrical,
irregular
TEM, XRD,
Mössbauer
spectroscopy
[369]
nZVIBlack tea,
grape mark vine L.e.
+FeCl3·6H2O
15–45-TEM[370]
(ZVI) NPsTerminalia chebula+FeSO4·7H2O <80AmorphousTEM, XRD,
UV–vis, FTIR
[371]
nZVI26 tree leaf extracts+FeCl3 10–20SphericalTEM[372]
Note: L.e.: Leaf extract.

6. Dye Pollution in Wastewater and Current Removal Strategies

Numerous organic compounds are introduced into the environment as waste pro-ducts stemming from diverse industries, including textiles, paper, leather, cosmetics, and more recently, photovoltaics, batteries, and light-emitting diodes. Of these industries, textiles take the lead as one of the most significant contributors to global water pollution [373]. The textile sector consumes millions of tons of water daily in its production processes, with a substantial portion of this water being discharged into rivers and lakes, frequently without prior treatment to eliminate pollutants that detrimentally impact aquatic ecosystems. Synthetic dyes constitute a substantial portion of the toxic compounds present in textile industry wastewater. This predicament is exacerbated by the fact that approximately 2% of the dyes manufactured within this industry are directly released into natural effluents [374].
Generally, dyes are recognized as potential sources of carcinogenic, mutagenic, and teratogenic substances, capable of infiltrating the food chain. Hence, it becomes imperative to exert stringent control over the industrial utilization of dyes to prevent their discharge in concentrations that exceed the environment’s assimilation capacity.
Dyes exhibit structural commonalities across different types, categorized based on their chemical structures as well as industrial applications [375]. Polymethine dyes, for instance, feature a conjugated polymethine chain with terminal functional groups and an odd number of π-centers and π-electrons [376,377]. Various functional groups, such as anthraquinone, azo, phthalocyanine, sulfur, indigo, nitro, and nitroso, also determine their structural classification [378]. The chemical structures of all the dyes referred to in this work, together with their common and IUPAC names can be found as Supplementary Material.
Chemical stability is a notable trait of dyes, rendering them resistant to light, temperature, and environmental degradation. This persistence poses ecological concerns, particularly for freshwater aquatic ecosystems [379]. Synthetic dyes, in particular, exhibit high resistivity towards washing, heat, light, and biological agents, exacerbating their environmental persistence [380]. Dyes’ adverse effects on aquatic life are well documented, including inhibition of photosynthesis, reduction of dissolved oxygen levels, and disruption of behavior and reproductive patterns in aquatic organisms [381]. Toxicity testing, measuring parameters like LC50 and EC50, confirms their detrimental impact on aquatic ecosystems, directly affecting photoautotrophic organisms and causing ecological imbalances [381,382]. In many instances, however, the pollution caused by dyes employed in sectors like the textile industry goes largely unnoticed. This lack of awareness among both workers and the exposed population makes it a silent and unaddressed issue. Consequently, the release of effluents containing these dyes into the environment is deemed unacceptable without prior treatment to bring their concentrations within permissible levels. The paramount challenge revolves around reducing the dye concentration in industrial effluents cost-effectively.
However, these types of wastewater-containing dyes are very difficult to treat, because the organic molecules that compose them have complex aromatic groups in their structure, which makes them resistant to light and temperature; these characteristics make them bio-accumulative and difficult to degrade in living beings, which causes illness or can even be fatal. In addition, the colloidal matter together with these colloids can cause diseases or even death. Also, the greasy foam present in the water increases the turbidity and gives the water a bad appearance. Evidently, the human being is not free of the polluting action of the colorants spilled in the environment. If the spillage occurs in crop fields, it can affect soil productivity and, through filtration or runoff, it can reach aquifers or surface waters (rivers, lakes, reservoirs) with the consequent problems in the water supply for human consumption.
Today, the manufacturing and utilization of synthetic dyes for fabric dyeing have evolved into a colossal industry. The allure of vibrant colors on textiles has a history dating back to at least 3500 B.C. However, it was in 1856 when Perkins unveiled the realm of synthetic dyes, ushering in a wide spectrum of vivid hues. Yet, the environmentally detrimental aspects of these dyes have raised profound concerns.
Dyes are water-soluble compounds that possess the ability to impart color to fibers while remaining unaffected by factors such as light, temperature, and soap. There are several ways to classify the colorants, according to their chemical structure, their characteristics, and their applications [383].
Treatment technologies for dyes differ only slightly from those generally used for other pollutants. Such differences are mainly due to some specific characteristics of dyes, such as, for instance, low biodegradability, which makes them persistent and refractory to many conventional treatments [384,385]. The typical methods for dye wastewater treatment include physical, chemical, and biological approaches, as well as advanced oxidation processes and bio-adsorption techniques [386,387,388,389,390]. Physicochemical methods and adsorption using activated carbon are commonly employed for dye abatement [391,392]. Bio-adsorption, particularly using bio-adsorbents derived from local sources, is favored for its long-term viability [393]. Bioremediation, including microbial treatment, enzyme-mediated dye removal, and phytoremediation, has emerged as a sustainable approach for dye degradation [389]. Furthermore, nanotechnology and advanced oxidation processes (AOPs) hold promise for effective dye removal [394], as will be detailed in the following paragraphs.
In some cases, conventional treatment methods for the removal of dyes in wastewater are not effective due to the recalcitrant nature of the pollutants. Such methods include adsorption, degradation, and mineralization.
Regarding adsorption and degradation methods, in many cases the treatment involves both. Adsorption using biomass as the source for creating the adsorbent has received special attention. Degradation methods focus on oxidation process of dyes. Some advanced oxidation processes involving the use of iron cations in solution have been used for the degradation of dyes. Specifically, these are the processes known as Fenton (if the cation used is Fe2+) or Fenton-like (if Fe3+ is used), in which the dissolved iron catalyzes the decomposition of hydrogen peroxide with the consequent formation of two highly oxidizing hydroxyl radicals (·OH), which result in the decomposition of the dye, the latter being reduced (ideally) to innocuous substances such as CO2 and water. Its major drawback is the large amount of peroxide that must be added if such a method is used alone [395]. nZVIs are currently being used in conjunction with advanced oxidation processes such as Fenton or Fenton-like processes for dye removal. The aim is to achieve a synergistic effect by increasing the efficiency of the nanoparticles and decreasing the peroxide concentrations in the oxidative processes used. The resulting combination shows promising effects, and its development is in full growth, especially focusing on the development of green chemistry methods for the synthesis of nZVIs.
Metallic iron nanoparticles have emerged as an exceedingly efficient alternative for mitigating this problem within the textile industry. Fe nanoparticles may act as adsorbent, degradation catalyst, and flocculant or additive for flocculation. As mentioned before, this work focuses mainly on degradation. Figure 5 shows no clear trend in dye removal efficiency as a function of either particle size, precursor, or dye. This huge variability probably is due to the multiple factors that condition the removal activity, such as experimental conditions, nanoparticle stabilization and accessibility, oxidant agent, etc.
Table 6 summarizes studies involving the synthesis of iron nanoparticles using plant extracts for use in dye removal. Parameters influencing the conditions include the nanoparticles synthesized, the different precursors, the plant extracts used, the concentration of the dyes, the yield, and the reaction time for dye removal to take place.
The plant extract, whose natural composition includes polyphenols, sugars, alkaloids, phenolic acids, proteins, and coenzymes, plays a triple role, acting as a reducing, coating, and stabilizing agent. The result is the rapid initiation of the reduction process and the generation of large quantities of stable nanoparticles. This multifunctionality expedites the onset of the reduction process and yields substantial quantities of enduring nanoparticles. These nanoparticles possess an extensive surface area and demonstrate thermal and electrical conductivity. Moreover, they have the capacity to acquire magnetic properties, facilitating convenient extraction from the environment and potential reuse [396]. When the concentration of polyphenols in the plant is high, the size of the nanoparticles decreases and they are highly reactive, thus the degradation rate increases and there are no iron oxides present in the nanoparticles.
According to the literature, the synthesis of nanoparticles is influenced not only by the plant extract, but also by various plant parts such as stems, leaves, bark, and fruits. The composition and percentage of biomolecules in these plant parts significantly influence the morphology of the nanoparticles and the properties of their metal precursors as they can act as stabilizing agents [397]. Biomolecules containing -OH and C=C groups serve as reductants for iron salts, transforming them into elemental iron while undergoing oxidation to COOH. It has been shown that if these biomolecules also have conjugated double bonds, they act as capping agents. This capping process can lead to steric hindrance around the nanoparticles, preventing their aggregation. Other large biomolecules such as cellulose, hemicellulose, proteins, etc., seem to play the role of stabilization, preventing agglomeration and favoring flocculation.
Grape leaf extracts have been analyzed, where biomolecules such as phytols, terpenoids (sitosterols), and antioxidants (vitamin E) play the dual role of reducing agent and coating [398,399]. The nanoparticles generated are quasi-spherical with a diameter of 20–60 nm. Other biomolecules such as hydroquinone, benzenediol, 2-methoxy-4-vinylphenol, and some polyphenolic compounds are also responsible for the reduction and coating of iron salts. In eucalyptus leaf extract, benothein B also forms complexes with iron and acts as a reducing and stabilizing agent [296,308].
As seen above, a promising strategy to achieve better adsorption/reduction performance together with increased stabilization of the nanoparticle is the use of zero-valent iron nanoparticles immobilized on porous supports. The use of ecological supports such as biochar or activated carbon, among others, is being sought to a greater extent. The advantages of using these types of carriers are the easy availability of biomass, as indicated in Table 6 (cellulose, corn stover, dried distillers’ grain, red oak, bamboo charcoal, and pine sawdust), and the synergistic effect produced in the adsorption mechanism on the surface of the activated carbon or biochar and the nanoparticle. The most advantageous carriers used are those derived from biomass, biochars, or pyrolyzed material. Another advantage is the high surface area and easy availability of biomass. It should be noted that if nanoparticles are supported on resins, polymers, and heavy metals they may contribute to toxic effects or give rise to other harmful pollutants.
When nanoparticles are synthesized in situ for use in decontamination in water or soil, the use of biopolymers (guar gum, chitosan, starch, xanthan gum, riboflavin, cellulose, and agar-agar) that have hydroxy groups on their surface prevents the nanoparticles from agglomerating due to possible repulsion effects, or to the increased nucleation speed of the iron atoms that gives rise to a thin negatively charged layer, so that electrostatic stabilization occurs.
Chitosan, which is a glucosamine biopolymer derived from crustaceans, modifies the surface of the nanoparticles significantly as it has -NH2 and -OH groups that provide bin-ding groups, resulting in greater stabilization of the nanoparticle. The addition of amino acids has also been used as a step after the synthesis of the nanoparticles. Groups in their chemical structure such as -NH2, -SH, and -COOH can interact with the nanoparticle surface of the zero-valent iron by ionic interaction or by formation of -OH bonds on their oxidized surface. It is very important to choose the correct amino acid to mediate the generation of the zero-valent iron nanoparticles. L-glutamic acid proved to be the most effective while L-cysteine gave rise to the transformation of zero-valent iron into oxides and oxyhydroxides [400].
Organic dyes used in the textile and leather industry are degraded by using nZVI nanoparticles formed from extracts of green tea, oolong tea, black tea, eucalyptus plant leaves, Ferula persica root, pomegranate leaf, and grape, following pseudo-first- and pseudo-second-order kinetic models (such as bromothymol blue, orange II, malachite green, methylene blue, methyl orange, crystal violet, and black acid 194). Activation energies less than 15 kJ/mol indicate that diffusion is the predominant phenomenon during adsorption. The degradation of orange II in the LC-MS analysis shows the presence of intermediates such as 1,2-dihydroxynaphthalene, 1-diazo-2-naphthol, 2-naphthol 4-hydrobenzenesulfonate, 4-sulfophenylhydroperoxide, and benzene sulfonate, suggesting that orange II is adsorbed on the surface of the nanoparticle [401]. Possibly, by the interaction of FeO(OH) groups and iron oxides followed by the excision of the azo bond by the zero-valent iron electrons present in the core.
The bleaching of methylene blue or methyl orange can be attributed to a process involving the cleavage of C=C and C=N bonds of the dyes. It can also be explained by an enhanced mass transfer rate at the surface of the nanoparticles followed by their adsorption on the reactive sites of the nanoparticles. Although ultrasound generates some ·OH radicals, the amount remains too low to drive the reaction to a significant extent. The decrease in total organic carbon (TOC) could be attributed to fixation of the dye molecule by long-chain capping agents, followed by co-precipitation and adsorption. The superior removal performance obtained with ultrasound was explained by Wang et al. through a better mass transfer rate of dyes on the nZVI surface, followed by adsorption on reactive sites of the nZVI particles [402]. The electrons generated by Fe0 can directly migrate to dyes. Subsequent trapping by H2O/H+ yields highly reactive hydrogen species (H*) or dissociation of dissolved oxygen, generating strong oxidative species such as CHO2, CO, and COH, followed by cleavage of the chromophore groups and conjugated systems. Furthermore, corrosion products deposited on functional nZVIs (TP-Fe) [402] could be expelled by ultrasound and replenish the surface of nZVIs. Coagulation of the corrosion products (i.e., iron hydroxides/oxides) leads to sedimentation and co-precipitation, which might be the mechanisms responsible for the TOC removal process.
In recent studies, green synthesis of iron nanoparticles by tea (Camellia sinensis) po-lyphenols has been used for the removal of bromothymol blue. The nanoparticles created were spherical with a size range of 5–15 nm. The degradation of bromothymol blue was performed with H2O2 and synthesized nZVIs, the kinetics of the reactions ran through a first-order equation depending on the concentration of bromothymol blue. The results suggested a decrease in the rate of bromothymol degradation with increasing amounts of iron up to a maximum of 0.55 mM. In addition, EDTA was used to increase the stability of hydrogen peroxide and slow down its decomposition and production of hydroxyl radicals [403].
According to Shahwan [404], the removal efficiency of dyes is higher with iron nanoparticles obtained from green tea extract and used as a Fenton-type catalyst than if iron nanoparticles are obtained with sodium borohydride. These nanoparticles act as a substrate for the generation of ferrous ions with a consequent decrease in pH (to ~3.5) and increase in the concentration of ·OH radicals. As for the kinetics of the removal process, it is faster when green-synthesized nanoparticles are used. It should be noted that the kinetics of the removal process runs through a second-order or first-order equation.
Huang et al. [405] used oolong tea extract for the reductive degradation of malachite green in aqueous solution. The removal rate decreased with aggregation and oxidation of the nanoparticles, so the size and reactivity of the nanoparticles played an important role in their degradation. It is very likely that the final structures and size of the nZVIs were related to the polyphenol/caffeine concentrations present in the tea extracts. Spherical nanoparticles within the range of size of 40–50 nm were synthesized. XRD showed characteristic peaks corresponding to iron structures such as maghemite (Fe2O3), magnetite (Fe3O4), and iron oxyhydroxide (FeO(OH)), and the Fe0 peak was small as it was not very crystalline in nature. When iron nanoparticles from oolong tea were used, the removal efficiency improved as the contact time increased, reaching equilibrium in 60 min and a removal rate of 75.5%. This is attributable to the fact that the polyphenol and caffeine content in the oolong tea extract not only act as coating agents, preventing the aggregation of the nanoparticles, but are also reducing agents in the synthesis of nanoparticles. The kinetics conformed to a pseudo-first-order kinetic equation with high correlation coefficients (r2 > 0.98) and whose activation energy was of the order of 23.86 kJ/mol, suggesting that the relationship was controlled by the chemical diffusion factor. According to the EDS data, the compositions of the oolong tea extract were 20.34% for O, 49.51% for C, 5.05% for S, and 15.10% for K.
Luo et al. [401] address the synthesis of iron nanoparticles using vine leaf extract and apply them to the elimination of orange acid II. The biomolecule extract in methanol and in aqueous medium is compared. The methanol extract has a high concentration of biomolecules capable of increasing the reduction of iron II to zero-valent iron, and thus, increasing the concentration of iron nanoparticles, which results in a high reactivity in the removal of the dye (removal is 80% for the methanolic extract and only 4% when using the aqueous extract). This is attributed to the high concentration of polyphenols and other biomolecules that not only act as reducing agents but also as coating agents, preventing the nanoparticles from aggregating. As a result, the nanoparticles are more stable and reactive. The authors propose a degradation mechanism based on asymmetric cleavage of azo bonds for the removal of orange II when using iron nanoparticles prepared from vine leaf extract.
From eucalyptus leaves, iron nanoparticles are obtained for the elimination of the acid black dye 194, forming an iron–polyphenol complex which is used as a flocculant [296], containing a large amount of hydroxyl groups and acting as chelating agents to form the iron–polyphenol complex. The formed iron–polyphenol nanoparticles (nZVIs) are amorphous, with sizes in the range of 60 nm and with an iron(III) ion located in the chelated nanoparticle. The nanoflocculant can remove acid black 194 more efficiently than FeCl3 for the same dose (1 mL) 80.5% and 63%, respectively.
Removal of crystal violet in aqueous media was carried out by Nasiri et al. [406] using zero-valent iron nanoparticles functionalized with β-cyclodextrin from Ferula persica root extracts. When a surfactant (β-cyclodextrin) was added to the plant extract, the BET surface area of the nanoparticle increased (from 46.68 to 47.10 m2/g), its size decreased (from 34.20 to 24.20 nm), and its morphology was spherical. Biomolecules extracted also served as a coating agent, stabilizing the zero-iron nanoparticle and decreasing the concentration of iron oxides.
Table 6. Synthesis of Fe nanoparticles by green chemistry for the removal of dyes.
Table 6. Synthesis of Fe nanoparticles by green chemistry for the removal of dyes.
NPsLeaf Extract
of Plant
Fe SourceDyeConcentration
(mgL−1)
Removal
(%)
React. Time
(min)
Reference
nZVIGreen teaFe (NO3)3∙9H2OMalachite green509360[294]
nZVIEucalyptusFeCl3·6H2OAcid black 19410080.5200[296]
nZVIEucalyptus L.e.FeSO4·7H2OCrystal violet3097.61800[311]
nZVIBlack teaFeCl3·6H2OReactive blue 23849.690.560[317]
nZVIRuellia tuberosaFeCl3·6H2OReactive black 525–40095.3–99.830[320]
nZVIArtocarpus
heterophyllus
FeCl3·6H2OFuchsin basic487.520[323]
nZVIRicinus communisFeCl3·6H2OMethylene blue25–20096.8160[326]
nZVIGreen teaFeCl3·6H2OReactive blue 23849.596.240[330]
nZVIMango peelFeCl3·6H2OMethyl orange10094.2360[341]
nZVIPomegranate L.e.FeCl3·6H2OMalachite green1569536[396]
nZVIVine L.e.FeCl2·4H2OOrange II10080120[398]
nZVIVine L.e.FeCl2·4H2OOrange II10092200[401]
nZVITeaFeSO4·7H2OMalachite green,
methylene blue
20090.7, 90.7530[402]
nZVICamellia sinensis teaFeCl3·6H2OBromothymol blue500-60[403]
nZVIGreen teaFeCl2·4H2OMethylene blue,
methyl orange
5036080[404]
nZVIOolong teaFeSO4·7H2OMalachite green5075.5-[405]
nZVIFerula persicaFeSO4·7H2OCrystal violet20099.8210[406]
nZVICatharanthus roseusFeSO4·7H2OMethyl orange5050360[407]
α-Fe2O3Phyllanthus niruri,
Moringa stenopetala
FeCl3·6H2OMethylene blue15692–9636[408]
nZVIChlorophytum
comosum
FeCl3·6H2OMethyl orange47725[409]
nZVIPartheniumFeSO4·7H2OCrystal violet1209530[410]
Fe3O4Thunbergia
grandiflora
FeSO4·7H2OAcid blue 1132594.38210[411]
Fe2O3Raphanus sativus L.e.FeCl3·6H2OMethylene blue,
methyl red
10010060[412]
Fe2O3Fan palm,
Dombeya wallichii,
Pyrus communis
FeSO4·7H2OReactive blue20094, 81, 88210[413]
Fe3O4Jatropha curcas,
Cinnamomum tamala
FeCl2·H2O+
FeCl3
Methylene blue20046.66120[414]
Iron oxideArtemisia vulgaris L.e.FeCl3·6H2OMethyl orange2598.6360[415]
Iron oxideDaphne mezereumFeCl3·6H2OMethyl orange2581360[416]
Fe3O4Fraxinus sinensis RoxbFeCl3·6H2O+
FeSO4·7H2O
Crystal violet498.5725[417]
nClusterCupressus sempervirensFeCl3·6H2OMethyl orange2595.8360[418]
nZVIPsidium guajava L.e.FeCl3·6H2OMethylene blue50945[419]
nZVIHibiscus sabdariffa,
roselle flower
FeCl3·6H2ORhodamine B171005[420]
nZVITrigonella
foenum-graecum
FeCl3·6H2OMethyl orange259590[421]
ZVICamellia sinensis and
pomegranate L.e.
FeSO4·7H2OTextile wastewater2.330
-
>95
pH 8.5
(Pt-Co)
120[422]
Fe3O4Azolla filiculoides
and fig. L.e.
FeCl3·6H2O+
FeSO4·4H2O
Crystal violet
Methylene blue
500100210[423]
Fe3O4Pissum sativum peelFeCl3·6H2OMethyl orange10096.2360[424]
nZVIEucalyptusFeSO4·7H2OMethyl orange1099.6180[425]
Iron oxidePomegranate L.e.FeCl3·6H2OCongo red1009360[426]
Ni-iron
oxide
Moringa oleiferaFeCl3·6H2O +
NiCl2·6H2O
Malachite green20~91.625[427]
Fe3O4Green tea L.e.FeCl3·6H2O+
FeSO4∙ 4H2O
Methylene blue3.59516[428]
Pd-iron
oxide
PepperFe (NO3)3∙9H2O+
FeCl2·4H2O
Acid black,
acid brown
20.297.85120[429]
Iron oxideCynometra ramiflora fruitFeCl2·H2O+FeCl3Methylene blue20100110[430]
Iron oxideCynometra ramifloraFeSO4·7H2ORhodamine B13410015[431]
Iron oxideS. cuminiFerrous oxalateReactive blue 2351098.75240[432]
nZVITeaFeSO4·7H2OMethylene blue5085.7, 34.45[433]
Fe3O4Ridge gourd peelsFeCl3·6H2OMethylene blue1209630[434]
nZVITieguanyin teaFeCl3·6H2OBromothymol blue100>9030[435]
nZVIGreen teaFeCl3·6H2ORBB-R,
DR80 mixture
2509020[436]
Fe3O4Maize cobFeCl2·4H2O+
2FeCl3·6H2O
Methylene blue15699.6336[437]
nZVIEucalyptus tereticornis,
Melaleuca nesophila, Rosemarinus officinalis
FeSO4·7H2OAcid black 194100100200[438]
nZVIGreen tea,
oolong tea,
black tea
FeSO4·7H2OMalachite green, methylene blue5081.2, 75.6,
67.1
60[439]
nZVIAspalathus linearisAcid mine
drainage
Orange II509430[440]
nZVICalotropis giganteaFe (NO3)3·9H2OMethylene blue50–40083.930[441]

nZVI
Amaranthus dubius L.e.FeCl3·6H2OMethyl orange2081360[442]
Note: L.e.: Leaf extract.

