Next Article in Journal
Formation of Thick Immersion Coatings and Residual Stress Evaluation in the System ZrB2–ZrO2: Experimental and Numerical Investigation
Next Article in Special Issue
Materials Make a Better Life: Functional Metals, Metal Oxides, and Metal Complexes
Previous Article in Journal
Review on Mechanoresponsive Smart Windows: Structures and Driving Modes
Previous Article in Special Issue
The Composites of PCL and Tetranuclear Titanium(IV)–Oxo Complex with Acetylsalicylate Ligands—Assessment of Their Biocompatibility and Antimicrobial Activity with the Correlation to EPR Spectroscopy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Application of Iron Nanoparticle-Based Materials in the Food Industry

by
Dariusz Góral
1,*,
Andrzej Marczuk
2,
Małgorzata Góral-Kowalczyk
2,
Iryna Koval
3 and
Dariusz Andrejko
1
1
Department of Biological Bases of Food and Feed Technologies, Faculty of Production Engineering, University of Life Sciences in Lublin, 20-612 Lublin, Poland
2
Department of Agricultural Forestry and Transport Machines, Faculty of Production Engineering, University of Life Sciences in Lublin, 20-950 Lublin, Poland
3
Department of Physical, Analytical and General Chemistry, Lviv Polytechnic National University, 79013 Lviv, Ukraine
*
Author to whom correspondence should be addressed.
Submission received: 24 November 2022 / Revised: 9 January 2023 / Accepted: 10 January 2023 / Published: 12 January 2023

Abstract

:
Due to their different properties compared to other materials, nanoparticles of iron and iron oxides are increasingly used in the food industry. Food technologists have especially paid attention to their ease of separation by magnetic fields and biocompatibility. Unfortunately, the consumption of increasing amounts of nanoparticles has raised concerns about their biotoxicity. Hence, knowledge about the applicability of iron nanoparticle-based materials in the food industry is needed not only among scientists, but also among all individuals who are involved in food production. The first part of this article describes typical methods of obtaining iron nanoparticles using chemical synthesis and so-called green chemistry. The second part of this article describes the use of iron nanoparticles and iron nanoparticle-based materials for active packaging, including the ability to eliminate oxygen and antimicrobial activity. Then, the possibilities of using the magnetic properties of iron nano-oxides for enzyme immobilization, food analysis, protein purification and mycotoxin and histamine removal from food are described. Other described applications of materials based on iron nanoparticles are the production of artificial enzymes, process control, food fortification and preserving food in a supercooled state. The third part of the article analyzes the biocompatibility of iron nanoparticles, their impact on the human body and the safety of their use.

1. Introduction

Nanotechnology is one of the relatively younger fields of science. Thus far, it has been mainly associated with medicine, including drug delivery and pharmacotherapy. The development of nanotechnology in food is progressing much faster than the development of nanomaterials and nanopharmaceuticals. Industrialists, scientists and researchers are meeting the needs of these sectors with nanotechnology solutions. Nanotechnology is used in many sectors of the food industry, from agricultural production (herbicides, fertilizers, pathogen analysis and detection and genetic engineering) to crop processing and food production (intelligent packaging, improved taste, improved texture or quality of food, new gelling or viscosity control agents, dietary supplements) [1]. The properties of many materials depend on the size of the particles at the nanoscale, and their structure results in changes in the coercive force in magnetic materials, surface reactivity, catalytic ability or a very important feature, such as mechanical strength. The surface effects of nanoparticles are relevant to structural issues [2]. Compared to other metallic NPs, iron nanoparticles are widely used. This is determined by such properties as their small size, superparamagnetism and biocompatibility. Iron nanoparticles are used in bioprocessing, tissue engineering and other aspects of modern medicine. There are different oxide forms of iron, such as hematite (α-Fe2O3), maghemite (γ-Fe2O3), FeOH (OH) goethite and magnetite (Fe3O4) [3,4]. The most commonly used crystalline iron oxide structures are maghemite (γ-Fe2O3) and magnetite (Fe3O4). They are used in many fields, including magnetic data storage, pigment production for electrochemistry, contrast production and many others [5]. The antimicrobial, anti-larvicide and antioxidant activities of some iron nanoparticles are used in packaging applications. Iron oxide nanoparticles (IONPs) exhibit a number of beneficial features, such as magnetic properties, high values of saturation magnetization and control by low-intensity magnetic fields, as well as their non-toxicity, ease of biodegradation and compatibility with biological structures that are used in food production processes. Furthermore, the exploitation of the iron-binding properties of the molecule has been used to eliminate hazardous substances from food—for example, citrinin or the nephrotoxic mycotoxin that is found during fermentation and is produced by the Monascus mold. In the literature, IONPs are classified into the following categories: ultrasmall superparamagnetic iron oxides (USPIOs), cross-linked iron oxides (CLIOs) and monocrystalline iron oxide nanoparticles (MIONs) [1]
IONPs are being used to develop new proposals for biomedicine, food safety and environmental protection, as well as solutions without the limitations and drawbacks of currently available techniques. Among others, user-friendly, sensitive, low-cost diagnostic tests have been developed to assist in the detection of animal pathogens and diseases [6]. In addition, iron oxide nanoparticles offer the ability to bind to ligands, creating effective delivery tools. Thanks to this property, magnetic iron nanoparticles can be used in many areas of the food and medical industries, particularly as gene and drug nanocarriers [1].
The current emphasis on the use of nanomaterials is well justified, given the enormous potential that they exhibit. At the same time, their use in the food industry should be closely monitored. Over the past few decades, iron nanoparticles have gained a significant advantage in many technologies due to the many benefits associated with their use, including simple synthesis, ease of application and easy recovery through the use of magnetic fields. However, iron nanoparticles also have some disadvantages, such as aggregation and easy oxidation. This paper focuses on summarizing the applications of iron nanoparticle-based materials in food production and analyzes the safety of their consumption by humans in depth. In doing so, we provide evidence of the key role that iron nanoparticles will play in broadly understanding food production and highlight the health risks to consumers.

2. Methods for Obtaining Iron Nanoparticles

2.1. Chemical Synthesis

Many methods are known for the chemical synthesis of iron nanoparticles. Methods such as the co-precipitation of an aqueous solution of iron (Fe2+) and iron (Fe3+) salts by addition of a weak base (for example, sodium hydroxide, ammonia, TMAOH), microemulsion techniques, hydrothermal synthesis, sonochemical methods, non-aqueous methods, the thermal decomposition of an organic iron precursor, sol–gel techniques, γ-ray synthesis, microwave plasma synthesis and the thermal decomposition of an alkaline Fe3+ chelate solution in the presence of a high-boiling-point solvent have been described in the literature [7,8] (Table 1).
The most common methods of chemical synthesis are thermal decomposition and chemical reduction. Thermal decomposition methods proceed at elevated temperatures by using water or organic solvents. The process can be carried out at atmospheric pressure [33] or under increased pressure [34]. The size of the nanoparticles can be controlled by varying the nucleation rate. At higher temperatures, nucleation is faster than crystallization, so smaller nanoparticles are formed under these conditions [7]. The use of hydrophilic polyhydroxy alcohols as solvents facilitates the subsequent dispersion of nanoparticles in water [8]. Homogeneous iron nanoparticles of small size with high purity can be obtained by this method. Unfortunately, this comes at a high cost, as it requires high temperatures and expensive precursors that are hazardous to health and the environment. For this reason, this method is not suitable for industrial applications where large quantities of IONs are needed (e.g., environmental clean-up or Ni-Fe battery production). The use of toxic organic solvents, such as diethyl ether, eliminates the possibility of using IONPs in biological processes and food production [9,35]. Comparing the two methods of thermal decomposition and chemical reduction, it should be noted that chemical reduction is more economical. The greatest advantage of chemical reduction is that the nanoparticles produced can be biocompatible. In a reduction reaction, an iron salt, such as FeCl3, is reduced in an aqueous solution environment, resulting in the formation of IONPs. In the literature, the most commonly described reducing substances are hydrazine hydrate (N2H4·XH2O) and sodium borohydride (NaBH4). Hydrazine hydrate is used in alkaline environments, while sodium borohydride is used in neutral or acidic environments [36]. The easiest to carry out and most commonly used is the co-precipitation method. It involves the co-precipitation of a mixture in appropriate proportions of an aqueous solution of iron(II) and iron(III) salts (e.g., FeCl3·6H2O, FeSO4·7H2O) under alkaline conditions and in the absence of oxygen; this leads to the formation of nanocrystals of magnetite or maghemite. The size of the nanoparticles is manipulated by changing the pH or salt concentration of the solution [37]. However, the size of the nanoparticles is very uneven [11]. IONPs are most often produced without surfactants or polymers dissolved in water. These agents are only used to prevent aggregation and agglomeration. They cause the iron nanoparticles to repel each other [9].