7. Future Perspectives and Challenges

Environmental issues are not inherent to chemistry per se; rather, they frequently originate from anthropogenic processes or the imprudent utilization of chemical knowledge, resulting in adverse environmental ramifications. Consequently, there exists a pressing need to adopt clean and sustainable methodologies, as well as to utilize raw materials derived from nature.
This review undertakes a comprehensive analysis of methodologies employed for the removal of dyes, with a specific focus on nanoparticles synthesized through both conventional and green chemistry techniques. Notably, this entails the utilization of plant extracts as reductants for the formation of nZVIs. Such an approach prioritizes environmental safety and long-term sustainability over immediate gains. This necessitates exploration into more environmentally benign alternatives, refinement of Fe nanoparticle synthesis processes for heightened efficacy, and advocacy for responsible chemical practices.
Some key aspects that need to be addressed in the near future are summarized below.
Critical to note is the considerable variability in extracted biomolecules from plant extracts, constituting an unpredictable factor in green chemistry nanoparticle synthesis. The diverse nature of plants, encompassing distinct plant parts and geographical locales, complicates the prognostication of extracted biomolecules and their quantities. Moreover, not all biomolecules involved in nanoparticle synthesis are comprehensively understood. Controllable factors such as solvent, temperature, pH, precursor iron salts, and solution agitation profoundly influence nanoparticle growth, morphology, size, aggregation, coating, and stability within green chemistry procedures.
Furthermore, it is imperative to recognize the industrial implications of nanoparticle synthesis, necessitating large-scale production. Consequently, material supply assumes critical importance, with the agrifood industry wielding significant influence due to the pronounced variability influenced by geographical location and climate.
A standardized protocol is imperative for achieving greater reproducibility in nanoparticle synthesis concerning size and shape, thereby optimizing dye decontamination processes in the environment. Furthermore, the development of methodologies aimed at minimizing waste generation or enabling waste reuse is critical for fostering more sustainable processes. The extraction processes from agrifood industries entail significant solvent consumption, resulting in solid residues, necessitating a reduction in energy consumption. Analogous to numerous other fields, the reutilization and recycling of these by-product streams are pivotal for practical industrial implementation.
Regarding applications, substantial challenges persist. Ongoing efforts are directed towards exploring novel applications in diverse sectors such as food, textiles, healthcare, and construction, extending beyond environmental objectives.
In the realm of water purification, wastewater treatment plants are transitioning towards pollutant mineralization and biorefinery models, aiming to derive commercially viable products. Accordingly, pollutant removal processes must align with this philosophy and be seamlessly integrated into wastewater treatment plants. This mandates that new nanoparticles, acting as catalysts, should adsorb and convert pollutants into products that can be easily desorbed at low energy costs.
Regarding the main challenges that must be faced in connection with the use of nZVIs in decontamination processes of waters polluted with dyes, some crucial aspects have to be taken into account.
The authors underscore the potential for the wastewater industry to operate as a biorefinery, extracting commercial products from wastewater. However, these treatment methodologies are still nascent, as depicted in Figure 6.
It is noteworthy that a mere 242 articles, less than 0.001%, focus on dye recovery, with the majority concentrating on repurposing waste for dye treatment using alternative methods. Only a fraction of these articles address the recovery of solid waste from the textile industry [443,444]. Marazzi et al. [445] propose the use of algae for the treatment of dyes and their subsequent transformation into energy in biogas plants through anaerobic digestion.
Since clean, high-quality water is a valuable and essential commodity, and given that one of the main and most obvious parameters indicating water quality is its color, the application of iron nanoparticles as an available technology for dye removal is easy to use, cost-effective, and very efficient. The two main removal processes are adsorption and decolorization. This implies that studies should be directed to investigate the optimal conditions of these processes, such as the influence of the effects of initial dye concentration, pH, temperature, as well as nanoparticle size, morphology, and dosage, and to generate new general trends based on these studies.
A plethora of literature exists on environmentally friendly methods, often associated with biomass or bio-resources as feedstock sources. While these methodologies hold significant promise, presuming their environmental friendliness without thorough environmental impact assessments would be erroneous [446]. Thus, it is imperative to evaluate the environmental footprint of the chosen method to ensure its sustainability and assess its economic viability for industrialization. Furthermore, methods employing industrial chemical reagents can be rendered environmentally friendly provided energy consumption is minimized, and by-products or residues are recycled or integrated into a circular economy framework.
Figure 6. Number of published articles for each treatment method (a), and schematic mechanism of each treatment method (b).
Figure 6. Number of published articles for each treatment method (a), and schematic mechanism of each treatment method (b).
Water 16 01607 g006

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/w16111607/s1, Table S1: Common names, chemical structure, and IUPAC names of the different dyes referred to in this work.

Author Contributions

Conceptualization, E.M.C.-C. and M.F.A.-F.; methodology, M.F.A.-F., C.R.-R., C.F.-G. and V.M.-J.; formal analysis, M.F.A.-F., V.M.-J. and C.R.-R.; investigation, E.M.C.-C., M.F.A.-F., C.R.-R., C.F.-G. and V.M.-J.; resources, M.F.A.-F., C.R.-R. and C.F.-G.; writing—original draft preparation, M.F.A.-F., C.R.-R., C.F.-G., V.M.-J. and J.P.-T.; writing—review and editing, M.F.A.-F. and C.R.-R.; supervision, E.M.C.-C. and M.F.A.-F. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zaki, M.S.; Abou Zaid, A.A.; Abdelzaher, M.F.; Shalaby, S.I. Clinicopathological Changes in Fish Exposed to Pollutants. Life Sci. J. 2014, 11, 271–278. [Google Scholar]
  2. Dwivedi, S.; Shikha, D. Water Pollution: Causes, Effects and Control. Biochem. Cell Arch. 2016, 16, 96–102. [Google Scholar]
  3. Agoha, E.E.C. Crude Oil in Drinking Water: Chitosan Intervention. IFMBE Proc. 2019, 68, 741–743. [Google Scholar] [CrossRef]
  4. Larios-Meoño, J.F.; Morales, Y.; Gonzalez-Taranco, C.; Mougenot, B. Impact of Water Quality on the Economy and Health of Latin America: The Case of Peru and Colombia. In Proceedings of the 33rd International Business Information Management Association Conference, IBIMA 2019: Education Excellence and Innovation Management through Vision 2020, Granada, Spain, 10–11 April 2019; pp. 8960–8969. [Google Scholar]
  5. Gupta, J.K.; Shah, K.; Mishra, P. Inadmissible Planktons in Potable Water: A Potential Risk for Human Health. Curr. Sci. 2020, 119, 1627–1632. [Google Scholar] [CrossRef]
  6. Russ, J.; Zaveri, E.; Desbureaux, S.; Damania, R.; Rodella, A.-S. The Impact of Water Quality of GDP Growth: Evidence from around the World. Water Secur. 2022, 17, 100130. [Google Scholar] [CrossRef]
  7. du Plessis, A. Global Water Quality Challenges; Springer: Cham, Switzerland, 2017. [Google Scholar] [CrossRef]
  8. Das, J.; Sarkar, S.; Saha, K.; Roy, S.; Nag, S. Emerging Environmental Contaminants: Sources, Consequences, and Future Challenges. In Bioremediation Technologies: For Wastewater and Sustainable Circular Bioeconomy; Walter de Gruyter: Berlin, Germany, 2023; ISBN 978-3-11101-682-5. [Google Scholar] [CrossRef]
  9. Snyder, S.A.; Anumol, T. Emerging Chemical Contaminants: How Chemical Development Outpaces Impact Assessment. In Still Only One Earth: Progress in the 40 Years Since the First UN Conference on the Environment; Royal Society of Chemistry: London, UK, 2015; Volume 2015, ISBN 978-1-78262-076-1. [Google Scholar] [CrossRef]
  10. Stefanakis, A.I.; Becker, J.A. A Review of Emerging Contaminants in Water: Classification, Sources, and Potential Risks. In Impact of Water Pollution on Human Health and Environmental Sustainability; IGI Global: Pennsylvania, PA, USA, 2019; ISBN 978-1-79981-211-1. [Google Scholar] [CrossRef]
  11. Sivasubramaniyan, S.G.; Kandasamy, S.; Manickam, N.K. Biotechnological Approaches for Removal of Emerging Contaminants. In Biotechnology for Zero Waste: Emerging Waste Management Techniques; Wiley: Hoboken, NJ, USA, 2021; ISBN 978-3-52783-206-4. [Google Scholar] [CrossRef]
  12. Khan, S.; Naushad, M.; Govarthanan, M.; Iqbal, J.; Alfadul, S.M. Emerging Contaminants of High Concern for the Environment: Current Trends and Future Research. Environ. Res. 2022, 207, 112609. [Google Scholar] [CrossRef]
  13. Ngeno, E.; Shikuku, V.O. Emerging Contaminants: Pollution Control and Abatement. In Research Anthology on Emerging Techniques in Environmental Remediation; IGI Global: Pennsylvania, PA, USA, 2022; Volume 2, ISBN 978-1-66843-886-2. [Google Scholar] [CrossRef]
  14. Lofrano, G.; Sacco, O.; Venditto, V.; Carotenuto, M.; Libralato, G.; Guida, M.; Meric, S.; Vaiano, V. Occurrence and Potential Risks of Emerging Contaminants in Water. In Visible Light Active Structured Photocatalysts for the Removal of Emerging Contaminants; Elsevier: Amsterdam, The Netherlands, 2020; ISBN 978-0-12818-334-2. [Google Scholar] [CrossRef]
  15. Morin-Crini, N.; Lichtfouse, E.; Liu, G.; Balaram, V.; Ribeiro, A.R.L.; Lu, Z.; Stock, F.; Carmona, E.; Teixeira, M.R.; Picos-Corrales, L.A.; et al. Worldwide Cases of Water Pollution by Emerging Contaminants: A Review. Environ. Chem. Lett. 2022, 20, 2311–2338. [Google Scholar] [CrossRef]
  16. Sauvé, S.; Desrosiers, M. A Review of What Is an Emerging Contaminant. Chem. Cent. J. 2014, 8, 15. [Google Scholar] [CrossRef]
  17. Noguera-Oviedo, K.; Aga, D.S. Lessons Learned from More than Two Decades of Research on Emerging Contaminants in the Environment. J. Hazard. Mater. 2016, 316, 242–251. [Google Scholar] [CrossRef] [PubMed]
  18. Lellis, B.; Fávaro-Polonio, C.Z.; Pamphile, J.A.; Polonio, J.C. Effects of Textile Dyes on Health and the Environment and Bioremediation Potential of Living Organisms. Biotechnol. Res. Innov. 2019, 3, 275–290. [Google Scholar] [CrossRef]
  19. Tkaczyk, A.; Mitrowska, K.; Posyniak, A. Synthetic Organic Dyes as Contaminants of the Aquatic Environment and Their Implications for Ecosystems: A Review. Sci. Total Environ. 2020, 717, 137222. [Google Scholar] [CrossRef]
  20. Jankowska, A.; Ejsmont, A.; Galarda, A.; Goscianska, J. The Outcome of Human Exposure to Environmental Contaminants. Importance of Water and Air Purification Processes. In Sustainable Materials for Sensing and Remediation of Noxious Pollutants; Elsevier: Amsterdam, The Netherlands, 2022; ISBN 978-0-32399-425-5. [Google Scholar] [CrossRef]
  21. Rahman, M.; Tabassum, Z. Biotechnological Approach to Treat Textile Dyeing Effluents: A Critical Review Analysing the Practical Applications. Text. Leather Rev. 2024, 7, 124–152. [Google Scholar] [CrossRef]
  22. Wang, S.; Wang, Q.; Zhang, T.; Liu, S.; Ho, S.S.H.; Tian, J.; Su, H.; Zhang, Y.; Wang, L.; Wu, T.; et al. Elaborations of the Influencing Factors on the Formation of Secondary Inorganic Aerosols in a Heavily Polluted Urban Area of China. J. Environ. Sci. 2024, 138, 406–417. [Google Scholar] [CrossRef]
  23. Bai, Y.; Lin, H.; Wang, C.; Wang, Q.; Qu, J. Digitalizing River Aquatic Ecosystems. J. Environ. Sci. 2024, 137, 677–680. [Google Scholar] [CrossRef]
  24. Brown, A.M.; Bass, A.M.; Skiba, U.; MacDonald, J.M.; Pickard, A.E. Urban Landscapes and Legacy Industry Provide Hotspots for Riverine Greenhouse Gases: A Source-to-Sea Study of the River Clyde. Water Res. 2023, 236, 119969. [Google Scholar] [CrossRef]
  25. Wei, Z.; Wei, Y.; Liu, Y.; Niu, S.; Xu, Y.; Park, J.-H.; Wang, J.J. Biochar-Based Materials as Remediation Strategy in Petroleum Hydrocarbon-Contaminated Soil and Water: Performances, Mechanisms, and Environmental Impact. J. Environ. Sci. 2024, 138, 350–372. [Google Scholar] [CrossRef]
  26. Shi, C.; Liu, Z.; Yu, B.; Zhang, Y.; Yang, H.; Han, Y.; Wang, B.; Liu, Z.; Zhang, H. Emergence of Nanoplastics in the Aquatic Environment and Possible Impacts on Aquatic Organisms. Sci. Total Environ. 2024, 906, 167404. [Google Scholar] [CrossRef]
  27. Castaño-Ortiz, J.; Gil-Solsona, R.; Ospina-Álvarez, N.; Alcaraz-Hernández, J.; Farré, M.; León, V.; Barceló, D.; Santos, L.; Rodríguez-Mozaz, S. Fate of Pharmaceuticals in the Ebro River Delta Region: The Combined Evaluation of Water, Sediment, Plastic Litter, and Biomonitoring. Sci. Total Environ. 2024, 906, 167467. [Google Scholar] [CrossRef]
  28. Khezami, F.; Gómez-Navarro, O.; Barbieri, M.V.; Khiari, N.; Chkirbene, A.; Chiron, S.; Khadhar, S.; Pérez, S. Occurrence of Contaminants of Emerging Concern and Pesticides and Relative Risk Assessment in Tunisian Groundwater. Sci. Total Environ. 2024, 906, 167319. [Google Scholar] [CrossRef]
  29. Hung, V.Q.; Egodawatta, P.; Gallage, C.; Dawes, L.; Nguyen-Xuan, T.; Nguyen-Viet, T.; Bui-Tien, T.; Nguyen-Quang, T.; De Roeck, G. Leaching Mechanism of Metals from Recycled Concrete Aggregates (RCA) and Potentially Environmental Issues. In Proceedings of the 4th International Conference on Sustainability in Civil Engineering, Hanoi, Vietnam, 25–27 November 2024; Volume 344, ISBN 978-981-99-2344-1. [Google Scholar] [CrossRef]
  30. Tudor, V.C.; Stoicea, P.; Chiurciu, I.-A.; Soare, E.; Iorga, A.M.; Dinu, T.A.; David, L.; Micu, M.M.; Smedescu, D.I.; Dumitru, E.A. The Use of Fertilizers and Pesticides in Wheat Production in the Main European Countries. Sustainability 2023, 15, 3038. [Google Scholar] [CrossRef]
  31. Bukomeko, H.; Taulya, G.; Schut, A.G.T.; van de Ven, G.W.J.; Kubiriba, J.; Giller, K. Evaluating Combined Effects of Pesticide and Crop Nutrition (with N, P, K and Si) on Weevil Damage in East African Highland Bananas. PLoS ONE 2023, 18, e0282493. [Google Scholar] [CrossRef] [PubMed]
  32. Mir, A.A.; Ahad, U.; Inayatullah, M.; Ali, U.; Ahmed, P. Evaluation of Water Quality Status of Pohru Watershed, Kashmir Valley, Jammu and Kashmir, India. Water Air Soil Pollut. 2023, 234, 154. [Google Scholar] [CrossRef]
  33. Holkar, C.R.; Jadhav, A.J.; Pinjari, D.V.; Mahamuni, N.M.; Pandit, A.B. A Critical Review on Textile Wastewater Treatments: Possible Approaches. J. Environ. Manage. 2016, 182, 351–366. [Google Scholar] [CrossRef] [PubMed]
  34. Anandita; Raees, K.; Shahadat, M.; Ali, S.W. Mechanistic Interaction of Microbe in Dye Degradation and the Role of Inherently Modified Organisms: A Review. Water Conserv. Sci. Eng. 2023, 8, 43. [Google Scholar] [CrossRef]
  35. Boretti, A.; Rosa, L. Reassessing the Projections of the World Water Development Report. NPJ Clean Water 2019, 2, 15. [Google Scholar] [CrossRef]
  36. Pokrajac, L.; Abbas, A.; Chrzanowski, W.; Dias, G.M.; Eggleton, B.J.; Maguire, S.; Maine, E.; Malloy, T.; Nathwani, J.; Nazar, L.; et al. Nanotechnology for a Sustainable Future: Addressing Global Challenges with the International Network4Sustainable Nanotechnology. ACS Nano 2021, 15, 18608–18623. [Google Scholar] [CrossRef] [PubMed]
  37. Saleh, H.M.; Hassan, A.I. Synthesis and Characterization of Nanomaterials for Application in Cost-Effective Electrochemical Devices. Sustainability 2023, 15, 10891. [Google Scholar] [CrossRef]
  38. Li, L.; Hu, J.; Shi, X.; Fan, M.; Luo, J.; Wei, X. Nanoscale Zero-Valent Metals: A Review of Synthesis, Characterization, and Applications to Environmental Remediation. Environ. Sci. Pollut. Res. 2016, 23, 17880–17900. [Google Scholar] [CrossRef] [PubMed]
  39. Crane, R.A.; Scott, T.B. Nanoscale Zero-Valent Iron: Future Prospects for an Emerging Water Treatment Technology. J. Hazard. Mater. 2012, 211–212, 112–125. [Google Scholar] [CrossRef]
  40. Malik, S.; Muhammad, K.; Waheed, Y. Nanotechnology: A Revolution in Modern Industry. Molecules 2023, 28, 661. [Google Scholar] [CrossRef]
  41. Hasirci, V.; Vrana, E.; Zorlutuna, P.; Ndreu, A.; Yilgor, P.; Basmanav, F.B.; Aydin, E. Nanobiomaterials: A Review of the Existing Science and Technology, and New Approaches. J. Biomater. Sci. 2006, 17, 1241–1268. [Google Scholar] [CrossRef]
  42. Khin, M.M.; Nair, A.S.; Babu, V.J.; Murugan, R.; Ramakrishna, S. A Review on Nanomaterials for Environmental Remediation. Energy Environ. Sci. 2012, 5, 8075–8109. [Google Scholar] [CrossRef]
  43. Mi, J.L.; Nørby, P.; Bremholm, M.; Becker, J.; Iversen, B.B. The Formation Mechanism of Bimetallic PtRu Alloy Nanoparticles in Solvothermal Synthesis. Nanoscale 2015, 7, 16170–16174. [Google Scholar] [CrossRef] [PubMed]
  44. Veronico, L.; Gentile, L. Removal of Pollutants by Ferrihydrite Nanoparticles Combined with Brij L4 Self-Assembled Nanostructures. ACS Appl. Nano Mater. 2023, 6, 720–728. [Google Scholar] [CrossRef]
  45. Ganesan, S.; Lakshmisankar, S.; Deepak, S.; Selvakumar, M.; Venkatesan, S.P.; Hemanandh, J. Environmental Impact of One-Step Hydrothermal Green Synthesized ZnO Nanoparticles in DI Diesel Engine Performance Analysis Using Duel Fuel. Nanotechnol. Environ. Eng. 2023, 8, 361–375. [Google Scholar] [CrossRef]
  46. Muraro, P.C.L.; Wouters, R.D.; Pavoski, G.; Espinosa, D.C.R.; Ruiz, Y.P.M.; Galembeck, A.; Rech, V.C.; da Silva, W.L. Ag/TiNPS Nanocatalyst: Biosynthesis, Characterization and Photocatalytic Activity. J. Photochem. Photobiol. A Chem. 2023, 439, 114598. [Google Scholar] [CrossRef]
  47. Khan, I.; Saeed, K.; Khan, I. Nanoparticles: Properties, Applications and Toxicities. Arab. J. Chem. 2019, 12, 908–931. [Google Scholar] [CrossRef]
  48. Guerra, F.D.; Attia, M.F.; Whitehead, D.C.; Alexis, F. Nanotechnology for Environmental Remediation: Materials and Applications. Molecules 2018, 23, 1760. [Google Scholar] [CrossRef] [PubMed]
  49. Westerhoff, P.; Song, G.; Hristovski, K.; Kiser, M.A. Occurrence and Removal of Titanium at Full Scale Wastewater Treatment Plants: Implications for TiO2 Nanomaterials. J. Environ. Monit. 2011, 13, 1195–1203. [Google Scholar] [CrossRef] [PubMed]
  50. Batley, G.E.; Kirby, J.K.; McLaughlin, M.J. Fate and Risks of Nanomaterials in Aquatic and Terrestrial Environments. Acc. Chem. Res. 2013, 46, 854–862. [Google Scholar] [CrossRef]
  51. Gallocchio, F.; Biancotto, G.; Moressa, A.; Pascoli, F.; Pretto, T.; Toffan, A.; Arcangeli, G.; Montesi, F.; Peters, R.; Ricci, A. Bioaccumulation and in Vivo Formation of Titanium Dioxide Nanoparticles in Edible Mussels. Food Chem. 2020, 323, 126841. [Google Scholar] [CrossRef] [PubMed]
  52. Baig, N.; Kammakakam, I.; Falath, W. Nanomaterials: A Review of Synthesis Methods, Properties, Recent Progress, and Challenges. Mater. Adv. 2021, 2, 1821–1871. [Google Scholar] [CrossRef]
  53. Khanna, P.; Kaur, A.; Goyal, D. Algae-Based Metallic Nanoparticles: Synthesis, Characterization and Applications. J. Microbiol. Methods 2019, 163, 105656. [Google Scholar] [CrossRef] [PubMed]
  54. Kumar, J.A.; Krithiga, T.; Manigandan, S.; Sathish, S.; Renita, A.A.; Prakash, P.; Prasad, B.S.N.; Kumar, T.R.P.; Rajasimman, M.; Hosseini-Bandegharaei, A.; et al. A Focus to Green Synthesis of Metal/Metal Based Oxide Nanoparticles: Various Mechanisms and Applications towards Ecological Approach. J. Clean. Prod. 2021, 324, 129198. [Google Scholar] [CrossRef]
  55. Alzahrani, H.A.H.; Almulaiky, Y.Q.; Alsaiari, A.O. The Photocatalytic Dye Degradation of Methylene Blue (MB) by Nanostructured ZnO under UV Irradiation. Phys. Scr. 2023, 98, 045703. [Google Scholar] [CrossRef]
  56. Soleimani-Gorgani, A.; Al-Sabahi, J.; Nejad, S.A.T.; Heydari, M.; Al-Abri, M.; Namaeighasemi, A. Visible-Light-Driven Super-Active Sn and GO Single- and Sn/Cu Co-Doped Nanophotocatalysts for Phenol Degradation: Thin-Film Printability, Thermal Stability, and Cytotoxicity Assay. J. Ind. Eng. Chem. 2023, 120, 514–528. [Google Scholar] [CrossRef]
  57. Hamad, M.T.M.H.; El-Sesy, M.E. Adsorptive Removal of Levofloxacin and Antibiotic Resistance Genes from Hospital Wastewater by Nano-Zero-Valent Iron and Nano-Copper Using Kinetic Studies and Response Surface Methodology. Bioresour. Bioprocess. 2023, 10, 1. [Google Scholar] [CrossRef] [PubMed]
  58. Ranjith, R.; Vignesh, S.; Balachandar, R.; Suganthi, S.; Raj, V.; Ramasundaram, S.; Kalyana Sundar, J.; Shkir, M.; Oh, T.H. Construction of Novel G-C3N4 Coupled Efficient Bi2O3 Nanoparticles for Improved Z-Scheme Photocatalytic Removal of Environmental Wastewater Contaminant: Insight Mechanism. J. Environ. Manag. 2023, 330, 117134. [Google Scholar] [CrossRef] [PubMed]
  59. Manikandan, V.; Anushkkaran, P.; Hwang, I.S.; Chae, W.S.; Lee, H.H.; Choi, S.H.; Mahadik, M.A.; Jang, J.S. Synergistic Role of In-Situ Zr-Doping and Cobalt Oxide Cocatalysts on Photocatalytic Bacterial Inactivation and Organic Pollutants Removal over Template-Free Fe2O3 Nanorods. Chemosphere 2023, 310, 136825. [Google Scholar] [CrossRef]
  60. Choi, C.J.; Dong, X.L.; Kim, B.K. Characterization of Fe and Co Nanoparticles Synthesized by Chemical Vapor Condensation. Scr. Mater. 2001, 44, 2225–2229. [Google Scholar] [CrossRef]
  61. Sun, H.; Zhou, G.; Liu, S.; Ang, H.M.; Tadé, M.O.; Wang, S. Nano-Fe0 Encapsulated in Microcarbon Spheres: Synthesis, Characterization, and Environmental Applications. ACS Appl. Mater. Interfaces 2012, 4, 6235–6241. [Google Scholar] [CrossRef] [PubMed]
  62. Li, J.; Chen, C.; Zhu, K.; Wang, X. Nanoscale Zero-Valent Iron Particles Modified on Reduced Graphene Oxides Using a Plasma Technique for Cd(II) Removal. J. Taiwan. Inst. Chem. Eng. 2016, 59, 389–394. [Google Scholar] [CrossRef]
  63. Amerhaider Nuar, N.N.; Siti Nurul, S.N.A.; Choong, T.S.Y.; Mat Azmi, I.D.; Abdul Romli, N.A.; Abdullah, L.C.; Chiang, P.C.; Li, F. Synthesis of Calcium Peroxide Nanoparticles with Starch as a Stabilizer for the Degradation of Organic Dye in an Aqueous Solution. Polymers 2023, 15, 1327. [Google Scholar] [CrossRef] [PubMed]
  64. Vijayan, M.; Manikandan, V.; Rajkumar, C.; Hatamleh, A.A.; Alnafisi, B.K.; Easwaran, G.; Liu, X.; Sivakumar, K.; Kim, H. Constructing Z-Scheme g-C3N4/TiO2 Heterostructure for Promoting Degradation of the Hazardous Dye Pollutants. Chemosphere 2023, 311, 136928. [Google Scholar] [CrossRef] [PubMed]
  65. Xia, Q.; Jiang, Z.; Wang, J.; Yao, Z. A Facile Preparation of Hierarchical Dendritic Zero-Valent Iron for Fenton-like Degradation of Phenol. Catal. Commun. 2017, 100, 57–61. [Google Scholar] [CrossRef]
  66. El-Sheekh, M.; Elshobary, M.; Abdullah, E.; Abdel-Basset, R.; Metwally, M. Application of a Novel Biological-Nanoparticle Pretreatment to Oscillatoria Acuminata Biomass and Coculture Dark Fermentation for Improving Hydrogen Production. Microb. Cell Fact. 2023, 22, 34. [Google Scholar] [CrossRef] [PubMed]
  67. Ribas, D.; Pešková, K.; Jubany, I.; Parma, P.; Černik, M.; Benito, J.A.; Martí, V. High Reactive Nano Zero-Valent Iron Produced via Wet Milling through Abrasion by Alumina. Chem. Eng. J. 2019, 366, 235–245. [Google Scholar] [CrossRef]
  68. Stefaniuk, M.; Oleszczuk, P.; Ok, Y.S. Review on Nano Zerovalent Iron (NZVI): From Synthesis to Environmental Applications. Chem. Eng. J. 2016, 287, 618–632. [Google Scholar] [CrossRef]
  69. Li, J.; Zhang, S.; Gui, S.; Chen, G.; Wang, Y.; Wang, Z.; Zheng, X.; Meng, S.; Ruan, C.; Chen, S. Photothermal Synergistic Engineering of CeO2 and Au Co-Modified VO2 for Efficient and Selective Oxidation of Aromatic Alcohols. Appl. Surf. Sci. 2023, 611, 155616. [Google Scholar] [CrossRef]
  70. Li, W.W.; Cheng, L.; Liu, J.; Yang, S.Y.; Zan, S.T.; Zhao, G.C. Recyclable Magnetic Fe3O4@C for Methylene Blue Removal under Microwave-Induced Reaction System. Chemosphere 2023, 310, 136821. [Google Scholar] [CrossRef]
  71. Shi, X.; Li, L.; Cao, W.; Xuan, X.; Zheng, J.; Wang, C. Piezoelectric Assisted Photocatalytic Degradation of Dyes by Plasmon Ag Nanoparticles Modified Na0.5Bi0.5TiO3 Nanospheres. Mater. Lett. 2023, 335, 133827. [Google Scholar] [CrossRef]
  72. Wang, J.; Zhang, J.; Song, Y.; Xu, X.; Cai, M.; Li, P.; Yuan, W.; Xiahou, Y. Functionalized Agarose Hydrogel with in Situ Ag Nanoparticles as Highly Recyclable Heterogeneous Catalyst for Aromatic Organic Pollutants. Environ. Sci. Pollut. Res. 2023, 30, 43950–43961. [Google Scholar] [CrossRef] [PubMed]
  73. Zhang, S.; Fang, K.; Liu, X.; Qiao, X.; Wang, J. Simplified and Efficient Inkjet Printing of Cotton Fabrics Using Cationic Colored Nanoparticles. Ind. Crops Prod. 2023, 193, 116217. [Google Scholar] [CrossRef]
  74. Zahid, M.; Segar, A.S.M.; Al-Majmaie, S.; Shather, A.H.; Khan, M.F.; Alguno, A.C.; Capangpangan, R.Y.; Ismail, A. Elettaria Cardamomum Seed Extract Synthesized Silver Nanoparticles for Efficient Catalytic Reduction of Toxic Dyes. Environ. Nanotechnol. Monit. Manag. 2023, 20, 100809. [Google Scholar] [CrossRef]
  75. Pandey, K.; Saha, S. Encapsulation of Zero Valent Iron Nanoparticles in Biodegradable Amphiphilic Janus Particles for Groundwater Remediation. J. Hazard. Mater. 2023, 445, 130501. [Google Scholar] [CrossRef] [PubMed]
  76. Rajesh, G.; Kumar, P.S.; Akilandeswari, S.; Rangasamy, G.; Mandal, A.; Shankar, V.U.; Ramya, M.; Nirmala, K.; Thirumalai, K. A Synergistic Consequence of Catalyst Dosage, PH Solution and Reactive Species of Fe-Doped CdAl2O4 Nanoparticles on the Degradation of Toxic Environmental Pollutants. Chemosphere 2023, 318, 137919. [Google Scholar] [CrossRef] [PubMed]
  77. Waheed, I.F.; Hamad, M.A.; Jasim, K.A.; Gesquiere, A.J. Degradation of Methylene Blue Using a Novel Magnetic CuNiFe2O4/g-C3N4 Nanocomposite as Heterojunction Photocatalyst. Diam. Relat. Mater. 2023, 133, 109716. [Google Scholar] [CrossRef]
  78. Sun, Q.; Zhang, B.; He, Y.; Sun, L.; Hou, P.; Gan, Z.; Yu, L.; Dong, L. Design and Synthesis of Black Phosphorus Quantum Dot Sensitized Inverse Opal TiO2 Photonic Crystal with Outstanding Photocatalytic Activities. Appl. Surf. Sci. 2023, 609, 155442. [Google Scholar] [CrossRef]
  79. Zheng, C.; Song, X.; Gan, Q.; Lin, J. High-Efficiency Removal of Organic Pollutants by Visible-Light-Driven Tubular Heterogeneous Micromotors through a Photocatalytic Fenton Process. J. Colloid. Interface Sci. 2023, 630, 121–133. [Google Scholar] [CrossRef]
  80. Gadge, S.; Tamboli, A.; Shinde, M.; Fouad, H.; Terashima, C.; Chauhan, R.; Gosavi, S. Sonocatalytic Degradation of Methylene Blue Using Spindle Shaped Cerium Oxide Nanoparticles. J. Solid. State Electrochem. 2023, 27, 2005–2015. [Google Scholar] [CrossRef]
  81. Al-Arjan, W.S. Self-Assembled Nanofibrous Membranes by Electrospinning as Efficient Dye Photocatalysts for Wastewater Treatment. Polymers 2023, 15, 340. [Google Scholar] [CrossRef] [PubMed]
  82. Wang, D.; Dong, S.; Hu, H.; He, Z.; Dong, F.; Tang, J.; Lu, X.; Wang, L.; Song, S.; Ma, J. Catalytic Ozonation of Atrazine with Stable Boron-Doped Graphene Nanoparticles Derived from Waste Polyvinyl Alcohol Film: Performance and Mechanism. Chem. Eng. J. 2023, 455, 140316. [Google Scholar] [CrossRef]
  83. Liu, B.; Wang, S.; Wang, H.; Wang, Y.; Xiao, Y.; Cheng, Y. Quaternary Ammonium Groups Modified Magnetic Cyclodextrin Polymers for Highly Efficient Dye Removal and Sterilization in Water Purification. Molecules 2023, 28, 167. [Google Scholar] [CrossRef] [PubMed]
  84. Akbarzadeh, A.; Khazani, Y.; Khaloo, S.S.; Ghalkhani, M. Highly Effectual Photocatalytic Degradation of Tartrazine by Using Ag Nanoparticles Decorated on Zn-Cu-Cr Layered Double Hydroxide@ 2D Graphitic Carbon Nitride (C3N5). Environ. Sci. Pollut. Res. 2023, 30, 12903–12915. [Google Scholar] [CrossRef] [PubMed]
  85. Cheng, A.; He, Y.; Liu, X.; He, C. Honeycomb-like Biochar Framework Coupled with Fe3O4/FeS Nanoparticles as Efficient Heterogeneous Fenton Catalyst for Phenol Degradation. J. Environ. Sci. 2024, 136, 390–399. [Google Scholar] [CrossRef] [PubMed]
  86. Yang, X.; Ming, F.; Wang, J.; Xu, L. Amino Acids Modified Nanoscale Zero-Valent Iron: Density Functional Theory Calculations, Experimental Synthesis and Application in the Fenton-like Degradation of Organic Solvents. J. Environ. Sci. 2024, 135, 296–309. [Google Scholar] [CrossRef] [PubMed]
  87. Xing, J.; Qi, Z.; Dong, W.; Chen, Q.; Wu, M.; Yi, P.; Pan, B.; Xing, B. Aggregation of Biochar Nanoparticles and the Impact on Bisphenol A Sorption: Experiments and Molecular Dynamics Simulations. Sci. Total Environ. 2023, 875, 162724. [Google Scholar] [CrossRef] [PubMed]
  88. Wang, M.; Feng, L. A Carbon Based-Screen-Printed Electrode Amplified with Two-Dimensional Reduced Graphene/Fe3O4 Nanocomposite as Electroanalytical Sensor for Monitoring 4-Aminophenol in Environmental Fluids. Chemosphere 2023, 323, 138238. [Google Scholar] [CrossRef] [PubMed]
  89. Hu, L.; Cui, J.; Wang, Y.; Jia, J. An Ultrasensitive Electrochemical Biosensor for Bisphenol A Based on Aptamer-Modified MrGO@AuNPs and SsDNA-Functionalized AuNP@MBs Synergistic Amplification. Chemosphere 2023, 311, 137154. [Google Scholar] [CrossRef]
  90. Naderi, A.; Hasham Firooz, M.; Gharibzadeh, F.; Giannakis, S.; Ahmadi, M.; Rezaei Kalantary, R.; Kakavandi, B. Anchoring ZnO on Spinel Cobalt Ferrite for Highly Synergic Sono-Photo-Catalytic, Surfactant-Assisted PAH Degradation from Soil Washing Solutions. J. Environ. Manag. 2023, 326, 116584. [Google Scholar] [CrossRef]
  91. Li, S.; Feng, D.; Liu, J.; Liu, Q.; Tang, J. Surfactant-Enhanced Reduction of Soil-Adsorbed Nitrobenzene by Carbon-Coated NZVI: Enhanced Desorption and Mechanism. Sci. Total Environ. 2023, 856, 159186. [Google Scholar] [CrossRef] [PubMed]
  92. Priyadarshini, I.; Chowdhury, A.; Rao, A.; Roy, B.; Chattopadhyay, P. Assessment of Bimetallic Zn/Fe0 Nanoparticles Stabilized Tween-80 and Rhamnolipid Foams for the Remediation of Diesel Contaminated Clay Soil. J. Environ. Manag. 2023, 325, 116596. [Google Scholar] [CrossRef] [PubMed]
  93. Banu, T.; Rashid, M.H.; Tofail, S.A.M.; Haq, E.U.; Gulshan, F. Photocatalytic Degradation of Metronidazole (MNZ) Antibiotic by CuO Nanoparticles for Environmental Protection from Pharmaceutical Pollution. Surf. Interface Anal. 2023, 55, 430–436. [Google Scholar] [CrossRef]
  94. Rawashdeh, R.Y.; Qabaja, G.; Albiss, B.A. Antibacterial Activity of Multi-Metallic (Ag–Cu–Li) Nanorods with Different Metallic Combination Ratios against Staphylococcus Aureus. BMC Res. Notes 2023, 16, 23. [Google Scholar] [CrossRef] [PubMed]
  95. Chen, L.; Arshad, M.; Chuang, Y.; Hong, Y.L.; Nguyen, T.B.; Wu, C.H.; Chen, C.W.; Dong, C. Di Facile Fabrication of Efficient Tungsten Disulfide Nanoparticles for Enhanced Photocatalytic Removal of Tetracycline (TC) and Pb (II) Photoreduction. Colloids Surf. A Physicochem. Eng. Asp. 2023, 662, 131004. [Google Scholar] [CrossRef]
  96. Zhang, K.; Wang, Y.; Li, L.; Jia, L. Fabrication of Alginate-Based Nanofibers Loaded with ZnO Nanoparticles for Adsorption of Tetracyclines from Environmental Waters. Mater. Today Commun. 2023, 34, 105214. [Google Scholar] [CrossRef]
  97. Gupta, G.; Kansal, S.K.; Umar, A.; Akbar, S. Visible-Light Driven Excellent Photocatalytic Degradation of Ofloxacin Antibiotic Using BiFeO3 Nanoparticles. Chemosphere 2023, 314, 137611. [Google Scholar] [CrossRef] [PubMed]
  98. Hemmati-Eslamlu, P.; Habibi-Yangjeh, A.; Khataee, A. Anchoring Spinel NiCr2O4 Nanoparticles on Tubular G-C3N4: Efficacious p-n Heterojunction Photocatalysts for Removal of Tetracycline Hydrochloride under Visible Light. J. Alloys Compd. 2023, 932, 167571. [Google Scholar] [CrossRef]
  99. Kokulnathan, T.; Wang, T.J.; Murugesan, T.; Anthuvan, A.J.; Kumar, R.R.; Ahmed, F.; Arshi, N. Structural Growth of Zinc Oxide Nanograins on Carbon Cloth as Flexible Electrochemical Platform for Hydroxychloroquine Detection. Chemosphere 2023, 312, 137186. [Google Scholar] [CrossRef]
  100. Liu, X.; Xu, J.; Zhang, T.; Zhang, J.; Xia, D.; Du, Y.; Jiang, Y.; Lin, K. Construction of Ag Nanocluster-Modified Ag3PO4 Containing Silver Vacancies via in-Situ Reduction: With Enhancing the Photocatalytic Degradation Activity of Sulfamethoxazole. J. Colloid. Interface Sci. 2023, 629, 989–1002. [Google Scholar] [CrossRef]
  101. Lei, X.; Xu, X.; Liu, L.; Xu, L.; Wang, L.; Kuang, H.; Xu, C. Gold-Nanoparticle-Based Multiplex Immuno-Strip Biosensor for Simultaneous Determination of 83 Antibiotics. Nano Res. 2023, 16, 1259–1268. [Google Scholar] [CrossRef]
  102. Ma, S.; Luo, X.; Kong, J.; Li, X.; Cao, Z.; Wang, X.; Cai, W.; Wang, L.; Ran, G. Plasmonic Silver Loaded Hybrid Bi-Ag Nanoalloys for Highly Efficient Disinfection by Enhancing Photothermal Performance and Interface Capability. Chem. Eng. J. 2022, 450, 138016. [Google Scholar] [CrossRef]
  103. Ahmed Shehab, M.; Szőri-Dorogházi, E.; Szabó, S.; Valsesia, A.; Chauhan, T.; Koós, T.; Muránszky, G.; Szabó, T.; Hernadi, K.; Németh, Z. Virus and Bacterial Removal Ability of TiO2 Nanowire-Based Self-Supported Hybrid Membranes. Arab. J. Chem. 2023, 16, 104388. [Google Scholar] [CrossRef]
  104. Ghani, S.B.A.; Al-Azzazy, M.M.; Alhewairini, S.S.; Al-Deghairi, M.A. The Miticidal Activity of Silver Nanoparticles towards Date Palm Mite (Oligonychus Afrasiaticus (McGregor)): Efficacy, Selectivity, and Risk Assessment. Braz. J. Biol. 2022, 84, e261262. [Google Scholar] [CrossRef] [PubMed]
  105. Hao, H.; Zhu, J.; Yang, B.; Peng, L.; Lou, S. Ovalbumin-Coated Gold Nanoparticles with Interesting Colloidal Stability for Colorimetric Detection of Carbaryl in Complex Media. Food Chem. 2023, 403, 134485. [Google Scholar] [CrossRef] [PubMed]
  106. Tran, H.N.; Nguyen, N.B.; Ly, N.H.; Joo, S.W.; Vasseghian, Y. Core-Shell Au@ZIF-67-Based Pollutant Monitoring of Thiram and Carbendazim Pesticides. Environ. Pollut. 2023, 317, 120775. [Google Scholar] [CrossRef] [PubMed]
  107. Hoang, N.H.; Le Thanh, T.; Sangpueak, R.; Thepbandit, W.; Saengchan, C.; Papathoti, N.K.; Treekoon, J.; Kamkaew, A.; Phansak, P.; Buensanteai, K. The Effect of Chitosan Nanoparticle Formulations for Control of Leaf Spot Disease on Cassava. Phytoparasitica 2023, 51, 621–636. [Google Scholar] [CrossRef]
  108. Cao, Z.; Ma, X.; Zou, A.; Shi, Z.; Xiang, S.; Xu, J.; Cai, L.; Huang, J.; Sun, X. Chitin Nanocrystals Supported Copper: A New Nanomaterial with High Activity with P. Syringae Pv. Tabaci. Pest. Manag. Sci. 2023, 79, 2017–2028. [Google Scholar] [CrossRef] [PubMed]
  109. Zhang, W.; Ahari, H.; Zhang, Z.; Jafari, S.M. Role of Silica (SiO2) Nano/Micro-Particles in the Functionality of Degradable Packaging Films/Coatings and Their Application in Food Preservation. Trends Food Sci. Technol. 2023, 133, 75–86. [Google Scholar] [CrossRef]
  110. Bouzidi, I.; Sellami, B.; Boulanger, A.; Joyeux, C.; Harrath, A.H.; Albeshr, M.F.; Pacioglu, O.; Boufahja, F.; Beyrem, H.; Mougin, K. Metallic Nanoparticles Affect Uptake of Polycyclic Aromatic Hydrocarbons and Impacts in the Mediterranean Mussels Mytilus Galloprovincialis. Mar. Pollut. Bull. 2023, 188, 114641. [Google Scholar] [CrossRef]
  111. Gong, L.; Zhang, Z.; Xia, C.; Zheng, J.; Gu, Y.; He, F. A Quantitative Study of the Effects of Particle’ Properties and Environmental Conditions on the Electron Efficiency of Pd and Sulfidated Nanoscale Zero-Valent Irons. Sci. Total Environ. 2022, 853, 158469. [Google Scholar] [CrossRef] [PubMed]
  112. Kung, T.A.; Chen, P.J. Exploring Specific Biomarkers Regarding Neurobehavioral Toxicity of Lead Dioxide Nanoparticles in Medaka Fish in Different Water Matrices. Sci. Total Environ. 2023, 856, 159268. [Google Scholar] [CrossRef] [PubMed]