2.2. Green Synthesis

2.2.1. Synthesis by Microorganisms

Metallic and non-metallic nanoparticles are most often obtained by chemical processes carried out under rigorously controlled conditions. In contrast, the production of nanoparticles under biological conditions is characterized by such conditions of temperature, pH or pressure, which can be easily obtained both in the laboratory and in industry [2]. In living organisms, the intracellular or extracellular generation and deposition of inorganic crystalline structures is common. Examples include CaCO3 and Ca3(PO4)2 deposited in bacterial structures; gold and silver deposited in mycelia and bacteria; and silica, magnetite and greigite found in magnetotactic bacteria [8,37]. The potential of microorganisms is increasingly being exploited in laboratory conditions for the preparation of iron nanoparticles (Table 2).
The combination of microorganisms with metal ion precursors leads to products that possess great potential for use in nanoparticle synthesis.

2.2.2. Synthesis Using Plant Parts

The simplest, cheapest and most reproducible means to produce biocompatible IONPs is green synthesis. Different parts of plants (leaves, stems, seeds and roots) are used for this purpose (Table 3).
The ease of synthesis of nanoparticles in biological conditions is caused by the high content of organic reducing complexes contained in plants. Plants, compared to microorganisms, synthesize stable nanoparticles that are useful for rapid, large-scale synthesis [66].
Many studies have investigated green plant parts for the environmentally safe production of iron nanoparticles using precursors such as iron(II) chloride or iron(III) chloride solution. This has been shown to be possible due to the high content of bioactive components in plants. These active substances are mainly polyphenols, which have reducing and encapsulating properties. In addition, the use of a water environment and normal pressure and temperature eliminates the possibility of the formation of toxic substances, which in turn results in a lack of problems regarding the disposal of hazardous waste. For example, citrus and peanut peels and stalks can be used in nanoparticle production. Chemical and physical methods of producing nanomolecules require the usage of very reactive and toxic reducing agents, e.g., sodium borohydride and hydrazine hydrate, that cause unwanted effects to the environment, plants and animals. Such agents are not used in green synthesis [67].

2.2.3. Synthesis Using Green Reagents

There have been many studies that have used non-toxic, biocompatible materials to produce IONPs. Analyses have been conducted using biopolymers, ascorbic and amino acid, hemoglobin and myoglobin, sugar, glucose, synthetic tannic and gallic acid, among others [66]. Examples of green reagent applications are presented in Table 4.