  113. Gao, M.; Chang, J.; Wang, Z.; Zhang, H.; Wang, T. Advances in Transport and Toxicity of Nanoparticles in Plants. J. Nanobiotechnol. 2023, 21, 75. [Google Scholar] [CrossRef] [PubMed]
  114. Zheng, N.; Sun, X.; Shi, Y.; Chen, L.; Wang, L.; Cai, H.; Han, C.; Liao, T.; Yang, C.; Zuo, Z.; et al. The Valence State of Iron-Based Nanomaterials Determines the Ferroptosis Potential in a Zebrafish Model. Sci. Total Environ. 2023, 855, 158715. [Google Scholar] [CrossRef] [PubMed]
  115. Mahjoubian, M.; Naeemi, A.S.; Moradi-Shoeili, Z.; Tyler, C.R.; Mansouri, B. Toxicity of Silver Nanoparticles in the Presence of Zinc Oxide Nanoparticles Differs for Acute and Chronic Exposures in Zebrafish. Arch. Environ. Contam. Toxicol. 2023, 84, 1–17. [Google Scholar] [CrossRef] [PubMed]
  116. Li, L.; Luo, H.; Shao, Z.; Zhou, H.; Lu, J.; Chen, J.; Huang, C.; Zhang, S.; Liu, X.; Xia, L.; et al. Converting Plastic Wastes to Naphtha for Closing the Plastic Loop. J. Am. Chem. Soc. 2023, 145, 1847–1854. [Google Scholar] [CrossRef] [PubMed]
  117. Beig, B.; Niazi, M.B.K.; Jahan, Z.; Haider, G.; Zia, M.; Shah, G.A.; Iqbal, Z.; Hayat, A. Development and Testing of Zinc Sulfate and Zinc Oxide Nanoparticle-Coated Urea Fertilizer to Improve N and Zn Use Efficiency. Front. Plant Sci. 2023, 13, 1058219. [Google Scholar] [CrossRef] [PubMed]
  118. Tao, X.; Xu, J.; Yang, K.; Lin, D. Novel Oxymagnesite/Green Rust Nanohybrids for Selective Removal and Slow Release of Phosphate in Water. Sci. Total Environ. 2023, 856, 159207. [Google Scholar] [CrossRef] [PubMed]
  119. Triolo, C.; Santangelo, S.; Petrovičovà, B.; Musolino, M.G.; Rincón, I.; Atxirika, A.; Gil, S.; Belaustegui, Y. Evaluation of the Specific Capacitance of High-Entropy Oxide-Based Electrode Materials in View of Their Use for Water Desalination via Capacitive Method. Appl. Sci. 2023, 13, 721. [Google Scholar] [CrossRef]
  120. Zhang, C.; Sun, L.; Ouyang, Y.; Xu, F.; Zou, Y.; Chu, H.; Zhang, K.; Li, B.; Pan, H. Improved Electrochemical Water Splitting by RuNi Nanoparticles Supported on RGO@mesoporous N-Doped Carbon Nanosheets. J. Alloys Compd. 2023, 937, 168334. [Google Scholar] [CrossRef]
  121. Benjak, P.; Radetić, L.; Tomaš, M.; Brnardić, I.; Radetić, B.; Špada, V.; Grčić, I. Rubber Tiles Made from Secondary Raw Materials with Immobilized Titanium Dioxide as Passive Air Protection. Processes 2023, 11, 125. [Google Scholar] [CrossRef]
  122. Wang, C.; Wang, Y.; Shi, W.D.; Yan, W.C. Electric Field Assisted Assembly of Nanoparticle Loaded Microspheres toward Industrial Applications for Organic Dye Removal. Sep. Purif. Technol. 2023, 306, 122565. [Google Scholar] [CrossRef]
  123. Zhang, G.; Yu, Y.; Tu, Y.; Liu, Y.; Huang, J.; Yin, X.; Feng, Y. Preparation of Reusable UHMWPE/TiO2 Photocatalytic Microporous Membrane Reactors for Efficient Degradation of Organic Pollutants in Water. Sep. Purif. Technol. 2023, 305, 122515. [Google Scholar] [CrossRef]
  124. Ma, J.; Yu, Z.; Shu, L.; Ke, S.; He, Q.; Zhao, Q.; Ke, Q. The Disinhibition Effect of Iron-Based Particles on Anaerobic Digestion of Florfenicol-Containing Cow Manure: Performance and Mechanism. Environ. Res. 2023, 223, 115471. [Google Scholar] [CrossRef] [PubMed]
  125. Markowicz, A.; Borymski, S.; Adamek, A.; Sułowicz, S. The Influence of ZnO Nanoparticles on Horizontal Transfer of Resistance Genes in Lab and Soil Conditions. Environ. Res. 2023, 223, 115420. [Google Scholar] [CrossRef] [PubMed]
  126. Li, X.; Lu, H.; Yang, K.; Zhu, L. Attenuation of Tetracyclines and Related Resistance Genes in Soil When Exposed to Nanoscale Zero-Valent Iron. J. Hazard. Mater. 2023, 448, 130867. [Google Scholar] [CrossRef] [PubMed]
  127. Senbill, H.; Hassan, S.M.; Eldesouky, S.E. Acaricidal and Biological Activities of Titanium Dioxide and Zinc Oxide Nanoparticles on the Two-Spotted Spider Mite, Tetranychus Urticae Koch (Acari: Tetranychidae) and Their Side Effects on the Predatory Mite, Neoseiulus Californicus (Acari: Phytoseiidae). J. Asia Pac. Entomol. 2023, 26, 102027. [Google Scholar] [CrossRef]
  128. Wang, Z.; Bi, Y.; Li, K.; Song, Z.; Pan, C.; Zhang, S.; Lan, X.; Foulkes, N.S.; Zhao, H. Nickel Oxide Nanoparticles Induce Developmental Neurotoxicity in Zebrafish by Triggering Both Apoptosis and Ferroptosis. Environ. Sci. Nano 2023, 10, 640–655. [Google Scholar] [CrossRef]
  129. Bakouei, M.; Sadeghizadeh Yazdi, J.; Ehrampoush, M.H.; Salari, M.; Madadizadeh, F. The Effect of Active Chitosan Films Containing Bacterial Cellulose Nanofiber and ZnO Nanoparticles on the Shelf Life of Loaf Bread. J. Food Qual. 2023, 2023, 1–12. [Google Scholar] [CrossRef]
  130. Alvan, Z.B.A.; Asgari, H.M.; Amanipoor, H.; Buazar, F.; Motaghed, S. Evaluation of the Effects of Zero-Valent Iron Nanoparticles in the Treatment of Soils Polluted with Refinery Effluent Hydrocarbons. Water Air Soil. Pollut. 2023, 234, 40. [Google Scholar] [CrossRef]
  131. Zhu, X.; Li, H.; Zhou, L.; Jiang, H.; Ji, M.; Chen, J. Evaluation of the Gut Microbiome Alterations in Healthy Rats after Dietary Exposure to Different Synthetic ZnO Nanoparticles. Life Sci. 2023, 312, 121250. [Google Scholar] [CrossRef] [PubMed]
  132. Aqeel, M.; Khalid, N.; Nazir, A.; Irshad, M.K.; Hakami, O.; Basahi, M.A.; Alamri, S.; Hashem, M.; Noman, A. Foliar Application of Silver Nanoparticles Mitigated Nutritional and Biochemical Perturbations in Chilli Pepper Fertigated with Domestic Wastewater. Plant Physiol. Biochem. 2023, 194, 470–479. [Google Scholar] [CrossRef] [PubMed]
  133. Liu, Y.; Li, N.; Du, C.; Wang, Y.; He, K.; Zheng, H.; Xue, Z.; Chen, Q.; Li, X. Various Hydrogen Bonds Make Different Fates of Pharmaceutical Contaminants on Oxygen-Rich Nanomaterials. Environ. Pollut. 2023, 316, 120572. [Google Scholar] [CrossRef] [PubMed]
  134. Albarano, L.; Toscanesi, M.; Trifuoggi, M.; Guida, M.; Lofrano, G.; Libralato, G. In Situ Microcosm Remediation of Polyaromatic Hydrocarbons: Influence and Effectiveness of Nano-Zero Valent Iron and Activated Carbon. Environ. Sci. Pollut. Res. 2023, 30, 3235–3251. [Google Scholar] [CrossRef] [PubMed]
  135. Nekoukhou, M.; Fallah, S.; Abbasi-Surki, A.; Pokhrel, L.R.; Rostamnejadi, A. Improved Efficacy of Foliar Application of Zinc Oxide Nanoparticles on Zinc Biofortification, Primary Productivity and Secondary Metabolite Production in Dragonhead. J. Clean. Prod. 2022, 379, 134803. [Google Scholar] [CrossRef]
  136. Yan, Z.; Liu, C.; Liu, Y.; Tan, X.; Li, X.; Shi, Y.; Ding, C. The Interaction of ZnO Nanoparticles, Cr(VI), and Microorganisms Triggers a Novel ROS Scavenging Strategy to Inhibit Microbial Cr(VI) Reduction. J. Hazard. Mater. 2023, 443, 130375. [Google Scholar] [CrossRef] [PubMed]
  137. Zeeshan, M.; Hu, Y.X.; Guo, X.H.; Sun, C.Y.; Salam, A.; Ahmad, S.; Muhammad, I.; Nasar, J.; Jahan, M.S.; Fahad, S.; et al. Physiological and Transcriptomic Study Reveal SeNPs-Mediated AsIII Stress Detoxification Mechanisms Involved Modulation of Antioxidants, Metal Transporters, and Transcription Factors in Glycine Max L. (Merr.) Roots. Environ. Pollut. 2023, 317, 120637. [Google Scholar] [CrossRef] [PubMed]
  138. Moghazy, M.A. Leidenfrost Green Synthesis Method for MoO3 and WO3 Nanorods Preparation: Characterization and Methylene Blue Adsorption Ability. BMC Chem. 2023, 17, 5. [Google Scholar] [CrossRef] [PubMed]
  139. Basumatary, S.; Kumar, K.J.; Daimari, J.; Mondal, A.; Kalita, S.; Dey, K.S.; Deka, A.K. Biosynthesis of Silver Nanoparticles Using Antidesma Acidum Leaf Extract: Its Application in Textile Organic Dye Degradation. Environ. Nanotechnol. Monit. Manag. 2023, 19, 100769. [Google Scholar] [CrossRef]
  140. Faryad, S.; Azhar, U.; Tahir, M.B.; Ali, W.; Arif, M.; Sagir, M. Spinach-Derived Boron-Doped g-C3N4/TiO2 Composites for Efficient Photo-Degradation of Methylene Blue Dye. Chemosphere 2023, 320, 138002. [Google Scholar] [CrossRef]
  141. Rather, M.Y.; Shincy, M.; Sundarapandian, S. Photocatalytic Degradation of Rhodamine-B by Phytosynthesized Gold Nanoparticles. Int. J. Environ. Sci. Technol. 2023, 20, 4073–4084. [Google Scholar] [CrossRef]
  142. Mansur, A.A.P.; Custódio, D.A.C.; Dorneles, E.M.S.; Coura, F.M.; Carvalho, I.C.; Lage, A.P.; Mansur, H.S. Nanoplexes of ZnS Quantum Dot-Poly-L-Lysine/Iron Oxide Nanoparticle-Carboxymethylcellulose for Photocatalytic Degradation of Dyes and Antibacterial Activity in Wastewater Treatment. Int. J. Biol. Macromol. 2023, 231, 123363. [Google Scholar] [CrossRef] [PubMed]
  143. Singh, K.; Nancy; Singh, G.; Singh, J. Sustainable Synthesis of Biogenic ZnO NPs for Mitigation of Emerging Pollutants and Pathogens. Environ. Res. 2023, 219, 114952. [Google Scholar] [CrossRef] [PubMed]
  144. Ashraf, I.; Singh, N.B.; Agarwal, A. Iron-Rich Coal Fly Ash-Polydopamine-Silver Nanocomposite (IRCFA-PDA-Ag NPs): Tailored Material for Remediation of Methylene Blue Dye from Aqueous Solution. Environ. Monit. Assess. 2023, 195, 322. [Google Scholar] [CrossRef] [PubMed]
  145. Munir, R.; Ali, K.; Naqvi, S.A.Z.; Maqsood, M.A.; Bashir, M.Z.; Noreen, S. Biosynthesis of Leucaena Leucocephala Leaf Mediated ZnO, CuO, MnO2, and MgO Based Nano-Adsorbents for Reactive Golden Yellow-145 (RY-145) and Direct Red-31 (DR-31) Dye Removal from Textile Wastewater to Reuse in Agricultural Purpose. Sep. Purif. Technol. 2023, 306, 122527. [Google Scholar] [CrossRef]
  146. Gautam, I.; Grady, T.; Fernando, H. Degradation of the Dye Methyl Orange Using Cow and Goat Milk Iron Nanoparticles. Green Chem. Lett. Rev. 2023, 16, 2174818. [Google Scholar] [CrossRef]
  147. Bhandari, Y.; Varma, S.; Sawant, A.; Beemagani, S.; Jaiswal, N.; Chaudhari, B.P.; Vamkudoth, K.R. Biosynthesis of Gold Nanoparticles by Penicillium Rubens and Catalytic Detoxification of Ochratoxin A and Organic Dye Pollutants. Int. Microbiol. 2023, 26, 765–780. [Google Scholar] [CrossRef] [PubMed]
  148. Shinde, S.B.; Dhengale, S.D.; Nille, O.S.; Jadhav, S.S.; Gore, A.H.; Bhosale, T.R.; Birajdar, N.B.; Kolekar, S.S.; Kolekar, G.B.; Anbhule, P.V. Template Free Synthesis of Mesoporous Carbon from Fire Cracker Waste and Designing of ZnO-Mesoporous Carbon Photocatalyst for Dye (MO) Degradation. Inorg. Chem. Commun. 2023, 147, 110242. [Google Scholar] [CrossRef]
  149. Manikanika; Chopra, L. Photo-Degradation of Dyes and Drugs Using Aloe Vera Synthesized Zinc Oxide Nanoparticles—A Review. Mater. Today Proc. 2023, 72, 1613–1617. [Google Scholar] [CrossRef]
  150. Nguyen, L.T.T.; Le, P.T.; Nguyen, T.A.; Doan, N.N.; No, K. Biochar from Cyperus Alternifolius Linn.: From a Waste of Phytoremediation Processing to Efficient Depolluting Agent. Environ. Sci. Pollut. Res. 2023, 30, 1898–1907. [Google Scholar] [CrossRef]
  151. Aryafar, A.; Ekrami-Kakhki, M.S.; Naeimi, A. Enhanced Electrocatalytic Activity of Pt-SnO2 Nanoparticles Supported on Natural Bentonite-Functionalized Reduced Graphene Oxide-Extracted Chitosan from Shrimp Wastes for Methanol Electro-Oxidation. Sci. Rep. 2023, 13, 3597. [Google Scholar] [CrossRef] [PubMed]
  152. Yuan, X.; Yu, S.; Xue, N.; Li, T.; Sun, M.; Zhang, L. Activation of Iron Based Persulfate Heterogeneous Nano Catalyst Using Plant Extract for Removal of Tetrabromobisphenol A from Soil. J. Environ. Chem. Eng. 2023, 11, 109493. [Google Scholar] [CrossRef]
  153. Das, S.; Panicker, N.J.; Rather, M.A.; Mandal, M.; Sahu, P.P. Green Synthesis of Cerium Oxide Nanoparticles Using Dillenia Indica Aqueous Extract and Its Anti-Oxidant Activity. Bull. Mater. Sci. 2023, 46, 3. [Google Scholar] [CrossRef]
  154. Doan, V.D.; Le, V.T.; Tran, D.L.; Nguyen, T.L.H.; Nguyen, D.C.; Nguyen, A.T.; Le, V.T. Catalytic Reduction of Nitrophenols Using Gnetum Montanum Extract Capped Silver Nanoparticles. Mol. Catal. 2023, 534, 112804. [Google Scholar] [CrossRef]
  155. Ganesan, S.; Nijohn, J.M.; lakshmisankar, S.; Venkatesan, S.P.; Hemanandh, J.; Muniappan, P. Environmental Analysis of One-Step Hydrothermal Green Synthesized MgO Nanoparticles on VCR Diesel Engine Performance Using Jatropha Oil and Lemongrass Oil. Nanotechnol. Environ. Eng. 2023, 8, 599–613. [Google Scholar] [CrossRef]
  156. Wang, Y.; Ding, Y.; Tan, Y.; Fu, L.; Qing, W. Preparation of Transition Metal Ions (Fe2+, Co2+ and Ni2+) Doped Carbon Nanoparticles from Biowaste for Cystine and Cr(VI) Detection and Fluorescence Ink. Inorg. Chem. Commun. 2023, 147, 110220. [Google Scholar] [CrossRef]
  157. Wali, S.; Zahra, M.; Okla, M.K.; Wahidah, H.A.; Tauseef, I.; Haleem, K.S.; Farid, A.; Maryam, A.; Abdelgawad, H.; Adetunji, C.O.; et al. Brassica Oleracea L. (Acephala Group) Based Zinc Oxide Nanoparticles and Their Efficacy as Antibacterial Agent. Braz. J. Biol. 2024, 84, e259351. [Google Scholar] [CrossRef]
  158. Ghasemi, S.; Harighi, B.; Ashengroph, M. Biosynthesis of Silver Nanoparticles Using Pseudomonas Canadensis, and Its Antivirulence Effects against Pseudomonas Tolaasii, Mushroom Brown Blotch Agent. Sci. Rep. 2023, 13, 3668. [Google Scholar] [CrossRef] [PubMed]
  159. Hu, C.; Zhu, W.; Lu, Y.; Ren, Y.; Gu, J.; Song, Y.; He, J. Alpinia Officinarum Mediated Copper Oxide Nanoparticles: Synthesis and Its Antifungal Activity against Colletotrichum Gloeosporioides. Environ. Sci. Pollut. Res. 2023, 30, 28818–28829. [Google Scholar] [CrossRef]
  160. Nikishina, M.B.; Ivanova, E.V.; Tretyakova, A.V.; Mukhtorov, L.G.; Atroshchenko, Y.M. Study of Biological Activity Colloidal Solutions of Iron Synthesized on the Basis of Aqueous Cuff Extract. Sib. J. Life Sci. Agric. 2022, 14, 388–403. [Google Scholar] [CrossRef]
  161. Palanisamy, D.S.; Gounder, B.S.; Selvaraj, K.; Kandhasamy, S.; Alqahtani, T.; Alqahtani, A.; Chidambaram, K.; Arunachalam, K.; Alkahtani, A.M.; Chandramoorthy, H.C.; et al. Synergistic Antibacterial and Mosquitocidal Effect of Passiflora Foetida Synthesized Silver Nanoparticles. Braz. J. Biol. 2024, 84, 263391. [Google Scholar] [CrossRef] [PubMed]
  162. Nawab, R.; Ali, M.; Haroon, U.; Kamal, A.; Akbar, M.; Anwar, F.; Ahmed, J.; Chaudhary, H.J.; Iqbal, A.; Hashem, M.; et al. Calotropis procera (L.) Mediated Synthesis of AgNPs and Their Application to Control Leaf Spot of Hibiscus rosa-sinensis (L.). Braz. J. Biol. 2024, 84, 261123. [Google Scholar] [CrossRef] [PubMed]
  163. Sultana, M.J.; Nibir, A.I.S.; Ahmed, F.R.S. Biosensing and Anti-Inflammatory Effects of Silver, Copper and Iron Nanoparticles from the Leaf Extract of Catharanthus Roseus. Beni Suef Univ. J. Basic Appl. Sci. 2023, 12, 26. [Google Scholar] [CrossRef]
  164. Agustin, M.B.; Lehtonen, M.; Kemell, M.; Lahtinen, P.; Oliaei, E.; Mikkonen, K.S. Lignin Nanoparticle-Decorated Nanocellulose Cryogels as Adsorbents for Pharmaceutical Pollutants. J. Environ. Manag. 2023, 330, 117210. [Google Scholar] [CrossRef] [PubMed]
  165. Hamdan, D.I.; Tawfeek, N.; El-Shiekh, R.A.; Khalil, H.M.A.; Mahmoud, M.Y.; Bakr, A.F.; Zaafar, D.; Farrag, N.; Wink, M.; El-Shazly, A.M. Salix Subserrata Bark Extract-Loaded Chitosan Nanoparticles Attenuate Neurotoxicity Induced by Sodium Arsenate in Rats in Relation with HPLC–PDA-ESI–MS/MS Profile. AAPS PharmSciTech 2023, 24, 15. [Google Scholar] [CrossRef] [PubMed]
  166. Deshmukh, P.; Sar, S.K.; Jindal, M.K. Plant Mediated Magnetic Nano Composite as Promising Scavenger’s Radionuclides for the Efficient Remediation in Aqueous Medium. Chemosphere 2023, 312, 137246. [Google Scholar] [CrossRef] [PubMed]
  167. Karthik Chandan, A.; Neha Mallika, G.; Bala Narsaiah, T. A Green Approach to Arsenic Removal Using ZnO Nanoparticles Synthesized from Acacia Catechu Leaf Extract. Mater. Today Proc. 2023, 72, 110–119. [Google Scholar] [CrossRef]
  168. Gong, Y.; Wang, Y.; Lin, N.; Wang, R.; Wang, M.; Zhang, X. Iron-Based Materials for Simultaneous Removal of Heavy Metal(Loid)s and Emerging Organic Contaminants from the Aquatic Environment: Recent Advances and Perspectives. Environ. Pollut. 2022, 299, 118871. [Google Scholar] [CrossRef] [PubMed]
  169. Papolu, P.; Bhogi, A. Green Synthesis of Various Metal Oxide Nanoparticles for the Environmental Remediation—An Overview. Mater. Today Proc. 2023, 92, 924–927. [Google Scholar] [CrossRef]
  170. Kobayashi, Y.; Yamamoto, K.; Shoji, R. A CaH2-Assisted Reduction Method to Prepare Nanoscale Zero-Valent Iron (NZVI) from Fe2O3 for Water Remediation Application. Minerals 2023, 13, 1385. [Google Scholar] [CrossRef]
  171. Couto, D.; Freitas, M.; Carvalho, F.; Fernandes, E. Iron Oxide Nanoparticles: An Insight into Their Biomedical Applications. Curr. Med. Chem. 2015, 22, 1808–1828. [Google Scholar] [CrossRef] [PubMed]
  172. Sangaiya, P.; Jayaprakash, R. A Review on Iron Oxide Nanoparticles and Their Biomedical Applications. J. Supercond. Nov. Magn. 2018, 31, 3397–3413. [Google Scholar] [CrossRef]
  173. Gautam, S.; Bansal, D.; Bhatnagar, D.; Sharma, C.; Goyal, N. Synthesis of Iron-Based Nanoparticles by Chemical Methods and Their Biomedical Applications. In Oxides for Medical Applications; Woodhead Publishing: Sawston, UK, 2023; ISBN 978-0-32390-538-1. [Google Scholar] [CrossRef]
  174. Manohar, A.K.; Yang, C.; Malkhandi, S.; Prakash, G.K.S.; Narayanan, S.R. Enhancing the Performance of the Rechargeable Iron Electrode in Alkaline Batteries with Bismuth Oxide and Iron Sulfide Additives. J. Electrochem. Soc. 2013, 160, A2078–A2084. [Google Scholar] [CrossRef]
  175. Hayashi, K.; Maeda, Y.; Suzuki, T.; Sakamoto, H.; Kugimiya, T.; Tan, W.K.; Kawamura, G.; Muto, H.; Matsuda, A. Development of Iron-Based Rechargeable Batteries with Sintered Porous Iron Electrodes. ECS Trans. 2016, 75, 111–116. [Google Scholar] [CrossRef]
  176. Yang, C.; Manohar, A.K.; Narayanan, S.R. A High-Performance Sintered Iron Electrode for Rechargeable Alkaline Batteries to Enable Large-Scale Energy Storage. J. Electrochem. Soc. 2017, 164, A418–A429. [Google Scholar] [CrossRef]
  177. Wei, D.; Darcel, C. Iron Catalysis in Reduction and Hydrometalation Reactions. Chem. Rev. 2019, 119, 2550–2610. [Google Scholar] [CrossRef] [PubMed]
  178. Rydel-Ciszek, K.; Pacześniak, T.; Zaborniak, I.; Błoniarz, P.; Surmacz, K.; Sobkowiak, A.; Chmielarz, P. Iron-Based Catalytically Active Complexes in Preparation of Functional Materials. Processes 2020, 8, 1683. [Google Scholar] [CrossRef]
  179. Palomo, J.M. Design of Iron Nanostructured Catalysts. In Iron Catalysis; World Scientific: Singapore, 2021; Volume 19. [Google Scholar] [CrossRef]
  180. Franken, T.; Heel, A. Are Fe Based Catalysts an Upcoming Alternative to Ni in CO2 Methanation at Elevated Pressure? J. CO2 Util. 2020, 39, 101175. [Google Scholar] [CrossRef]
  181. Phuong, D.H.L.; Alsaiari, M.; Pham, C.Q.; Hieu, N.H.; T․ Pham, T.-P.; Rajamohan, N.; Pham, D.D.; Vo, D.-V.N.; Trinh, T.H.; Setiabudi, H.D.; et al. Carbon Dioxide Reforming of Methane over Modified Iron-Cobalt Alumina Catalyst: Role of Promoter. J. Taiwan. Inst. Chem. Eng. 2024, 155, 105253. [Google Scholar] [CrossRef]
  182. Guo, N.; Zhu, S. Iron-Catalyzed Hydrogenation Reactions. Chin. J. Org. Chem. 2015, 35, 1383–1398. [Google Scholar] [CrossRef]
  183. Cheng, Y.; Pham, H.; Huo, J.; Johnson, R.; Datye, A.K.; Shanks, B. High Activity Pd-Fe Bimetallic Catalysts for Aqueous Phase Hydrogenations. Mol. Catal. 2019, 477, 110546. [Google Scholar] [CrossRef]
  184. Amiri, R.; Bourezgui, A.; Djeridi, W.; Dappozze, F.; Houas, A.; Guillard, C.; Elsellami, L. Surface Modification of TiO2 with a Less Expensive Metal (Iron) to Exploit Solar Energy in Photocatalysis: An Ecological and Economical Solution. Int. J. Hydrog. Energy 2024, 51, 638–647. [Google Scholar] [CrossRef]
  185. Zhang, M.; Zhang, K.; Zhou, R.; Wang, J. Hydroxylamine Enhanced Activation of Peroxymonosulfate by Fe(III)/Cu(II) Bimetallic for High-Efficiency Degradation of AO7. Water Sci. Technol. 2022, 85, 2038–2050. [Google Scholar] [CrossRef]
  186. Zhang, L.; Huo, S.; Li, J.; Gao, M. Preparation of CuFeO2 as Heterogeneous Fenton-like Catalyst and Its Degradation Performance on Ofloxacin in the Water. Chin. J. Environ. Eng. 2022, 16, 2144–2155. [Google Scholar] [CrossRef]
  187. Hou, Q.; Qin, L.; Peng, X.; Zhou, C.; Li, X.; Zhang, J.; Gao, L. Insight to an Efficient and Magnetic α-Fe2O3/γ-Fe2O3/Cu2O Hybrid Catalysis for Peroxymonosulfate: Preparation, Performance, and Mechanism. Res. Chem. Intermed. 2022, 48, 3753–3772. [Google Scholar] [CrossRef]
  188. Mendoza-Carrasco, R.; Cuerda-Correa, E.M.; Alexandre-Franco, M.F.; Fernández-González, C.; Gómez-Serrano, V. Preparation of High-Quality Activated Carbon from Polyethyleneterephthalate (PET) Bottle Waste. Its Use in the Removal of Pollutants in Aqueous Solution. J. Environ. Manag. 2016, 181, 522–535. [Google Scholar] [CrossRef] [PubMed]
  189. Tang, C.; Wang, X.; Zhang, Y.; Liu, N.; Hu, X. Corrosion Behaviors and Kinetics of Nanoscale Zero-Valent Iron in Water: A Review. J. Environ. Sci. 2024, 135, 391–406. [Google Scholar] [CrossRef] [PubMed]
  190. Gomes, H.I.; Dias-Ferreira, C.; Ribeiro, A.B.; Pamukcu, S. Influence of Electrolyte and Voltage on the Direct Current Enhanced Transport of Iron Nanoparticles in Clay. Chemosphere 2014, 99, 171–179. [Google Scholar] [CrossRef] [PubMed]
  191. Grieger, K.D.; Fjordbøge, A.; Hartmann, N.B.; Eriksson, E.; Bjerg, P.L.; Baun, A. Environmental Benefits and Risks of Zero-Valent Iron Nanoparticles (NZVI) for in Situ Remediation: Risk Mitigation or Trade-Off? J. Contam. Hydrol. 2010, 118, 165–183. [Google Scholar] [CrossRef] [PubMed]
  192. Kahraman, B.F. Recent Advances in Soil Remediation Using Nano Zero-Valent Iron (NZVI) Activated Peroxydisulfate (PDS) and Peroxymonosulfate (PMS). In Advances in Environmental Research; Nova Publishers: Hauppauge, NY, USA, 2023; Volume 94, ISBN 979-8-88697-665-6. [Google Scholar]
  193. Li, Q.; Chen, Z.; Wang, H.; Yang, H.; Wen, T.; Wang, S.; Hu, B.; Wang, X. Removal of Organic Compounds by Nanoscale Zero-Valent Iron and Its Composites. Sci. Total Environ. 2021, 792, 148546. [Google Scholar] [CrossRef]
  194. Ma, Z.; Cao, H.; Lv, F.; Yang, Y.; Chen, C.; Yang, T.; Zheng, H.; Wu, D. Preparation of NZVI Embedded Modified Mesoporous Carbon for Catalytic Persulfate to Degradation of Reactive Black 5. Front. Environ. Sci. Eng. 2021, 15, 98. [Google Scholar] [CrossRef]
  195. Soubh, A.M.; Abdoli, M.A.; Ahmad, L.A. Optimizing the Removal of Methylene Blue from Aqueous Solutions Using Persulfate Activated with Nanoscale Zero Valent Iron (nZVI) Supported by Reduced Expanded Graphene Oxide (REGO). Environ. Health Eng. Manag. 2021, 8, 15–24. [Google Scholar] [CrossRef]
  196. Li, Y.-T.; Liu, X.-Y.; Li, X.; Liu, H.; Du, W.-Y.; Chen, J.-L. Oxidative Degradation of Rhodamine B Solution with NZVI Persulfate Activation. Environ. Eng. Res. 2024, 29, 230405. [Google Scholar] [CrossRef]
  197. Mu, Y.; Jia, F.; Ai, Z.; Zhang, L. Molecular Oxygen Activation with Nano Zero-Valent Iron for Aerobic Degradation of Organic Contaminants and the Performance Enhancement. Acta Chim. Sin. 2017, 75, 538–543. [Google Scholar] [CrossRef]
  198. Dada, A.O.; Adekola, F.A.; Odebunmi, E.O.; Dada, F.E.; Bello, O.S.; Ogunlaja, A.S. Bottom-up Approach Synthesis of Core-Shell Nanoscale Zerovalent Iron (CS-NZVI): Physicochemical and Spectroscopic Characterization with Cu(II) Ions Adsorption Application. MethodsX 2020, 7, 100976. [Google Scholar] [CrossRef] [PubMed]
  199. Liu, A.; Zhang, W.-X. Fine Structural Features of Nanoscale Zero-Valent Iron Characterized by Spherical Aberration Corrected Scanning Transmission Electron Microscopy (CS-STEM). Analyst 2014, 139, 4512–4518. [Google Scholar] [CrossRef] [PubMed]
  200. Mathew, N.K.; Rohith, V.K.; Theertharaman, G.; Navaneethan, M.; Balakumar, S. Probing the Influence of Liquid Nitrogen Assisted Chemical Reduction on the Nature of Passivation Layer, Magnetic Properties, and Cr (VI) Remediation Performance of Nanoscale Zero Valent Iron. J. Environ. Chem. Eng. 2023, 11, 109096. [Google Scholar] [CrossRef]
  201. Wang, Q.; Luo, W.; Chen, X.; Fan, J.; Jiang, W.; Wang, L.; Jiang, W.; Zhang, W.-X.; Yang, J. Porous-Carbon-Confined Formation of Monodisperse Iron Nanoparticle Yolks toward Versatile Nanoreactors for Metal Extraction. Chem. Eur. J. 2018, 24, 15663–15668. [Google Scholar] [CrossRef] [PubMed]
  202. Li, X.Q.; Elliott, D.W.; Zhang, W.X. Zero-Valent Iron Nanoparticles for Abatement of Environmental Pollutants: Materials and Engineering Aspects. Crit. Rev. Solid. State Mater. Sci. 2006, 31, 111–122. [Google Scholar] [CrossRef]
  203. Yan, W.; Herzing, A.A.; Kiely, C.J.; Zhang, W.-X. Nanoscale Zero-Valent Iron (NZVI): Aspects of the Core-Shell Structure and Reactions with Inorganic Species in Water. J. Contam. Hydrol. 2010, 118, 96–104. [Google Scholar] [CrossRef]
  204. Huber, D.L. Synthesis, Properties, and Applications of Iron Nanoparticles. Small 2005, 1, 482–501. [Google Scholar] [CrossRef] [PubMed]
  205. Teja, A.S.; Koh, P.-Y. Synthesis, Properties, and Applications of Magnetic Iron Oxide Nanoparticles. Prog. Cryst. Growth Charact. Mater. 2009, 55, 22–45. [Google Scholar] [CrossRef]
  206. Nemati, Z.; Khurshid, H.; Alonso, J.; Phan, M.H.; Mukherjee, P.; Srikanth, H. From Core/Shell to Hollow Fe/γ-Fe2O3 Nanoparticles: Evolution of the Magnetic Behavior. Nanotechnology 2015, 26, 405705. [Google Scholar] [CrossRef] [PubMed]
  207. 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] [PubMed]
  208. Ansari, A.; Siddiqui, V.U.; Akram, M.K.; Siddiqi, W.A.; Khan, A.; Al-Romaizan, A.N.; Hussein, M.A.; Puttegowda, M. Synthesis of Atmospherically Stable Zero-Valent Iron Nanoparticles (nZVI) for the Efficient Catalytic Treatment of High-Strength Domestic Wastewater. Catalysts 2022, 12, 26. [Google Scholar] [CrossRef]
  209. Carpenter, E.E.; Calvin, S.; Stroud, R.M.; Harris, V.G. Passivated Iron as Core-Shell Nanoparticles. Chem. Mater. 2003, 15, 3245–3246. [Google Scholar] [CrossRef]
  210. Barreto-Rodrigues, M.; Silveira, J.; Zazo, J.A.; Rodriguez, J.J. Synthesis, Characterization and Application of Nanoscale Zero-Valent Iron in the Degradation of the Azo Dye Disperse Red 1. J. Environ. Chem. Eng. 2017, 5, 628–634. [Google Scholar] [CrossRef]
  211. Ken, D.S.; Sinha, A. Recent Developments in Surface Modification of Nano Zero-Valent Iron (NZVI): Remediation, Toxicity and Environmental Impacts. Environ. Nanotechnol. Monit. Manag. 2020, 14, 100344. [Google Scholar] [CrossRef]
  212. Idris, D.S.; Roy, A. Synthesis of Bimetallic Nanoparticles and Applications—An Updated Review. Crystals 2023, 13, 637. [Google Scholar] [CrossRef]
  213. He, Y.; Lin, H.; Dong, Y.; Li, B.; Wang, L.; Chu, S.; Luo, M.; Liu, J. Zeolite Supported Fe/Ni Bimetallic Nanoparticles for Simultaneous Removal of Nitrate and Phosphate: Synergistic Effect and Mechanism. Chem. Eng. J. 2018, 347, 669–681. [Google Scholar] [CrossRef]
  214. Dongsheng, Z.; Wenqiang, G.; Guozhang, C.; Shuai, L.; Weizhou, J.; Youzhi, L. Removal of Heavy Metal Lead(II) Using Nanoscale Zero-Valent Iron with Different Preservation Methods. Adv. Powder Technol. 2019, 30, 581–589. [Google Scholar] [CrossRef]
  215. Jiang, G.; Lan, M.; Zhang, Z.; Lv, X.; Lou, Z.; Xu, X.; Dong, F.; Zhang, S. Identification of Active Hydrogen Species on Palladium Nanoparticles for an Enhanced Electrocatalytic Hydrodechlorination of 2,4-Dichlorophenol in Water. Environ. Sci. Technol. 2017, 51, 7599–7605. [Google Scholar] [CrossRef] [PubMed]
  216. Izadi, N.; Sangani, M.M.M.; Yavari, M.A.; Baghdadi, M. Optimization of a Simple Continuous System for the Preparation of Zero-Valent Iron Nanoparticles Coated with Flaxseed Gum: Effect of Groundwater Quality on the Aggregation. Environ. Technol. Innov. 2023, 30, 103119. [Google Scholar] [CrossRef]
  217. Parimala, L.; Santhanalakshmi, J. Studies on the Iron Nanoparticles Catalyzed Reduction of Substituted Aromatic Ketones to Alcohols. J. Nanopart. 2014, 2014, 156868. [Google Scholar] [CrossRef]
  218. Nunez Garcia, A.; Boparai, H.K.; de Boer, C.V.; Chowdhury, A.I.A.; Kocur, C.M.D.; Austrins, L.M.; Herrera, J.; O’Carroll, D.M. Fate and Transport of Sulfidated Nano Zerovalent Iron (S-nZVI): A Field Study. Water Res. 2020, 170, 115319. [Google Scholar] [CrossRef] [PubMed]
  219. Kania, G.; Sternak, M.; Jasztal, A.; Chlopicki, S.; Błażejczyk, A.; Nasulewicz-Goldeman, A.; Wietrzyk, J.; Jasiński, K.; Skórka, T.; Zapotoczny, S.; et al. Uptake and Bioreactivity of Charged Chitosan-Coated Superparamagnetic Nanoparticles as Promising Contrast Agents for Magnetic Resonance Imaging. Nanomedicine 2018, 14, 131–140. [Google Scholar] [CrossRef] [PubMed]
  220. Yang, C.; Ge, C.; Li, X.; Li, L.; Wang, B.; Lin, A.; Yang, W. Does Soluble Starch Improve the Removal of Cr(VI) by nZVI Loaded on Biochar? Ecotoxicol. Environ. Saf. 2021, 208, 111552. [Google Scholar] [CrossRef] [PubMed]
  221. Wei, L.; Tielong, L.; Fangchun, L.; Jinjuan, Z. Synthesis of Guar Gum Stabilized Nanoscale Zero-Valent Iron for Cr (VI) Removal in Water. IOP Conf. Ser. Earth Environ. Sci. 2021, 647, 012101. [Google Scholar] [CrossRef]
  222. Xu, Y.; Zhao, D. Reductive Immobilization of Chromate in Water and Soil Using Stabilized Iron Nanoparticles. Water. Res. 2007, 41, 2101–2108. [Google Scholar] [CrossRef]
  223. Berge, N.D.; Ramsburg, C.A. Iron-Mediated Trichloroethene Reduction within Nonaqueous Phase Liquid. J. Contam. Hydrol. 2010, 118, 105–116. [Google Scholar] [CrossRef]
  224. Quinn, J.; Geiger, C.; Clausen, C.; Brooks, K.; Coon, C.; O’Hara, S.; Krug, T.; Major, D.; Yoon, W.S.; Gavaskar, A.; et al. Field Demonstration of DNAPL Dehalogenation Using Emulsified Zero-Valent Iron. Environ. Sci. Technol. 2005, 39, 1309–1318. [Google Scholar] [CrossRef] [PubMed]
  225. Zhang, W.; Qian, L.; Chen, Y.; Ouyang, D.; Han, L.; Shang, X.; Li, J.; Gu, M.; Chen, M. Nanoscale Zero-Valent Iron Supported by Attapulgite Produced at Different Acid Modification: Synthesis Mechanism and the Role of Silicon on Cr(VI) Removal. Chemosphere 2021, 267, 129183. [Google Scholar] [CrossRef] [PubMed]
  226. Zarime, N.‘A.; Solemon, B.; Wan Yaacob, W.Z.; Jamil, H.; Che Omar, R.; Oyekanmi, A.A. Effectiveness of Artificially Synthesized Granitic Residual Soil-Supported Nano Zero-Valent Iron (Gr-ZVI) as Effective Heavy Metal Contaminant Adsorbent. Inorganics 2023, 11, 131. [Google Scholar] [CrossRef]
  227. Tajabadi, M.; Rahmani, I.; Mirkazemi, S.M.; Goran Orimi, H. Insights into the Synthesis Optimization of Fe@SiO2 Core-Shell Nanostructure as a Highly Efficient Nano-Heater for Magnetic Hyperthermia Treatment. Adv. Powder Technol. 2022, 33, 103366. [Google Scholar] [CrossRef]
  228. Xu, Y.; Liu, Z.; Ma, K.; Qin, Q. Facile Synthesis of High Iron Content Activated Carbon-Supported Nanoscale Zero-Valent Iron for Enhanced Cr(VI) Removal in Aqueous Solution. Chemosphere 2022, 291, 132709. [Google Scholar] [CrossRef] [PubMed]
  229. Falyouna, O.; Eljamal, O.; Maamoun, I.; Tahara, A.; Sugihara, Y. Magnetic Zeolite Synthesis for Efficient Removal of Cesium in a Lab-Scale Continuous Treatment System. J. Colloid Interface Sci. 2020, 571, 66–79. [Google Scholar] [CrossRef] [PubMed]
  230. Leovac Maćerak, A.; Kulić Mandić, A.; Pešić, V.; Tomašević Pilipović, D.; Bečelić-Tomin, M.; Kerkez, D. “Green” nZVI-Biochar as Fenton Catalyst: Perspective of Closing-the-Loop in Wastewater Treatment. Molecules 2023, 28, 1425. [Google Scholar] [CrossRef] [PubMed]
  231. Liu, H.; Xu, J.; Li, Y.; Li, Y. Aggregate Nanostructures of Organic Molecular Materials. Acc. Chem. Res. 2010, 43, 1496–1508. [Google Scholar] [CrossRef]
  232. Kumari, N.; Behera, M.; Singh, R. Facile Synthesis of Biopolymer Decorated Magnetic Coreshells for Enhanced Removal of Xenobiotic Azo Dyes through Experimental Modelling. Food Chem. Toxicol. 2023, 171, 113518. [Google Scholar] [CrossRef]
  233. Antony, J.; Meera, V.; Raphael, V.P.; Vinod, P. Facile Encapsulation of Nano Zero-Valent Iron with Calcium Carbonate: Synthesis, Characterization and Application for Iron Remediation. J. Environ. Health Sci. Eng. 2022, 20, 915–930. [Google Scholar] [CrossRef]
  234. Hierrezuelo, J.; Sadeghpour, A.; Szilagyi, I.; Vaccaro, A.; Borkovec, M. Electrostatic Stabilization of Charged Colloidal Particles with Adsorbed Polyelectrolytes of Opposite Charge. Langmuir 2010, 26, 15109–15111. [Google Scholar] [CrossRef] [PubMed]
  235. Suk, J.S.; Xu, Q.; Kim, N.; Hanes, J.; Ensign, L.M. PEGylation as a Strategy for Improving Nanoparticle-Based Drug and Gene Delivery. Adv. Drug Deliv. Rev. 2016, 99, 28–51. [Google Scholar] [CrossRef] [PubMed]
  236. Moura, C.C.; Salazar-Bryam, A.M.; Piazza, R.D.; Carvalho dos Santos, C.; Jafelicci, M.; Marques, R.F.C.; Contiero, J. Rhamnolipids as Green Stabilizers of nZVI and Application in the Removal of Nitrate From Simulated Groundwater. Front. Bioeng. Biotechnol. 2022, 10, 794460. [Google Scholar] [CrossRef] [PubMed]
  237. Izadi, N.; Ali, B.H.; Shahin, M.S.; Baghdadi, M. The Removal of Cr(VI) from Aqueous and Saturated Porous Media by Nanoscale Zero-Valent Iron Stabilized with Flaxseed Gum Extract: Synthesis by Continuous Flow Injection Method. Korean J. Chem. Eng. 2022, 39, 2217–2228. [Google Scholar] [CrossRef]
  238. Shaikh, A.A.; Bera, S.; Paul, S.; Mondal, S.; Saha, A.; Roy, S. Synthesis and Characterization of Inorganic Nanoparticles Luminophores for Environmental Remediation. 4open 2022, 5, 19. [Google Scholar] [CrossRef]
  239. Bhawna; Acharya, A.D.; Kaur, S. Rapid Reductive Degradation of Dye Contaminated Water by Using a Core-Shell Nano Zerovalent Iron (NZVI). J. Indian Chem. Soc. 2022, 99, 100598. [Google Scholar] [CrossRef]
  240. Moreno-Bárcenas, A.; Arizpe-Zapata, J.A.; Rivera Haro, J.A.; Sepúlveda, P.; Garcia-Garcia, A. Jute Fibers Synergy with nZVI/GO: Superficial Properties Enhancement for Arsenic Removal in Water with Possible Application in Dynamic Flow Filtration Systems. Nanomaterials 2022, 12, 3974. [Google Scholar] [CrossRef]
  241. Selvan, B.K.; Thiyagarajan, K.; Das, S.; Jaya, N.; Jabasingh, S.A.; Saravanan, P.; Rajasimman, M.; Vasseghian, Y. Synthesis and Characterization of Nano Zerovalent Iron-Kaolin Clay (NZVI-Kaol) Composite Polyethersulfone (PES) Membrane for the Efficacious As2O3 Removal from Potable Water Samples. Chemosphere 2022, 288, 132405. [Google Scholar] [CrossRef] [PubMed]
  242. Song, M.; Hu, X.; Gu, T.; Zhang, W.X.; Deng, Z. Nanocelluloses Affixed Nanoscale Zero-Valent Iron (NZVI) for Nickel Removal: Synthesis, Characterization and Mechanisms. J. Environ. Chem. Eng. 2022, 10, 107466. [Google Scholar] [CrossRef]
  243. Xia, J.; Shen, Y.; Zhang, H.; Hu, X.; Mian, M.M.; Zhang, W.H. Synthesis of Magnetic NZVI@biochar Catalyst from Acid Precipitated Black Liquor and Fenton Sludge and Its Application for Fenton-like Removal of Rhodamine B Dye. Ind. Crops Prod. 2022, 187, 115449. [Google Scholar] [CrossRef]
  244. Liu, Y.; Zhang, X.; Zhou, Y.; Ma, H.; Cheng, X.; Wei, L.; Hou, Z. MFO@NZVI/Hydrogel for Sulfasalazine Degradation: Performance, Mechanism and Degradation Pathway. Sep. Purif. Technol. 2022, 282, 120054. [Google Scholar] [CrossRef]
  245. Shukla, F.; Kikani, T.; Khan, A.; Thakore, S. α-Hydroxy Acids Modified β-Cyclodextrin Capped Iron Nanocatalyst for Rapid Reduction of Nitroaromatics: A Sonochemical Approach. Int. J. Biol. Macromol. 2022, 209, 1504–1515. [Google Scholar] [CrossRef] [PubMed]
  246. Akoto, J.D.; Chai, F.; Repo, E.; Yang, Z.; Wang, D.; Zhao, F.; Liao, Q.; Chai, L. Polyethyleneimine Stabilized Nanoscale Zero-Valent Iron-Magnetite (Fe3O4@nZVI-PEI) for the Enhanced Removal of Arsenic from Acidic Aqueous Solution: Performance and Mechanisms. J. Environ. Chem. Eng. 2022, 10, 108589. [Google Scholar] [CrossRef]
  247. Cheng, F.; Hou, J.; Wu, J.; Xu, Y.; Miao, L.; Xia, J.; Huo, Z. Kinetics, Products and Pathways for the Removal of Pentachlorophenol (PCP) by Sulfidated Nanoscale Zero-Valent Iron (S-NZVI). Environ. Sci. 2022, 8, 2567–2579. [Google Scholar] [CrossRef]
  248. Abd El-Monaem, E.M.; Omer, A.M.; El-Subruiti, G.M.; Mohy-Eldin, M.S.; Eltaweil, A.S. Zero-Valent Iron Supported-Lemon Derived Biochar for Ultra-Fast Adsorption of Methylene Blue. Biomass Convers. Biorefin. 2022, 14, 1697–1709. [Google Scholar] [CrossRef]
  249. Okonji, S.O.; Achari, G.; Pernitsky, D. Removal of Organoselenium from Aqueous Solution by Nanoscale Zerovalent Iron Supported on Granular Activated Carbon. Water 2022, 14, 987. [Google Scholar] [CrossRef]