3. Application of Iron Nanoparticle-Based Materials in Food Production

3.1. Application in the Production of Food Packaging

The use of nanomaterials in food packaging alters the penetration properties, improves the barrier properties, increases thermal resistance and leads to strong antimicrobial activity [77]. Nanoparticles are characterized by their large surface area and high surface energy. This creates strong interfacial interactions between the polymer bonds and the nanoparticles. This significantly improves the properties of the bio/polymers used for packaging [78].
One of the main problems during food storage is its contact with oxygen. This agent is responsible for many degradation reactions in food, such as lipid oxidation, microbial increase, enzymatic browning and nutritional degradation. Changes in food as a result of oxygen exposure result in a significantly reduced shelf life, especially for oxygen-sensitive foods [79]. In order to limit the access of oxygen to food, it is most often packaged using materials that provide an active barrier against oxygen penetration [80]. One way in which food can be protected from oxygen is through the use of iron nanoparticles. Two strategies are possible for the use of nano-iron in packaging. The first is to insert sachets containing nZVI into the packaging, and the second method is to incorporate the nanoparticles into the packaging film using a monolayer or multilayer structure [81,82]. An example is bentonite and kaolinite modified with Fe0. These substances have been approved by the European Food Safety Authority (EFSA) as non-nanoparticles [83]. Unlike normal iron, Fe0 reacts with oxygen under non-humid conditions [80]. Khalay et al. [84] investigated the suitability of PP nanocomposites having in their compositions montmorillonite (OMMT) with IONP modification for food packaging. They found that the nanoparticles were active as an active oxygen scavenger, and were able to capture and absorb oxygen by chemically reacting with it. Similarly, Busolo and Lagaron [82] reported that iron kaolinite contained in active composites is a passive barrier due to its winding pathway, which impedes gas diffusion and captures and reacts with molecular oxygen.
The most important problem during food storage is spoilage by microorganisms. The use of IONPs and nanocomposites with antimicrobial functions is an effective means to minimize the impact of microorganisms on food during processing or storage. In this way, they provide an extended shelf life and improved food safety [85]. Iron nanoparticles can be used to address food oxidation by applying antimicrobial coatings to the inner surface of the packaging. Such a coating gradually releases the antimicrobial substances, or they are immobilized on the active surface of the packaging [86]. Song et al. [87] produced composites of polydopamine with iron oxide nanoparticles (IONPs@pDA). Nisin was conjugated on the IONP@pDA nanoparticles. The newly synthesized material showed good efficiency in reducing Alicyclobacillus acidoterrestris, a food-spoilage-causing bacterium that is a serious problem in the food industry. Recent research has focused on producing IONP composites with surfaces modified with other nanometals. There is strong interest in silver-coated magnetic nanocomposites because of their unique antimicrobial properties [88].

3.2. Edible Coatings on Food

The use of films and coatings produced from natural edible polymers (biopolymers) on food is increasingly being used to extend the shelf life of fresh produce. These films and coatings have similar functions to conventional protective packaging or synthetic coatings [89]. The use of food-approved nanoparticles has enabled the creation of functional edible coatings. These coatings consist of nanoemulsions, polymer nanoparticles, nanofibers, solid lipid nanoparticles (SLNs), nanostructured lipid carriers (NLCs) and traditional polymer nanocomposites (inorganic–organic) [90]. One of the most commonly used ingredients in edible films is chitosan. Chitosan is a natural polymer composed of (1,4)-linked 2-amino-deoxy-b-D-glucan. It has been shown in many publications that chitosan is non-toxic, biodegradable, biofunctional, biocompatible and has antimicrobial properties. Furthermore, due to its ability to form transparent films and the good mechanical properties of films that can meet various packaging needs, chitosan is an excellent ingredient for edible packaging [91]. Combining chitosan with nanoparticles seems to be an excellent solution. Nehra et al. [92] reported that chitosan coated with Fe3O4 nanoparticles is an antimicrobial agent against E. coli, B. subtilis, Candida albicans, Aspergillus niger and Fusarium solani. Shrifian-Esfahni et al. [93] found that under in vitro conditions, a chitosan–Fe3O4 nanoparticle film showed inhibitory activity against E. coli and P. aeruginosa. SPIONs composed of magnetite (Fe3O4) are characterized by such properties, which can be used for antimicrobial applications [94], especially since the U.S. Food and Drug Administration has found that SPIONPs are biocompatible (BC) with the human body [95].

3.3. Immobilization of Enzymes

Enzymes are used as biocatalysts for large-scale production in many areas of industry. Their great advantage is their limited environmental impact. In the food industry, they make it possible to modify the structures of foods. For example, this is used in the production of lactose-free milk and calorie-reduced fats [96]. Placing enzymes on magnetic nanoparticle carriers ensures a strong covalent bond between the enzyme and the carrier. By doing so, it minimizes the leaching of the enzyme in the aqueous environment and prevents the contamination of the product with other proteins. An important advantage of this solution is that a magnetic field can be used to separate the biocatalyst from the reactants. This protects against particle losses and provides high mechanical strength, thermal stability and resistance to chemical and microbial degradation [97]. Another area where iron SPIONs have found application is in protein/enzyme immobilization. Here, the characteristic magnetic properties of SPIONs and the ease of separation in a magnetic field have been exploited. However, the most important factor is the ability to orient proteins/enzymes in the carriers; this protects against internal diffusion, which is a problem in traditional methods [98].

3.4. Artificial Enzymes

Nanomaterials that exhibit enzyme-like activity are proposed as a new generation of artificial enzymes and referred to as “nanozymes” [99]. The main advantage of nanozymes over their natural counterparts is the ability to regulate activity by changing the size or morphology, or through doping or surface modification. The most typical nanozymes are iron oxide nanoparticles (Fe3O4 and Fe2O3). They mimic the activity of peroxidase and catalase [100]. Peroxidase-like activity is possible under acidic conditions. Ferromagnetic nanoparticles can catalyze the reaction of hydrogen peroxide with hydroxyl radicals. This means that IONPs can be referred to as IONzymes. As reported by Qin et al. [101], IONzymes are antiviral agents because they induce the peroxidation of membrane lipids in synthesized liposomes. Their large surface area enables them to exert strong catalytic activity. This activity matches and sometimes exceeds that of natural enzymes. IONzymes are more stable and less expensive than natural enzymes. In addition, proteolytic degradation and environmental factors (temperature, pH, ionic strength and heavy metals) are no longer an issue. One of the most important features is that they do not degrade during storage at room temperature. The disadvantages of IONzymes are that they only bind to a specific substrate (substrate specificity) and show different catalytic activity, which is determined by the size and structure [102].

3.5. Food Analysis

Iron nanoparticles have been widely used in food analysis methods [103]. Magnetic nanoparticles are usually integrated into detection techniques in two ways: as an electrode modifier and as a sample preconcentrator [98]. An example is the method based on magnetic microparticles coated with protein A (a component of the cell wall produced by several strains of Staphylococcus aureus). This has been used for the accurate (0.2 mg/kg) detection of the allergen Ara h3 (a heat-stable protein that comprises 19% of the total protein in peanut extracts) in food. The method has high repeatability and reproducibility [104]. Another example is the use of magnetic nanoparticles in glucose biosensors. The biosensors use immobilized oxidase to convert the target analytes into electrochemically detectable products. This makes the determination of glucose in food easier [105]. Magnetite nanoparticles enable food analysis for the detection of certain elements (e.g., silver, copper, lead, cadmium and mercury). Magnetite nanoparticles are coated with silanes. This allows them to be functionalized with ligands that are selective for metal ions [103] (Table 5).
In a study by Yang et al. [117], it was shown that SPIONs can accelerate the preconcentration process for the detection of bacteria in food. Target bacteria used for PCR detection are concentrated using submicron superparamagnetic anion exchangers. This enabled the PCR detection limit to be shifted from 104–105 CFU mL−1 to 102 CFU mL−1. Escherichia coli and Agrobacterium tumefaciens were used as models of bacteria. Wen et al. [115] described a one-step assay for the detection of Salmonella typhimurium by fluorescent identification using IONPs with functional additives. The work of Chen et al. in [118] used silica-coated SPIONs to extract DNA from pathogenic bacteria. In this method, preincubation of the sample was avoided and the influence of the food matrix was eliminated.