  250. Idham, M.F.; Falyouna, O.; Eljamal, R.; Maamoun, I.; Eljamal, O. Chloramphenicol Removal from Water by Various Precursors to Enhance Graphene Oxide–Iron Nanocomposites. J. Water Process Eng. 2022, 50, 103289. [Google Scholar] [CrossRef]
  251. Xie, Y.; Lu, G.; Tao, X.; Wen, Z.; Dang, Z. A Collaborative Strategy for Elevated Reduction and Immobilization of Cr(VI) Using Nano Zero Valent Iron Assisted by Schwertmannite: Removal Performance and Mechanism. J. Hazard. Mater. 2022, 422, 126952. [Google Scholar] [CrossRef] [PubMed]
  252. Orbuleţ, O.D.; Dăncilă, A.M.; Căprărescu, S.; Modrogan, C.; Purcar, V. Nitrates Removal from Simulated Groundwater Using Nano Zerovalent Iron Supported by Polystyrenic Gel. Polymers 2023, 15, 61. [Google Scholar] [CrossRef]
  253. Fan, X.; Ma, L.; Liu, S.; Xie, Y.; Lu, S.; Tan, Z.; Ji, J.; Fu, M.L.; Yuan, B.; Hu, Y.-B. Facile Synthesis of Lattice-Defective and Recyclable Zirconium Hydroxide Coated Nanoscale Zero-Valent Iron for Robust Arsenite Removal. Sep. Purif. Technol. 2022, 302, 122085. [Google Scholar] [CrossRef]
  254. Wen, J.; Fu, W.; Ding, S.; Zhang, Y.; Wang, W. Pyrogallic Acid Modified Nanoscale Zero-Valent Iron Efficiently Removed Cr(VI) by Improving Adsorption and Electron Selectivity. Chem. Eng. J. 2022, 443, 136510. [Google Scholar] [CrossRef]
  255. Dai, G.; Li, X.; Fu, H.; Wang, F.; Cui, Z.; Zhao, R.; Wang, L. A Novel Oxalated Zero-Valent Iron Nanoparticle for Pb(II) Removal from Aqueous Solution: Performance and Synergistic Mechanisms. Sep. Purif. Technol. 2022, 302, 122017. [Google Scholar] [CrossRef]
  256. Li, X.; Gao, M.; Huo, Y.; Liu, H.; Li, J.; Huang, T.; Ye, R.; Li, W. Impacts of Shell Structure on Nitrate-Reduction Activity and Air Stability of Nanoscale Zero-Valent Iron. Environ. Sci. Pollut. Res. 2022, 29, 80683–80692. [Google Scholar] [CrossRef] [PubMed]
  257. Mirhosseininia, J.; Sabbaghi, S.; Mirbagheri, N.S.; Zerafat, M.M. Treatment of As-Contaminated Drinking Water Using a Nano Zero-Valent Iron/Copper Slag Nanocomposite. J. Water Process Eng. 2022, 49, 103011. [Google Scholar] [CrossRef]
  258. Liu, J.; Shi, S.; Shu, J.; Li, C.; He, H.; Xiao, C.; Dong, X.; He, Y.; Liao, J.; Liu, N.; et al. Synthesis and Characterization of Waste Commercially Available Polyacrylonitrile Fiber-Based New Composites for Efficient Removal of Uranyl from U(VI)–CO3 Solutions. Sci. Total Environ. 2022, 822, 153507. [Google Scholar] [CrossRef] [PubMed]
  259. Shanableh, A.; Bhattacharjee, S.; Sadik, S. Evaluating Iron-Based Nanoparticles for Ciprofloxacin Removal: Date Seed Extract as a Biostabilizing and a Bioreducing Agent. J. Water Process Eng. 2021, 44, 102419. [Google Scholar] [CrossRef]
  260. Pillai, R.; Balan, R.; Lekshmi, I.C.; Milina, K. Facile Auto-Combustion Synthesis and Characterization of Stable Amorphous Nanoscale Zero-Valent Iron (NZVI). Int. J. Self-Propagating High-Temp. Synth. 2021, 30, 251–256. [Google Scholar] [CrossRef]
  261. Havelka, O.; Cvek, M.; Urbánek, M.; Łukowiec, D.; Jašíková, D.; Kotek, M.; Černík, M.; Amendola, V.; Torres-Mendieta, R. On the Use of Laser Fragmentation for the Synthesis of Ligand-Free Ultra-Small Iron Nanoparticles in Various Liquid Environments. Nanomaterials 2021, 11, 1538. [Google Scholar] [CrossRef] [PubMed]
  262. Xu, H.; Gao, M.; Hu, X.; Chen, Y.; Li, Y.; Xu, X.; Zhang, R.; Yang, X.; Tang, C.; Hu, X. A Novel Preparation of S-NZVI and Its High Efficient Removal of Cr(VI) in Aqueous Solution. J. Hazard. Mater. 2021, 416, 125924. [Google Scholar] [CrossRef] [PubMed]
  263. Yang, Y.; Xu, L.; Shen, H.; Wang, J. Construction of Three-Dimensional Reduced Graphene Oxide Wrapped NZVI Doped with Al2O3 as the Ternary Fenton-like Catalyst: Optimization, Characterization and Catalytic Mechanism. Sci. Total Environ. 2021, 780, 146576. [Google Scholar] [CrossRef]
  264. Peng, F.Y.; Wang, P.W.; Liao, W.; Yu, I.S. Lignin Biopolymer for the Synthesis of Iron Nanoparticles and the Composite Applied for the Removal of Methylene Blue. Polymers 2021, 13, 3847. [Google Scholar] [CrossRef] [PubMed]
  265. Zhan, J.; Yang, X.; Zhang, X.; Wang, Y.; Cai, X.; Chen, H. Bioprecipitation Facilitates the Green Synthesis of Sulfidated Nanoscale Zero-Valent Iron Particles for Highly Selective Dechlorination of Trichloroethene. J. Environ. Chem. Eng. 2021, 9, 106050. [Google Scholar] [CrossRef]
  266. Flores-Rojas, E.; Schnabel, D.; Justo-Cabrera, E.; Solorza-Feria, O.; Poggi-Varaldo, H.M.; Breton-Deval, L. Using Nano Zero-Valent Iron Supported on Diatomite to Remove Acid Blue Dye: Synthesis, Characterization, and Toxicology Test. Sustainability 2021, 13, 899. [Google Scholar] [CrossRef]
  267. Jismy, A.; Meera, V.; RaphaelVinod, P. Comparative Study on Iron Removal Using Chemically and Greenly Synthesised Zero-Valent Iron Nanoparticles. IOP Conf. Ser. Mater. Sci. Eng. 2021, 1114, 012082. [Google Scholar] [CrossRef]
  268. Wang, B.; Zhu, C.; Ai, D.; Fan, Z. Activation of Persulfate by Green Nano-Zero-Valent Iron-Loaded Biochar for the Removal of p-Nitrophenol: Performance, Mechanism and Variables Effects. J. Hazard. Mater. 2021, 417, 126106. [Google Scholar] [CrossRef] [PubMed]
  269. Patiño-Ruiz, D.A.; Meramo-Hurtado, S.I.; González-Delgado, Á.D.; Herrera, A. Environmental Sustainability Evaluation of Iron Oxide Nanoparticles Synthesized via Green Synthesis and the Coprecipitation Method: A Comparative Life Cycle Assessment Study. ACS Omega 2021, 6, 12410–12423. [Google Scholar] [CrossRef] [PubMed]
  270. Farhadi, S.; Sohrabi, M.R.; Motiee, F.; Davallo, M. Organophosphorus Diazinon Pesticide Removing from Aqueous Solution by Zero-Valent Iron Supported on Biopolymer Chitosan: RSM Optimization Methodology. J. Polym. Environ. 2021, 29, 103–120. [Google Scholar] [CrossRef]
  271. Cheng, Y.; Chen, R.; Wang, P.; Wang, Q.; Wan, S.; Huang, S.; Su, R.; Song, Y.; Yang, X.; Fu, X. Synthesis of a Novel Biochar-Supported Polycarboxylic Acid-Functionalized Nanoiron Oxide-Encapsulated Composite for Wastewater Treatment: Removal of Cd(II), EDTA and Cd-EDTA. J. Mater. Sci. 2021, 56, 18031–18049. [Google Scholar] [CrossRef]
  272. Farooq, U.; Zhuang, J.; Wang, X.; Lyu, S. A Recyclable Polydopamine-Functionalized Reduced Graphene Oxide/Fe Nanocomposite (PDA@Fe/RGO) for the Enhanced Degradation of 1,1,1-Trichloroethane. Chem. Eng. J. 2021, 403, 126405. [Google Scholar] [CrossRef]
  273. Rončević, S.; Nemet, I.; Zagorec, V.; Selmani, A. A Facile Size Tunable One-Pot Synthesis of Dipicolinate@nZVI Core-Shell Nanoparticles: Material Properties for Trace Cadmium Ion Removal. New J. Chem. 2020, 44, 17840–17848. [Google Scholar] [CrossRef]
  274. Francy, N.; Shanthakumar, S.; Chiampo, F.; Sekhar, Y.R. Remediation of Lead and Nickel Contaminated Soil Using Nanoscale Zero-Valent Iron (nZVI) Particles Synthesized Using Green Leaves: First Results. Processes 2020, 8, 1453. [Google Scholar] [CrossRef]
  275. Fan, J.; Chen, X.; Xu, Z.; Xu, X.; Zhao, L.; Qiu, H.; Cao, X. One-Pot Synthesis of NZVI-Embedded Biochar for Remediation of Two Mining Arsenic-Contaminated Soils: Arsenic Immobilization Associated with Iron Transformation. J. Hazard. Mater. 2020, 398, 122901. [Google Scholar] [CrossRef] [PubMed]
  276. Eljamal, R.; Eljamal, O.; Maamoun, I.; Yilmaz, G.; Sugihara, Y. Enhancing the Characteristics and Reactivity of NZVI: Polymers Effect and Mechanisms. J. Mol. Liq. 2020, 315, 113714. [Google Scholar] [CrossRef]
  277. Li, S.; Yang, F.; Li, J.; Cheng, K. Porous Biochar-Nanoscale Zero-Valent Iron Composites: Synthesis, Characterization and Application for Lead Ion Removal. Scie. Total Environ. 2020, 746, 141037. [Google Scholar] [CrossRef] [PubMed]
  278. Mdlovu, N.V.; Lin, K.S.; Hsien, M.J.; Chang, C.J.; Kunene, S.C. Synthesis, Characterization, and Application of Zero-Valent Iron Nanoparticles for TNT, RDX, and HMX Explosives Decontamination in Wastewater. J. Taiwan. Inst. Chem. Eng. 2020, 114, 186–198. [Google Scholar] [CrossRef]
  279. Mortazavian, S.; Bandala, E.R.; Bae, J.H.; Chun, D.; Moon, J. Assessment of P-Nitroso Dimethylaniline (PNDA) Suitability as a Hydroxyl Radical Probe: Investigating Bleaching Mechanism Using Immobilized Zero-Valent Iron Nanoparticles. Chem. Eng. J. 2020, 385, 123748. [Google Scholar] [CrossRef]
  280. Karam, A.; Zaher, K.; Mahmoud, A.S. Comparative Studies of Using Nano Zerovalent Iron, Activated Carbon, and Green Synthesized Nano Zerovalent Iron for Textile Wastewater Color Removal Using Artificial Intelligence, Regression Analysis, Adsorption Isotherm, and Kinetic Studies. Air Soil Water Res. 2020, 13, 1178622120908273. [Google Scholar] [CrossRef]
  281. Dehghani, M.H.; Karri, R.R.; Alimohammadi, M.; Nazmara, S.; Zarei, A.; Saeedi, Z. Insights into Endocrine-Disrupting Bisphenol-A Adsorption from Pharmaceutical Effluent by Chitosan Immobilized Nanoscale Zero-Valent Iron Nanoparticles. J. Mol. Liq. 2020, 311, 113317. [Google Scholar] [CrossRef]
  282. Lahoz, R.; Natividad, E.; Mayoral, Á.; Rentenberger, C.; Díaz-Fernández, D.; Félix, E.J.; Soriano, L.; Kautek, W.; Bomati-Miguel, O. Pursuit of Optimal Synthetic Conditions for Obtaining Colloidal Zero-Valent Iron Nanoparticles by Scanning Pulsed Laser Ablation in Liquids. J. Ind. Eng. Chem. 2020, 81, 340–351. [Google Scholar] [CrossRef]
  283. Yang, L.; Chen, Y.; Ouyang, D.; Yan, J.; Qian, L.; Han, L.; Chen, M.; Li, J.; Gu, M. Mechanistic Insights into Adsorptive and Oxidative Removal of Monochlorobenzene in Biochar-Supported Nanoscale Zero-Valent Iron/Persulfate System. Chem. Eng. J. 2020, 400, 125811. [Google Scholar] [CrossRef]
  284. Wang, X.; Zhang, B.; Ma, J.; Ning, P. Novel Synthesis of Aluminum Hydroxide Gel-Coated Nano Zero-Valent Iron and Studies of Its Activity in Flocculation-Enhanced Removal of Tetracycline. J. Environ. Sci. 2020, 89, 194–205. [Google Scholar] [CrossRef] [PubMed]
  285. Bin, Q.; Lin, B.; Zhu, K.; Shen, Y.; Man, Y.; Wang, B.; Lai, C.; Chen, W. Superior Trichloroethylene Removal from Water by Sulfide-Modified Nanoscale Zero-Valent Iron/Graphene Aerogel Composite. J. Environ. Sci. 2020, 88, 90–102. [Google Scholar] [CrossRef] [PubMed]
  286. Xu, J.; Avellan, A.; Li, H.; Clark, E.A.; Henkelman, G.; Kaegi, R.; Lowry, G.V. Iron and Sulfur Precursors Affect Crystalline Structure, Speciation, and Reactivity of Sulfidized Nanoscale Zerovalent Iron. Environ. Sci. Technol. 2020, 54, 13294–13303. [Google Scholar] [CrossRef] [PubMed]
  287. Xu, R.; Li, J.; Tang, J.; Wang, Y.; Niu, Y.; Lu, H. Research on the Characterization, Reactivity, and Transportability of Porous Silicon-Coated Nanoscale Zero-Valent Iron. Environ. Sci. Pollut. Res. 2020, 27, 31567–31577. [Google Scholar] [CrossRef] [PubMed]
  288. Goswami, A.; Kadam, R.G.; Tuček, J.; Sofer, Z.; Bouša, D.; Varma, R.S.; Gawande, M.B.; Zbořil, R. Fe(0)-Embedded Thermally Reduced Graphene Oxide as Efficient Nanocatalyst for Reduction of Nitro Compounds to Amines. Chem. Eng. J. 2020, 382, 122469. [Google Scholar] [CrossRef]
  289. Sawafta, R.; Shahwan, T. A Comparative Study of the Removal of Methylene Blue by Iron Nanoparticles from Water and Water-Ethanol Solutions. J. Mol. Liq. 2019, 273, 274–281. [Google Scholar] [CrossRef]
  290. Sharma, G.; Kumar, A.; Naushad, M.; Kumar, A.; Al-Muhtaseb, A.H.; Dhiman, P.; Ghfar, A.A.; Stadler, F.J.; Khan, M.R. Photoremediation of Toxic Dye from Aqueous Environment Using Monometallic and Bimetallic Quantum Dots Based Nanocomposites. J. Clean. Prod. 2016, 172, 2919–2930. [Google Scholar] [CrossRef]
  291. Hamdy, A.; Mostafa, M.K.; Nasr, M. Zero-Valent Iron Nanoparticles for Methylene Blue Removal from Aqueous Solutions and Textile Wastewater Treatment, with Cost Estimation. Water Sci. Technol. 2018, 78, 367–378. [Google Scholar] [CrossRef] [PubMed]
  292. Mikhailov, I.; Levina, V.; Leybo, D.; Masov, V.; Tagirov, M.; Kuznetsov, D. Synthesis, Characterization and Reactivity of Nanostructured Zero-Valent Iron Particles for Degradation of Azo Dyes. Int. J. Nanosci. 2017, 16, 1750017. [Google Scholar] [CrossRef]
  293. Mondal, P.; Anweshan, A.; Purkait, M.K. Green Synthesis and Environmental Application of Iron-Based Nanomaterials and Nanocomposite: A Review. Chemosphere 2020, 259, 127509. [Google Scholar] [CrossRef]
  294. Plachtová, P.; Medříková, Z.; Zbořil, R.; Tuček, J.; Varma, R.S.; Maršálek, B. Iron and Iron Oxide Nanoparticles Synthesized with Green Tea Extract: Differences in Ecotoxicological Profile and Ability to Degrade Malachite Green. ACS Sustain. Chem. Eng. 2018, 6, 8679–8687. [Google Scholar] [CrossRef]
  295. Prasad, A.S. Iron Oxide Nanoparticles Synthesized by Controlled Bio-Precipitation Using Leaf Extract of Garlic Vine (Mansoa Alliacea). Mater. Sci. Semicond. Process 2016, 53, 79–83. [Google Scholar] [CrossRef]
  296. Wang, Z.; Yu, C.; Fang, C.; Mallavarapu, M. Dye Removal Using Iron-Polyphenol Complex Nanoparticles Synthesized by Plant Leaves. Environ. Technol. Innov. 2014, 1–2, 29–34. [Google Scholar] [CrossRef]
  297. Weng, X.; Huang, L.; Chen, Z.; Megharaj, M.; Naidu, R. Synthesis of Iron-Based Nanoparticles by Green Tea Extract and Their Degradation of Malachite. Ind. Crops Prod. 2013, 51, 342–347. [Google Scholar] [CrossRef]
  298. Suazo-Hernández, J.; Sepúlveda, P.; Cáceres-Jensen, L.; Castro-Rojas, J.; Poblete-Grant, P.; Bolan, N.; de la Luz Mora, M. NZVI-Based Nanomaterials Used for Phosphate Removal from Aquatic Systems. Nanomaterials 2023, 13, 399. [Google Scholar] [CrossRef] [PubMed]
  299. Panić, S.; Petronijević, M.; Vukmirović, J.; Grba, N.; Savić, S. Green Synthesis of Nanoscale Zero-Valent Iron Aggregates for Catalytic Degradation of Textile Dyes. Catal. Lett. 2023, 153, 3605–3619. [Google Scholar] [CrossRef]
  300. Rashtbari, Y.; Sher, F.; Afshin, S.; Hamzezadeh, A.; Ahmadi, S.; Azhar, O.; Rastegar, A.; Ghosh, S.; Poureshgh, Y. Green Synthesis of Zero-Valent Iron Nanoparticles and Loading Effect on Activated Carbon for Furfural Adsorption. Chemosphere 2022, 287, 132114. [Google Scholar] [CrossRef] [PubMed]
  301. Le, N.T.; Dang, T.D.; Hoang Binh, K.; Nguyen, T.M.; Xuan, T.N.; La, D.D.; Kumar Nadda, A.; Chang, S.W.; Nguyen, D.D. Green Synthesis of Highly Stable Zero-Valent Iron Nanoparticles for Organic Dye Treatment Using Cleistocalyx Operculatus Leaf Extract. Sustain. Chem. Pharm. 2022, 25, 100598. [Google Scholar] [CrossRef]
  302. Zhang, C.; Tang, J.; Gao, F.; Yu, C.; Li, S.; Lyu, H.; Sun, H. Tetrahydrofuran Aided Self-Assembly Synthesis of NZVI@gBC Composite as Persulfate Activator for Degradation of 2,4-Dichlorophenol. Chem. Eng. J. 2022, 431, 134063. [Google Scholar] [CrossRef]
  303. Van Hoang, N.; Nguyen-Thi, L.; Kim, G.M.; Dang, T.D.; Ngoc Toan, V.; La, D.D. Green Synthesis of Zero-Valent Iron Nanoparticles by Cleistocalyx Operculatus Leaf Extract Using Microfluidic Device for Degradation of the Rhodamine B Dye. Adv. Nat. Sci. Nanosci. Nanotechnol. 2022, 13, 045007. [Google Scholar] [CrossRef]
  304. Zhou, Y.; Li, X. Green Synthesis of Modified Polyethylene Packing Supported Tea Polyphenols-NZVI for Nitrate Removal from Wastewater: Characterization and Mechanisms. Sci. Total Environ. 2022, 806, 150596. [Google Scholar] [CrossRef] [PubMed]
  305. Yang, L.; Shen, J.; Zhang, W.; Wu, W.; Wei, Z.; Chen, M.; Yan, J.; Qian, L.; Han, L.; Li, J.; et al. Hydrothermally Assisted Synthesis of Nano Zero-Valent Iron Encapsulated in Biomass-Derived Carbon for Peroxymonosulfate Activation: The Performance and Mechanisms for Efficient Degradation of Monochlorobenzene. Sci. Total Environ. 2022, 829, 154645. [Google Scholar] [CrossRef] [PubMed]
  306. Van Hoang, N.; Thi Xuan Quynh, N.; Dang, T.D.; Nguyen Xuan, T.; Ngoc Toan, V.; Duc La, D. Green Synthesis of Fe/Graphene Nanocomposite Using Cleistocalyx Operculatus Leaf Extract as a Reducing Agent: Removal of Pollutants (RhB Dye and Cr6+ Ions) in Aqueous Media. ChemistrySelect 2022, 7, e202203499. [Google Scholar] [CrossRef]
  307. Deewan, R.; Yan, D.Y.S.; Khamdahsag, P.; Tanboonchuy, V. Remediation of Arsenic-Contaminated Water by Green Zero-Valent Iron Nanoparticles. Environ. Sci. Pollut. Res. 2022, 30, 90352–90361. [Google Scholar] [CrossRef] [PubMed]
  308. Puiatti, G.A.; de Carvalho, J.P.; de Matos, A.T.; Lopes, R.P. Green Synthesis of Fe0 Nanoparticles Using Eucalyptus Grandis Leaf Extract: Characterization and Application for Dye Degradation by a (Photo)Fenton-like Process. J. Environ. Manag. 2022, 311, 114828. [Google Scholar] [CrossRef] [PubMed]
  309. Hür, C.; Erken, E. Assessment of Green Tea-Enabled Iron/Calcined Bentonite Nanocomposites for Phosphate Removal and Recovery. J. Environ. Chem. Eng. 2022, 10, 108519. [Google Scholar] [CrossRef]
  310. Sun, H.; Hua, Y.; Zhao, Y. Synchronous Efficient Reduction of Cr (VI) and Removal of Total Chromium by Corn Extract / Fe (III) System. Environ. Sci. Pollut. Res. 2022, 29, 28552–28564. [Google Scholar] [CrossRef] [PubMed]
  311. Lin, J.; Xue, C.; Guo, S.; Owens, G.; Chen, Z. Effects of Green Synthesized and Commercial NZVI on Crystal Violet Degradation by Burkholderia Vietnamiensis C09V: Dose-Dependent Toxicity and Biocompatibility. Chemosphere 2021, 279, 130612. [Google Scholar] [CrossRef] [PubMed]
  312. Qu, J.; Liu, Y.; Cheng, L.; Jiang, Z.; Zhang, G.; Deng, F.; Wang, L.; Han, W.; Zhang, Y. Green Synthesis of Hydrophilic Activated Carbon Supported Sulfide NZVI for Enhanced Pb(II) Scavenging from Water: Characterization, Kinetics, Isotherms and Mechanisms. J. Hazard. Mater. 2021, 403, 123607. [Google Scholar] [CrossRef]
  313. Kadhum, S.T.; Alkindi, G.Y.; Albayati, T.M. Eco Friendly Adsorbents for Removal of Phenol from Aqueous Solution Employing Nanoparticle Zero-Valent Iron Synthesized from Modified Green Tea Bio-Waste and Supported on Silty Clay. Chin. J. Chem. Eng. 2021, 36, 19–28. [Google Scholar] [CrossRef]
  314. Tan, W.; Ruan, Y.; Diao, Z.; Song, G.; Su, M.; Hou, L.; Chen, D.; Kong, L.; Deng, H. Removal of Levofloxacin through Adsorption and Peroxymonosulfate Activation Using Carbothermal Reduction Synthesized NZVI/Carbon Fiber. Chemosphere 2021, 280, 130626. [Google Scholar] [CrossRef] [PubMed]
  315. Yang, J.; Wang, S.; Xu, N.; Ye, Z.; Yang, H.; Huangfu, X. Synthesis of Montmorillonite-Supported Nano-Zero-Valent Iron via Green Tea Extract: Enhanced Transport and Application for Hexavalent Chromium Removal from Water and Soil. J. Hazard. Mater. 2021, 419, 126461. [Google Scholar] [CrossRef] [PubMed]
  316. Abdelfatah, A.M.; Fawzy, M.; El-Khouly, M.E.; Eltaweil, A.S. Efficient Adsorptive Removal of Tetracycline from Aqueous Solution Using Phytosynthesized Nano-Zero Valent Iron. J. Saudi Chem. Soc. 2021, 25, 101365. [Google Scholar] [CrossRef]
  317. Al Kindi, G.Y.; Hassan, A.K.; Yahya, D.G.H.; Alhaidri, H.A. The Nanoparticles Zero-Valent Synthesis by Black Tea Extract to Remove Rb 238 Using Synthetic and Natural Wastewater by Packed Bed Reactor. IOP Conf. Ser. Earth Environ. Sci. 2021, 779, 012092. [Google Scholar] [CrossRef]
  318. Ye, Z.; Xu, N.; Li, D.; Qian, J.; Du, C.; Chen, M. Vitamin C Mediates the Activation of Green Tea Extract to Modify Nanozero-Valent Iron Composites: Enhanced Transport in Heterogeneous Porous Media and the Removal of Hexavalent Chromium. J. Hazard. Mater. 2021, 411, 125042. [Google Scholar] [CrossRef] [PubMed]
  319. Salmani, M.H.; Abedi, M.; Mozaffari, S.A.; Mahvi, A.H.; Sheibani, A.; Jalili, M. Simultaneous Reduction and Adsorption of Arsenite Anions by Green Synthesis of Iron Nanoparticles Using Pomegranate Peel Extract. J. Environ. Health Sci. Eng. 2021, 19, 603–612. [Google Scholar] [CrossRef] [PubMed]
  320. Khunjan, U.; Kasikamphaiboon, P. Green Synthesis of Kaolin-Supported Nanoscale Zero-Valent Iron Using Ruellia Tuberosa Leaf Extract for Effective Decolorization of Azo Dye Reactive Black 5. Arab. J. Sci. Eng. 2021, 46, 383–394. [Google Scholar] [CrossRef]
  321. Panagou, I.; Noutsopoulos, C.; Mystrioti, C.; Barka, E.; Koumaki, E.; Kalli, M.; Malamis, S.; Papassiopi, N.; Mamais, D. Assessing the Performance of Environmentally Friendly-Produced Zerovalent Iron Nanoparticles to Remove Pharmaceuticals from Water. Sustainability 2021, 13, 12708. [Google Scholar] [CrossRef]
  322. Mousazadeh, S.; Shariati, S.; Yousefi, M.; Baniyaghoob, S.; Kefayati, H. Hexavalent Chromium Removal Using Ionic Liquid Coated Magnetic Nano Zero-Valent Iron Biosynthesized by Camellia Sinensis Extract. Int. J. Environ. Res. 2021, 15, 1017–1036. [Google Scholar] [CrossRef]
  323. Jain, R.; Mendiratta, S.; Kumar, L.; Srivastava, A. Green Synthesis of Iron Nanoparticles Using Artocarpus Heterophyllus Peel Extract and Their Application as a Heterogeneous Fenton-like Catalyst for the Degradation of Fuchsin Basic Dye. Curr. Res. Green Sustain. Chem. 2021, 4, 100086. [Google Scholar] [CrossRef]
  324. Guha, T.; Gopal, G.; Das, H.; Mukherjee, A.; Kundu, R. Nanopriming with Zero-Valent Iron Synthesized Using Pomegranate Peel Waste: A “Green” Approach for Yield Enhancement in Oryza Sativa L. Cv. Gonindobhog. Plant Physiol. Biochem. 2021, 163, 261–275. [Google Scholar] [CrossRef] [PubMed]
  325. Ma, L.; Du, Y.; Chen, S.; Zhang, F.; Zhan, W.; Du, D.; Zhang, T.C. Nanoscale Zero-Valent Iron Coupling with Shewanella Oneidensis MR-1 for Enhanced Reduction/Removal of Aqueous Cr(VI). Sep. Purif. Technol. 2021, 277, 119488. [Google Scholar] [CrossRef]
  326. Abdelfatah, A.M.; Fawzy, M.; Eltaweil, A.S.; El-Khouly, M.E. Green Synthesis of Nano-Zero-Valent Iron Using Ricinus Communis Seeds Extract: Characterization and Application in the Treatment of Methylene Blue-Polluted Water. ACS Omega 2021, 6, 25397–25411. [Google Scholar] [CrossRef] [PubMed]
  327. Liu, H.; Sun, Y.; He, X.; Zhang, H.; Wei, J.; Zhu, L. Microbiological Synthesis of Denitrifying Bacteria-Iron Nanopartical Composite Material and Its Eminent Performance in Removal of Nitrate-N. Sep. Purif. Technol. 2021, 267, 118663. [Google Scholar] [CrossRef]
  328. Ali, I.; Afshinb, S.; Poureshgh, Y.; Azari, A.; Rashtbari, Y.; Feizizadeh, A.; Hamzezadeh, A.; Fazlzadeh, M. Green Preparation of Activated Carbon from Pomegranate Peel Coated with Zero-Valent Iron Nanoparticles (NZVI) and Isotherm and Kinetic Studies of Amoxicillin Removal in Water. Environ. Sci. Pollut. Res. 2020, 27, 36732–36743. [Google Scholar] [CrossRef] [PubMed]
  329. Rashtbari, Y.; Américo-Pinheiro, J.H.P.; Bahrami, S.; Fazlzadeh, M.; Arfaeinia, H.; Poureshgh, Y. Efficiency of Zeolite Coated with Zero-Valent Iron Nanoparticles for Removal of Humic Acid from Aqueous Solutions. Water Air Soil Pollut. 2020, 231, 514. [Google Scholar] [CrossRef]
  330. Hassan, A.K.; Al-Kindi, G.Y.; Ghanim, D. Green Synthesis of Bentonite-Supported Iron Nanoparticles as a Heterogeneous Fenton-like Catalyst: Kinetics of Decolorization of Reactive Blue 238 Dye. Water Sci. Eng. 2020, 13, 286–298. [Google Scholar] [CrossRef]
  331. Rashtbari, Y.; Hazrati, S.; Azari, A.; Afshin, S.; Fazlzadeh, M.; Vosoughi, M. A Novel, Eco-Friendly and Green Synthesis of PPAC-ZnO and PPAC-NZVI Nanocomposite Using Pomegranate Peel: Cephalexin Adsorption Experiments, Mechanisms, Isotherms and Kinetics. Adv. Powder Technol. 2020, 31, 1612–1623. [Google Scholar] [CrossRef]
  332. Liu, X.; Yang, L.; Zhao, H.; Wang, W. Pyrolytic Production of Zerovalent Iron Nanoparticles Supported on Rice Husk-Derived Biochar: Simple, in Situ Synthesis and Use for Remediation of Cr(VI)-Polluted Soils. Sci. Total Environ. 2020, 708, 134479. [Google Scholar] [CrossRef]
  333. Lu, J.; Zhang, C.; Wu, J. One-Pot Synthesis of Magnetic Algal Carbon/Sulfidated Nanoscale Zerovalent Iron Composites for Removal of Bromated Disinfection by-Product. Chemosphere 2020, 250, 126257. [Google Scholar] [CrossRef]
  334. Zhang, Y.; Liu, N.; Yang, Y.; Li, J.; Wang, S.; Lv, J.; Tang, R. Novel Carbothermal Synthesis of Fe, N Co-Doped Oak Wood Biochar (Fe/N-OB) for Fast and Effective Cr(VI) Removal. Colloids Surf. A Physicochem. Eng. Asp. 2020, 600, 124926. [Google Scholar] [CrossRef]
  335. Mahmoud, A.S.; Farag, R.S.; Elshfai, M.M. Reduction of Organic Matter from Municipal Wastewater at Low Cost Using Green Synthesis Nano Iron Extracted from Black Tea: Artificial Intelligence with Regression Analysis. Egypt. J. Pet. 2020, 29, 9–20. [Google Scholar] [CrossRef]
  336. Shad, S.; Belinga-Desaunay-Nault, M.F.A.; Sohail; Bashir, N.; Lynch, I. Removal of Contaminants from Canal Water Using Microwave Synthesized Zero Valent Iron Nanoparticles. Environ. Sci. 2020, 6, 3057–3065. [Google Scholar] [CrossRef]
  337. Li, Q.; Liu, D.; Wang, T.; Chen, C.; Gadd, G.M. Iron Coral: Novel Fungal Biomineralization of Nanoscale Zerovalent Iron Composites for Treatment of Chlorinated Pollutants. Chem. Eng. J. 2020, 402, 126263. [Google Scholar] [CrossRef]
  338. Zhang, Y.; Jiao, X.; Liu, N.; Lv, J.; Yang, Y. Enhanced Removal of Aqueous Cr(VI) by a Green Synthesized Nanoscale Zero-Valent Iron Supported on Oak Wood Biochar. Chemosphere 2020, 245, 125542. [Google Scholar] [CrossRef] [PubMed]
  339. Romeh, A.A.; Ibrahim Saber, R.A. Green Nano-Phytoremediation and Solubility Improving Agents for the Remediation of Chlorfenapyr Contaminated Soil and Water. J. Environ. Manag. 2020, 260, 110104. [Google Scholar] [CrossRef] [PubMed]
  340. Yi, Y.; Tu, G.; Tsang, P.E.; Xiao, S.; Fang, Z. Green Synthesis of Iron-Based Nanoparticles from Extracts of Nephrolepis Auriculata and Applications for Cr(VI) Removal. Mater. Lett. 2019, 234, 388–391. [Google Scholar] [CrossRef]
  341. Desalegn, B.; Megharaj, M.; Chen, Z.; Naidu, R. Green Synthesis of Zero Valent Iron Nanoparticle Using Mango Peel Extract and Surface Characterization Using XPS and GC-MS. Heliyon 2019, 5, e01750. [Google Scholar] [CrossRef] [PubMed]
  342. Chen, X.; Yi, Z.; Chen, G.; Ma, X.; Su, W.; Cui, X.; Li, X. DOX-Assisted Functionalization of Green Tea Polyphenol Nanoparticles for Effective Chemo-Photothermal Cancer Therapy. J. Mater. Chem. B 2019, 7, 4066–4078. [Google Scholar] [CrossRef]
  343. Afsheen, S.; Tahir, M.B.; Iqbal, T.; Liaqat, A.; Abrar, M. Green Synthesis and Characterization of Novel Iron Particles by Using Different Extracts. J. Alloys Compd. 2018, 732, 935–944. [Google Scholar] [CrossRef]
  344. Gan, L.; Lu, Z.; Cao, D.; Chen, Z. Effects of Cetyltrimethylammonium Bromide on the Morphology of Green Synthesized Fe3O4 Nanoparticles Used to Remove Phosphate. Mater. Sci. Eng. C 2018, 82, 41–45. [Google Scholar] [CrossRef] [PubMed]
  345. Jagathesan, G.; Rajiv, P. Biosynthesis and Characterization of Iron Oxide Nanoparticles Using Eichhornia Crassipes Leaf Extract and Assessing Their Antibacterial Activity. Biocatal. Agric. Biotechnol. 2018, 13, 90–94. [Google Scholar] [CrossRef]
  346. Sebastian, A.; Nangia, A.; Prasad, M.N.V. A Green Synthetic Route to Phenolics Fabricated Magnetite Nanoparticles from Coconut Husk Extract: Implications to Treat Metal Contaminated Water and Heavy Metal Stress in Oryza Sativa L. J. Clean Prod. 2018, 174, 355–366. [Google Scholar] [CrossRef]
  347. Rana, A.; Kumari, N.; Tyagi, M.; Jagadevan, S. Leaf-Extract Mediated Zero-Valent Iron for Oxidation of Arsenic (III): Preparation, Characterization and Kinetics. Chem. Eng. J. 2018, 347, 91–100. [Google Scholar] [CrossRef]
  348. Machado, S.; Pacheco, J.G.; Nouws, H.P.A.; Albergaria, J.T.; Delerue-Matos, C. Green Zero-Valent Iron Nanoparticles for the Degradation of Amoxicillin. Int. J. Environ. Sci. Technol. 2017, 14, 1109–1118. [Google Scholar] [CrossRef]
  349. Sathya, K.; Saravanathamizhan, R.; Baskar, G. Ultrasound Assisted Phytosynthesis of Iron Oxide Nanoparticle. Ultrason. Sonochem. 2017, 39, 446–451. [Google Scholar] [CrossRef] [PubMed]
  350. Nithya, K.; Sathish, A.; Senthil Kumar, P.; Ramachandran, T. Fast Kinetics and High Adsorption Capacity of Green Extract Capped Superparamagnetic Iron Oxide Nanoparticles for the Adsorption of Ni(II) Ions. J. Ind. Eng. Chem. 2018, 59, 230–241. [Google Scholar] [CrossRef]
  351. Rajiv, P.; Bavadharani, B.; Kumar, M.N.; Vanathi, P. Synthesis and Characterization of Biogenic Iron Oxide Nanoparticles Using Green Chemistry Approach and Evaluating Their Biological Activities. Biocatal. Agric. Biotechnol. 2017, 12, 45–49. [Google Scholar] [CrossRef]
  352. Sangami, S.; Manu, B. Synthesis of Green Iron Nanoparticles Using Laterite and Their Application as a Fenton-like Catalyst for the Degradation of Herbicide Ametryn in Water. Environ. Technol. Innov. 2017, 8, 150–163. [Google Scholar] [CrossRef]
  353. Jin, X.; Liu, Y.; Tan, J.; Owens, G.; Chen, Z. Removal of Cr(VI) from Aqueous Solutions via Reduction and Absorption by Green Synthesized Iron Nanoparticles. J. Clean. Prod. 2018, 176, 929–936. [Google Scholar] [CrossRef]
  354. Saikia, I.; Hazarika, M.; Hussian, N.; Das, M.R.; Tamuly, C. Biogenic Synthesis of Fe2O3@SiO2 Nanoparticles for Ipso-Hydroxylation of Boronic Acid in Water. Tetrahedron Lett. 2017, 58, 4255–4259. [Google Scholar] [CrossRef]
  355. Xiao, Z.; Zhang, H.; Xu, Y.; Yuan, M.; Jing, X.; Huang, J.; Li, Q.; Sun, D. Ultra-Efficient Removal of Chromium from Aqueous Medium by Biogenic Iron Based Nanoparticles. Sep. Purif. Technol. 2017, 174, 466–473. [Google Scholar] [CrossRef]
  356. Manquián-Cerda, K.; Cruces, E.; Angélica Rubio, M.; Reyes, C.; Arancibia-Miranda, N. Preparation of Nanoscale Iron (Oxide, Oxyhydroxides and Zero-Valent) Particles Derived from Blueberries: Reactivity, Characterization and Removal Mechanism of Arsenate. Ecotoxicol. Environ. Saf. 2017, 145, 69–77. [Google Scholar] [CrossRef] [PubMed]
  357. Madhubala, V.; Kalaivani, T. Phyto and Hydrothermal Synthesis of Fe3O4@ZnO Core-Shell Nanoparticles Using Azadirachta Indica and Its Cytotoxicity Studies. Appl. Surf. Sci. 2018, 449, 584–590. [Google Scholar] [CrossRef]
  358. Wei, Y.; Fang, Z.; Zheng, L.; Tsang, E.P. Biosynthesized Iron Nanoparticles in Aqueous Extracts of Eichhornia Crassipes and Its Mechanism in the Hexavalent Chromium Removal. Appl. Surf. Sci. 2017, 399, 322–329. [Google Scholar] [CrossRef]
  359. Katata-Seru, L.; Moremedi, T.; Aremu, O.S.; Bahadur, I. Green Synthesis of Iron Nanoparticles Using Moringa Oleifera Extracts and Their Applications: Removal of Nitrate from Water and Antibacterial Activity against Escherichia Coli. J. Mol. Liq. 2018, 256, 296–304. [Google Scholar] [CrossRef]
  360. Ali, I.; Alothman, Z.A.; Alwarthan, A. Uptake of Propranolol on Ionic Liquid Iron Nanocomposite Adsorbent: Kinetic, Thermodynamics and Mechanism of Adsorption. J. Mol. Liq. 2017, 236, 205–213. [Google Scholar] [CrossRef]
  361. Gottimukkala, K.S.V.; Harika-Reddy, P.; Deeveka, Z. Green Synthesis of Iron Nanoparticles Using Green Tea Leaves Extract. J. Nanomed. Biother. Discov. 2017, 7, 1–4. [Google Scholar] [CrossRef]
  362. Fazlzadeh, M.; Rahmani, K.; Zarei, A.; Abdoallahzadeh, H.; Nasiri, F.; Khosravi, R. A Novel Green Synthesis of Zero Valent Iron Nanoparticles (NZVI) Using Three Plant Extracts and Their Efficient Application for Removal of Cr(VI) from Aqueous Solutions. Adv. Powder Technol. 2017, 28, 122–130. [Google Scholar] [CrossRef]
  363. Guo, M.; Weng, X.; Wang, T.; Chen, Z. Biosynthesized Iron-Based Nanoparticles Used as a Heterogeneous Catalyst for the Removal of 2,4-Dichlorophenol. Sep. Purif. Technol. 2017, 175, 222–228. [Google Scholar] [CrossRef]
  364. Weng, X.; Jin, X.; Lin, J.; Naidu, R.; Chen, Z. Removal of Mixed Contaminants Cr(VI) and Cu(II) by Green Synthesized Iron Based Nanoparticles. Ecol. Eng. 2016, 97, 32–39. [Google Scholar] [CrossRef]
  365. Jassal, V.; Shanker, U.; Gahlot, S. Green Synthesis of Some Iron Oxide Nanoparticles and Their Interaction with 2-Amino, 3-Amino and 4-Aminopyridines. Mater. Today Proc. 2016, 3, 1874–1882. [Google Scholar] [CrossRef]
  366. 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]
  367. Devatha, C.P.; Thalla, A.K.; Katte, S.Y. Green Synthesis of Iron Nanoparticles Using Different Leaf Extracts for Treatment of Domestic Waste Water. J. Clean. Prod. 2016, 139, 1425–1435. [Google Scholar] [CrossRef]
  368. Wang, T.; Lin, J.; Chen, Z.; Megharaj, M.; Naidu, R. Green Synthesized Iron Nanoparticles by Green Tea and Eucalyptus Leaves Extracts Used for Removal of Nitrate in Aqueous Solution. J. Clean. Prod. 2014, 83, 413–419. [Google Scholar] [CrossRef]
  369. Machado, S.; Grosso, J.P.; Nouws, H.P.A.; Albergaria, J.T.; Delerue-Matos, C. Utilization of Food Industry Wastes for the Production of Zero-Valent Iron Nanoparticles. Sci. Total Environ. 2014, 496, 233–240. [Google Scholar] [CrossRef] [PubMed]
  370. Machado, S.; Stawiński, W.; Slonina, P.; Pinto, A.R.; Grosso, J.P.; Nouws, H.P.A.; Albergaria, J.T.; Delerue-Matos, C. Application of Green Zero-Valent Iron Nanoparticles to the Remediation of Soils Contaminated with Ibuprofen. Sci. Total Environ. 2013, 461–462, 323–329. [Google Scholar] [CrossRef] [PubMed]
  371. Mohan Kumar, K.; Mandal, B.K.; Siva Kumar, K.; Sreedhara Reddy, P.; Sreedhar, B. Biobased Green Method to Synthesise Palladium and Iron Nanoparticles Using Terminalia Chebula Aqueous Extract. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2013, 102, 128–133. [Google Scholar] [CrossRef] [PubMed]
  372. Machado, S.; Pinto, S.L.; Grosso, J.P.; Nouws, H.P.A.; Albergaria, J.T.; Delerue-Matos, C. Green Production of Zero-Valent Iron Nanoparticles Using Tree Leaf Extracts. Sci. Total Environ. 2013, 445–446, 1–8. [Google Scholar] [CrossRef]
  373. Mathur, N.; Bhatnagar, P.; Verma, H. Genotoxicity of Vegetables Irrigated by Industrial Wastewater. J. Environ. Sci. 2006, 18, 964–968. [Google Scholar] [CrossRef]
  374. Pizzicato, B.; Pacifico, S.; Cayuela, D.; Mijas, G.; Riba-Moliner, M. Advancements in Sustainable Natural Dyes for Textile Applications: A Review. Molecules 2023, 28, 5954. [Google Scholar] [CrossRef] [PubMed]
  375. Ataei, M.; Maghsoudi, A.S.; Hassani, S. Dyes and Colorants. In Encyclopedia of Toxicology; Elsevier: Amsterdam, The Netherlands, 2023; Volume 3, ISBN 978-0-12824-315-2. [Google Scholar] [CrossRef]
  376. Mustroph, H. Streptocyanine Dyes. Phys. Sci. Rev. 2021, 6, 137–147. [Google Scholar] [CrossRef]
  377. Mustroph, H. Polymethine Dyes. Phys. Sci. Rev. 2020, 5, 20190084. [Google Scholar] [CrossRef]
  378. Benkhaya, S.; M’rabet, S.; El Harfi, A. A Review on Classifications, Recent Synthesis and Applications of Textile Dyes. Inorg. Chem. Commun. 2020, 115, 107891. [Google Scholar] [CrossRef]
  379. Croce, R.; Cinà, F.; Lombardo, A.; Crispeyn, G.; Cappelli, C.I.; Vian, M.; Maiorana, S.; Benfenati, E.; Baderna, D. Aquatic Toxicity of Several Textile Dye Formulations: Acute and Chronic Assays with Daphnia Magna and Raphidocelis Subcapitata. Ecotoxicol. Environ. Saf. 2017, 144, 79–87. [Google Scholar] [CrossRef] [PubMed]
  380. Kumar, L.; Bharadvaja, N. Microorganisms: A Remedial Source for Dye Pollution. In Removal of Toxic Pollutants through Microbiological and Tertiary Treatment; Elsevier: Amsterdam, The Netherlands, 2020; ISBN 978-0-12821-014-7. [Google Scholar] [CrossRef]
  381. Mojiri, A.; Zhou, J.L.; KarimiDermani, B.; Razmi, E.; Kasmuri, N. Anaerobic Membrane Bioreactor (AnMBR) for the Removal of Dyes from Water and Wastewater: Progress, Challenges, and Future Perspectives. Processes 2023, 11, 855. [Google Scholar] [CrossRef]
  382. Gita, S.; Shukla, S.P.; Deshmukhe, G.; Choudhury, T.G.; Saharan, N.; Singh, A.K. Toxicity Evaluation of Six Textile Dyes on Growth, Metabolism and Elemental Composition (C, H, N, S) of Microalgae Spirulina Platensis: The Environmental Consequences. Bull. Environ. Contam. Toxicol. 2021, 106, 302–309. [Google Scholar] [CrossRef] [PubMed]
  383. Rápó, E.; Tonk, S. Factors Affecting Synthetic Dye Adsorption; Desorption Studies: A Review of Results from the Last Five Years (2017–2021). Molecules 2021, 26, 5419. [Google Scholar] [CrossRef] [PubMed]
  384. Dave, S.; Dave, S.; Das, J. Photocatalytic Degradation of Dyes in Textile Effluent: A Green Approach to Eradicate Environmental Pollution. In The Future of Effluent Treatment Plants; Elsevier: Amsterdam, The Netherlands, 2021; ISBN 978-0-12822-956-9. [Google Scholar] [CrossRef]
  385. Atalay, S.; Ersöz, G. Hybrid Application of Advanced Oxidation Processes to Dyes’ Removal. In Green Chemistry and Water Remediation: Research and Applications; Elsevier: Amsterdam, The Netherlands, 2020; ISBN 978-0-12817-742-6. [Google Scholar] [CrossRef]
  386. Shindhal, T.; Rakholiya, P.; Varjani, S.; Pandey, A.; Ngo, H.H.; Guo, W.; Ng, H.Y.; Taherzadeh, M.J. A Critical Review on Advances in the Practices and Perspectives for the Treatment of Dye Industry Wastewater. Bioengineered 2021, 12, 70–87. [Google Scholar] [CrossRef]
  387. Samsami, S.; Mohamadi, M.; Sarrafzadeh, M.-H.; Rene, E.R.; Firoozbahr, M. Recent Advances in the Treatment of Dye-Containing Wastewater from Textile Industries: Overview and Perspectives. Process Saf. Environ. Prot. 2020, 143, 138–163. [Google Scholar] [CrossRef]
  388. Li, Z.-D.; Cheng, P.; Zhang, H.-X.; Bao, C.-H.; Zhu, X.-M.; Shao, B.-Y.; Ma, J. Perspectives on Dye Wastewater Treatment. Oxid. Commun. 2015, 38, 1470–1479. [Google Scholar]
  389. Bayineni, V.K. Bioremediation of Toxic Dyes for Zero Waste. In Biotechnology for Zero Waste: Emerging Waste Management Techniques; Wiley: Hoboken, NJ, USA, 2021; ISBN 978-3-52783-206-4. [Google Scholar] [CrossRef]
  390. Roy, M.; Saha, R. Dyes and Their Removal Technologies from Wastewater: A Critical Review. In Intelligent Environmental Data Monitoring for Pollution Management; Academic Press: Cambridge, MA, USA, 2020; ISBN 978-0-12819-671-7. [Google Scholar] [CrossRef]
  391. Thakare, S.R.; Thakare, J.; Kosankar, P.T.; Pal, M.R. A Chief, Industrial Waste, Activated Red Mud for Subtraction of Methylene Blue Dye from Environment. Mater. Today Proc. 2019, 29, 822–827. [Google Scholar] [CrossRef]
  392. Vishnu, D.; Dhandapani, B.; Authilingam, S.; Sivakumar, S.V. A Comprehensive Review of Effective Adsorbents Used for the Removal of Dyes from Wastewater. Curr. Anal. Chem. 2022, 18, 255–268. [Google Scholar] [CrossRef]
  393. Ernawati, L.; Reza, M.; Synthia, A.C.; Kartikasari, D.A.; Maharsih, I.K.; Halim, A. Role of Chemical Activating Agent on the Characteristics of Activated Carbon Derived from Fruit-Peel Waste for Aqueous Dye Removal. Key Eng. Mater. 2022, 937, 165–180. [Google Scholar] [CrossRef]
  394. Tripathi, A.K.; Tiwari, A.; Shankar, R.; Khare, P. Recent Trends of Hybrid Systems and Their Importance in Dye Degradation. In Photocatalytic Degradation of Dyes; Elsevier: Amsterdam, The Netherlands, 2021; ISBN 978-0-12823-876-9. [Google Scholar] [CrossRef]
  395. Yew, G.Y.; Tan, X.; Chew, K.W.; Chang, J.-S.; Tao, Y.; Jiang, N.; Show, P.L. Thermal-Fenton Mechanism with Sonoprocessing for Rapid Non-Catalytic Transesterification of Microalgal to Biofuel Production. Chem. Eng. J. 2021, 408, 127264. [Google Scholar] [CrossRef]
  396. Hu, Y.; Zhou, S.; Pan, X.; Zhou, F.; Sun, Y.; Liu, M.; Zhang, D.; Zhang, L. Fe Nanoparticles Synthesized by Pomegranate Leaves for Treatment of Malachite Green. J. Wuhan Univ. Technol. Mater. Sci. Ed. 2022, 37, 350–354. [Google Scholar] [CrossRef]
  397. Kirti; Kamsonlian, S.; Agarwal, V. Review on Synthesis of Plant-Mediated Green Iron Nanoparticles and Their Application for Decolorization of Dyes. Mater. Today Proc. 2022, 78, 99–107. [Google Scholar] [CrossRef]
  398. Luo, F.; Chen, Z.; Megharaj, M.; Naidu, R. Biomolecules in Grape Leaf Extract Involved in One-Step Synthesis of Iron-Based Nanoparticles. RSC Adv. 2014, 4, 53467–53474. [Google Scholar] [CrossRef]
  399. Luo, F.; Yang, D.; Chen, Z.; Megharaj, M.; Naidu, R. Characterization of Bimetallic Fe/Pd Nanoparticles by Grape Leaf Aqueous Extract and Identification of Active Biomolecules Involved in the Synthesis. Sci. Total Environ. 2016, 562, 526–532. [Google Scholar] [CrossRef]
  400. Monga, Y.; Kumar, P.; Sharma, R.K.; Filip, J.; Varma, R.S.; Zbořil, R.; Gawande, M.B. Sustainable Synthesis of Nanoscale Zerovalent Iron Particles for Environmental Remediation. ChemSusChem 2020, 13, 3288–3305. [Google Scholar] [CrossRef]
  401. Luo, F.; Yang, D.; Chen, Z.; Megharaj, M.; Naidu, R. The Mechanism for Degrading Orange II Based on Adsorption and Reduction by Ion-Based Nanoparticles Synthesized by Grape Leaf Extract. J. Hazard. Mater. 2015, 296, 37–45. [Google Scholar] [CrossRef] [PubMed]
  402. Wang, X.; Wang, A.; Ma, J.; Fu, M. Facile Green Synthesis of Functional Nanoscale Zero-Valent Iron and Studies of Its Activity toward Ultrasound-Enhanced Decolorization of Cationic Dyes. Chemosphere 2017, 166, 80–88. [Google Scholar] [CrossRef] [PubMed]
  403. Hoag, G.E.; Collins, J.B.; Holcomb, J.L.; Hoag, J.R.; Nadagouda, M.N.; Varma, R.S. Degradation of Bromothymol Blue by “greener” Nano-Scale Zero-Valent Iron Synthesized Using Tea Polyphenols. J. Mater. Chem. 2009, 19, 8671–8677. [Google Scholar] [CrossRef]
  404. Shahwan, T.; Abu Sirriah, S.; Nairat, M.; Boyaci, E.; Eroĝlu, A.E.; Scott, T.B.; Hallam, K.R. Green Synthesis of Iron Nanoparticles and Their Application as a Fenton-like Catalyst for the Degradation of Aqueous Cationic and Anionic Dyes. Chem. Eng. J. 2011, 172, 258–266. [Google Scholar] [CrossRef]
  405. Huang, L.; Weng, X.; Chen, Z.; Megharaj, M.; Naidu, R. Synthesis of Iron-Based Nanoparticles Using Oolong Tea Extract for the Degradation of Malachite Green. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2014, 117, 801–804. [Google Scholar] [CrossRef] [PubMed]
  406. Nasiri, J.; Motamedi, E.; Naghavi, M.R.; Ghafoori, M. Removal of Crystal Violet from Water Using β-Cyclodextrin Functionalized Biogenic Zero-Valent Iron Nanoadsorbents Synthesized via Aqueous Root Extracts of Ferula Persica. J. Hazard. Mater. 2019, 367, 325–338. [Google Scholar] [CrossRef] [PubMed]
  407. Roy, A.; Singh, V.; Sharma, S.; Ali, D.; Azad, A.K.; Kumar, G.; Emran, T.B. Antibacterial and Dye Degradation Activity of Green Synthesized Iron Nanoparticles. J. Nanomater. 2022, 2022, 3636481. [Google Scholar] [CrossRef]
  408. Haseena, S.; Shanavas, S.; Ahamad, T.; Alshehri, S.M.; Baskaran, P.; Duraimurugan, J.; Acevedo, R.; Khan, M.A.M.; Anbarasan, P.M.; Jayamani, N. Investigation on Photocatalytic Activity of Bio-Treated α-Fe2O3 Nanoparticles Using Phyllanthus Niruri and Moringa Stenopetala Leaf Extract against Methylene Blue and Phenol Molecules: Kinetics, Mechanism and Stability. J. Environ. Chem. Eng. 2021, 9, 104996. [Google Scholar] [CrossRef]
  409. Shaker Ardakani, L.; 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, e06159. [Google Scholar] [CrossRef]
  410. Rawat, S.; Samreen, K.; Nayak, A.K.; Singh, J.; Koduru, J.R. Fabrication of Iron Nanoparticles Using Parthenium: A Combinatorial Eco-Innovative Approach to Eradicate Crystal Violet Dye and Phosphate from the Aqueous Environment. Environ. Nanotechnol. Monit. Manag. 2021, 15, 100426. [Google Scholar] [CrossRef]
  411. Pai, S.; Kini, S.M.; Narasimhan, M.K.; Pugazhendhi, A.; Selvaraj, R. Structural Characterization and Adsorptive Ability of Green Synthesized Fe3O4 Nanoparticles to Remove Acid Blue 113 Dye. Surf. Interfaces 2021, 23, 100947. [Google Scholar] [CrossRef]
  412. Balu, P.; Asharani, I.V.; Thirumalai, D. Catalytic Degradation of Hazardous Textile Dyes by Iron Oxide Nanoparticles Prepared from Raphanus Sativus Leaves’ Extract: A Greener Approach. J. Mater. Sci. Mater. Electron. 2020, 31, 10669–10676. [Google Scholar] [CrossRef]
  413. Noreen, S.; Mustafa, G.; Ibrahim, S.M.; Naz, S.; Iqbal, M.; Yaseen, M.; Javed, T.; Nisar, J. Iron Oxide (Fe2O3) Prepared via Green Route and Adsorption Efficiency Evaluation for an Anionic Dye: Kinetics, Isotherms and Thermodynamics Studies. J. Mater. Res. Technol. 2020, 9, 4206–4217. [Google Scholar] [CrossRef]
  414. Das, C.; Sen, S.; Singh, T.; Ghosh, T.; Paul, S.S.; Kim, T.W.; Jeon, S.; Maiti, D.K.; Im, J.; Biswas, G. Green Synthesis, Characterization and Application of Natural Product Coated Magnetite Nanoparticles for Wastewater Treatment. Nanomaterials 2020, 10, 1615. [Google Scholar] [CrossRef] [PubMed]
  415. Jadidi Kouhbanani, M.A.; Beheshtkhoo, N.; Amani, A.M.; Taghizadeh, S.; Beigi, V.; Zakeri Bazmandeh, A.; Khalaf, N. Green Synthesis of Iron Oxide Nanoparticles Using Artemisia Vulgaris Leaf Extract and Their Application as a Heterogeneous Fenton-like Catalyst for the Degradation of Methyl Orange. Mater. Res. Express 2018, 5, 115013. [Google Scholar] [CrossRef]
  416. Beheshtkhoo, N.; Kouhbanani, M.A.J.; Savardashtaki, A.; Amani, A.M.; Taghizadeh, S. Green Synthesis of Iron Oxide Nanoparticles by Aqueous Leaf Extract of Daphne Mezereum as a Novel Dye Removing Material. Appl. Phys. A Mater. Sci. Process 2018, 124, 363. [Google Scholar] [CrossRef]
  417. Ali, I.; Peng, C.; Khan, Z.M.; Sultan, M.; Naz, I. Green Synthesis of Phytogenic Magnetic Nanoparticles and Their Applications in the Adsorptive Removal of Crystal Violet from Aqueous Solution. Arab. J. Sci. Eng. 2018, 43, 6245–6259. [Google Scholar] [CrossRef]
  418. Ebrahiminezhad, A.; Taghizadeh, S.; Ghasemi, Y.; Berenjian, A. Green Synthesized Nanoclusters of Ultra-Small Zero Valent Iron Nanoparticles as a Novel Dye Removing Material. Sci. Total Environ. 2018, 621, 1527–1532. [Google Scholar] [CrossRef] [PubMed]
  419. Somchaidee, P.; Tedsree, K. Green Synthesis of High Dispersion and Narrow Size Distribution of Zero-Valent Iron Nanoparticles Using Guava Leaf (Psidium guajava L.) Extract. Adv. Nat. Sci. Nanosci. Nanotechnol. 2018, 9, 035006. [Google Scholar] [CrossRef]
  420. Khan, Z.; Al-Thabaiti, S.A. Green Synthesis of Zero-Valent Fe-Nanoparticles: Catalytic Degradation of Rhodamine B, Interactions with Bovine Serum Albumin and Their Enhanced Antimicrobial Activities. J. Photochem. Photobiol. B 2018, 180, 259–267. [Google Scholar] [CrossRef]
  421. Radini, I.A.; Hasan, N.; Malik, M.A.; Khan, Z. Biosynthesis of Iron Nanoparticles Using Trigonella Foenum-Graecum Seed Extract for Photocatalytic Methyl Orange Dye Degradation and Antibacterial Applications. J. Photochem. Photobiol. B 2018, 183, 154–163. [Google Scholar] [CrossRef]
  422. Ozkan, Z.Y.; Cakirgoz, M.; Kaymak, E.S.; Erdim, E. Rapid Decolorization of Textile Wastewater by Green Synthesized Iron Nanoparticles. Water Sci. Technol. 2018, 77, 511–517. [Google Scholar] [CrossRef] [PubMed]
  423. Alizadeh, N.; Shariati, S.; Besharati, N. Adsorption of Crystal Violet and Methylene Blue on Azolla and Fig Leaves Modified with Magnetite Iron Oxide Nanoparticles. Int. J. Environ. Res. 2017, 11, 197–206. [Google Scholar] [CrossRef]
  424. 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]
  425. Weng, X.; Guo, M.; Luo, F.; Chen, Z. One-Step Green Synthesis of Bimetallic Fe/Ni Nanoparticles by Eucalyptus Leaf Extract: Biomolecules Identification, Characterization and Catalytic Activity. Chem. Eng. J. 2017, 308, 904–911. [Google Scholar] [CrossRef]
  426. Prasad, C.; Karlapudi, S.; Venkateswarlu, P.; Bahadur, I.; Kumar, S. Green Arbitrated Synthesis of Fe3O4 Magnetic Nanoparticles with Nanorod Structure from Pomegranate Leaves and Congo Red Dye Degradation Studies for Water Treatment. J. Mol. Liq. 2017, 240, 322–328. [Google Scholar] [CrossRef]
  427. Prasad, C.; Sreenivasulu, K.; Gangadhara, S.; Venkateswarlu, P. Bio Inspired Green Synthesis of Ni/Fe3O4magnetic Nanoparticles Using Moringa Oleifera Leaves Extract: A Magnetically Recoverable Catalyst for Organic Dye Degradation in Aqueous Solution. J. Alloys Compd. 2017, 700, 252–258. [Google Scholar] [CrossRef]
  428. Singh, K.K.; Senapati, K.K.; Sarma, K.C. Synthesis of Superparamagnetic Fe3O4 Nanoparticles Coated with Green Tea Polyphenols and Their Use for Removal of Dye Pollutant from Aqueous Solution. J. Environ. Chem. Eng. 2017, 5, 2214–2221. [Google Scholar] [CrossRef]
  429. Khaghani, S.; Ghanbari, D.; Khaghani, S. Green Synthesis of Iron Oxide-Palladium Nanocomposites by Pepper Extract and Its Application in Removing of Colored Pollutants from Water. J. Nanostruct. 2017, 7, 175–182. [Google Scholar] [CrossRef]
  430. 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]
  431. Groiss, S.; Selvaraj, R.; Varadavenkatesan, T.; Vinayagam, R. Structural Characterization, Antibacterial and Catalytic Effect of Iron Oxide Nanoparticles Synthesised Using the Leaf Extract of Cynometra Ramiflora. J. Mol. Struct. 2017, 1128, 572–578. [Google Scholar] [CrossRef]
  432. Natarajan, E.; Ponnaiah, G.P. Optimization of Process Parameters for the Decolorization of Reactive Blue 235 Dye by Barium Alginate Immobilized Iron Nanoparticles Synthesized from Aluminum Industry Waste. Environ. Nanotechnol. Monit. Manag. 2017, 7, 73–88. [Google Scholar] [CrossRef]
  433. Lin, J.; Weng, X.; Dharmarajan, R.; Chen, Z. Characterization and Reactivity of Iron Based Nanoparticles Synthesized by Tea Extracts under Various Atmospheres. Chemosphere 2017, 169, 413–417. [Google Scholar] [CrossRef] [PubMed]
  434. Cheera, P.; Karlapudi, S.; Sellola, G.; Ponneri, V. A Facile Green Synthesis of Spherical Fe3O4 Magnetic Nanoparticles and Their Effect on Degradation of Methylene Blue in Aqueous Solution. J. Mol. Liq. 2016, 221, 993–998. [Google Scholar] [CrossRef]
  435. Xin, H.; Yang, X.; Liu, X.; Tang, X.; Weng, L.; Han, Y. Biosynthesis of Iron Nanoparticles Using Tie Guanyin Tea Extract for Degradation of Bromothymol Blue. J. Nanotechnol. 2016, 2016, 4059591. [Google Scholar] [CrossRef]
  436. Truskewycz, A.; Shukla, R.; Ball, A.S. Iron Nanoparticles Synthesized Using Green Tea Extracts for the Fenton-like Degradation of Concentrated Dye Mixtures at Elevated Temperatures. J. Environ. Chem. Eng. 2016, 4, 4409–4417. [Google Scholar] [CrossRef]
  437. Tan, K.A.; Morad, N.; Teng, T.T.; Norli, I. Synthesis of Magnetic Nanocomposites (AMMC-Fe3O4) for Cationic Dye Removal: Optimization, Kinetic, Isotherm, and Thermodynamics Analysis. J. Taiwan. Inst. Chem. Eng. 2015, 54, 96–108. [Google Scholar] [CrossRef]
  438. Wang, Z.; Fang, C.; Megharaj, M. Characterization of Iron-Polyphenol Nanoparticles Synthesized by Three Plant Extracts and Their Fenton Oxidation of Azo Dye. ACS Sustain. Chem. Eng. 2014, 2, 1022–1025. [Google Scholar] [CrossRef]
  439. Huang, L.; Weng, X.; Chen, Z.; Megharaj, M.; Naidu, R. Green Synthesis of Iron Nanoparticles by Various Tea Extracts: Comparative Study of the Reactivity. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2014, 130, 295–301. [Google Scholar] [CrossRef]
  440. Kimpiab, E.; Kapiamba, K.F.; Folifac, L.; Oyekola, O.; Petrik, L. Synthesis of Stabilized Iron Nanoparticles from Acid Mine Drainage and Rooibos Tea for Application as a Fenton-like Catalyst. ACS Omega 2022, 7, 24423–24431. [Google Scholar] [CrossRef]
  441. Sravanthi, K.; Ayodhya, D.; Yadgiri Swamy, P. Green Synthesis, Characterization of Biomaterial-Supported Zero-Valent Iron Nanoparticles for Contaminated Water Treatment. J. Anal. Sci. Technol. 2018, 9, 3. [Google Scholar] [CrossRef]
  442. Harshiny, M.; Iswarya, C.N.; Matheswaran, M. Biogenic Synthesis of Iron Nanoparticles Using Amaranthus Dubius Leaf Extract as a Reducing Agent. Powder Technol. 2015, 286, 744–749. [Google Scholar] [CrossRef]
  443. Parmakoğlu, E.Ü.; Çay, A.; Yanık, J. Valorization of Solid Wastes from Textile Industry as an Adsorbent Through Activated Carbon Production. AATCC J. Res. 2023, 10, 133–143. [Google Scholar] [CrossRef]
  444. Cho, E.J.; Lee, Y.G.; Song, Y.; Kim, H.Y.; Nguyen, D.T.; Bae, H.J. Converting Textile Waste into Value-Added Chemicals: An Integrated Bio-Refinery Process. Environ. Sci. Ecotechnol. 2023, 15, 100238. [Google Scholar] [CrossRef] [PubMed]
  445. Marazzi, F.; Fornaroli, R.; Clagnan, E.; Brusetti, L.; Ficara, E.; Bellucci, M.; Mezzanotte, V. Wastewater from Textile Digital Printing as a Substrate for Microalgal Growth and Valorization. Bioresour. Technol. 2023, 375, 128828. [Google Scholar] [CrossRef]
  446. Fiorentino, G.; Ripa, M.; Ulgiati, S. Chemicals from Biomass: Technological versus Environmental Feasibility. A Review. Biofuels Bioprod. Biorefin. 2017, 11, 195–214. [Google Scholar] [CrossRef]
Figure 1. The synthesis of nanomaterials via top-down, bottom-up, and mixed methods (top–bottom approaches) [53,54] (left). Green vs. gray synthesis processes (right).
Figure 1. The synthesis of nanomaterials via top-down, bottom-up, and mixed methods (top–bottom approaches) [53,54] (left). Green vs. gray synthesis processes (right).
Water 16 01607 g001
Figure 2. Theoretical relationship between particle size and surface area considering a non-porous sphere (a). Typical structures of zero-valent nanoparticles of iron (b,c).
Figure 2. Theoretical relationship between particle size and surface area considering a non-porous sphere (a). Typical structures of zero-valent nanoparticles of iron (b,c).
Water 16 01607 g002
Figure 3. Published papers (a), typical crystal size (b), and yield (c) obtained in each synthetic method. Source: Scopus.
Figure 3. Published papers (a), typical crystal size (b), and yield (c) obtained in each synthetic method. Source: Scopus.
Water 16 01607 g003
Figure 5. Relationship of removal efficiency with the particle size (left) and with the iron salt precursor (right). Source: Scopus and references cited in Table 5.
Figure 5. Relationship of removal efficiency with the particle size (left) and with the iron salt precursor (right). Source: Scopus and references cited in Table 5.
Water 16 01607 g005
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

Rodríguez-Rasero, C.; Montes-Jimenez, V.; Alexandre-Franco, M.F.; Fernández-González, C.; Píriz-Tercero, J.; Cuerda-Correa, E.M. Use of Zero-Valent Iron Nanoparticles (nZVIs) from Environmentally Friendly Synthesis for the Removal of Dyes from Water—A Review. Water 2024, 16, 1607. https://doi.org/10.3390/w16111607

AMA Style

Rodríguez-Rasero C, Montes-Jimenez V, Alexandre-Franco MF, Fernández-González C, Píriz-Tercero J, Cuerda-Correa EM. Use of Zero-Valent Iron Nanoparticles (nZVIs) from Environmentally Friendly Synthesis for the Removal of Dyes from Water—A Review. Water. 2024; 16(11):1607. https://doi.org/10.3390/w16111607

Chicago/Turabian Style

Rodríguez-Rasero, Cristina, Vicente Montes-Jimenez, María F. Alexandre-Franco, Carmen Fernández-González, Jesús Píriz-Tercero, and Eduardo Manuel Cuerda-Correa. 2024. "Use of Zero-Valent Iron Nanoparticles (nZVIs) from Environmentally Friendly Synthesis for the Removal of Dyes from Water—A Review" Water 16, no. 11: 1607. https://doi.org/10.3390/w16111607

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