3.6. Protein Purification

One increasingly used process in the food industry is the purification of proteins from isolates or concentrates. The resulting proteins can then provide the basis for new forms of food. For example, milk proteins fall into two distinct categories: casein micelles, which are often associated with cheese production, and serum proteins, which are involved in many functional properties, such as foaming, gelling and emulsification. In high-tech food production, it is necessary to use only one type of protein. Therefore, there is a need for a low-cost and efficient technique to obtain pure proteins [119]. Traditionally, proteins are purified by chromatography, precipitation, ultrafiltration, centrifugation and dialysis. The inherent limitations of these methods are the long pretreatment times, equipment that is often expensive to purchase and operate and the need to hire experienced lab technicians. Magnetic separation, which has been known for many years, can be used for this purpose. It is characterized by several advantages. The technique using SPIONs is fast, scalable, sensitive and easy to automate and no special sample preparation is needed (Figure 1).
SPIONs used for protein purification are immobilized with ligands. Primarily, these are ligands with confirmed affinity for the proteins to be purified [105].
Another example is the use of SPIONs for the analysis and separation of hormones, drugs and contaminants from food. This technique is used in the study of food of animal origin. Traces of drugs may be present in this type of product. Such contaminated products, when consumed by humans, endanger their health [103]. Estrogen hormones are administered as feed additives (called growth promoters) to accelerate growth and increase meat weight in cattle, poultry and pigs [120]. As reported by Gao et al. [100], functionalized Fe3O4 nanoparticles have been successfully used as sorbents for the specific separation and enrichment of proteins from milk. The authors report that they were able to recover estradiol in the range of 88.9 to 92.1%. Luo et al. [121] used carbon nanotubes with SPIONs because of their excellent adsorption capacity against hydrophobic compounds. They were used as adsorbents in the extraction of phthalate acid esters. Esters such as phthalates and bisphenol A are derived from plastic packaging and are considered as substances with potential carcinogenic properties.

3.7. Iron Oxides as Ingredients in Foods and Dietary Supplements

The most important biological function of iron is oxygen transport, as it is part of the heme nucleus in the proteins hemoglobin and myoglobin. In living organisms, iron is involved in more than 200 enzyme systems that are essential for cellular function, particularly cellular energy utilization and DNA, RNA and protein synthesis. In addition, iron is involved in neurotransmitter metabolism, vitamin D activation, collagen metabolism and cholesterol catabolism. Therefore, iron is essential for the proper functioning of organisms, including humans [122]. According to Dave and Gao [33], ferritin is an intracellular protein found in the cells of the liver and immune system. It stores inactive iron in the body and releases it when needed. As reported by Kumari and Chauhan [123], commonly used iron enrichment additives with high bioavailability (FeSO4) are chemically reactive. These reactions cause color changes, a metallic aftertaste, rancidity in food products, etc. Due to these difficulties, FeSO4 is replaced by water-insoluble compounds FePO4, Fe4(P2O7)3 or ions of iron. These additives do not affect the properties of food but are poorly absorbed by the human body [124]. Therefore, IONPs are an alternative to traditional iron supplements, preventing iron-deficiency-related diseases [125], as similarly stated by von Moos et al. [124] and Perfecto et al. [126], using FePO4 NPs as an iron source. In addition to iron oxides and phosphates, Hilty et al. [127] used a Ca and Mg complex in iron oxide compounds. These minerals are essential for humans, so adding them to food poses no health risk. The authors showed that the addition of calcium and magnesium to Fe2O3 gives i.d.a. solubility > 85%, suggesting a high level of bioavailability. In addition, no caking was found during storage under various humidity conditions.

3.8. Colorants

Iron, in the form of various compounds, is used as a colorant (food additive E172). A range of colors can be obtained, such as black, red, yellow, blue, orange and brown, depending on the chemical composition. For example, magnetite (Fe3O4), with a particle size of 11.7 nm, is characterized by its black color. The red pigment is associated with hematite (α-Fe2O3) nanoparticles with a particle size of 39.5 nm, and the yellow pigment is attributed to goethite (α-FeOOH) nanoparticles with a particle size of 48.7 nm [128]. Iron oxides are approved food pigments and have been shown to contain significant amounts of nanoparticles [129].

3.9. Mycotoxin Removal

Mycotoxins are produced by filamentous fungi belonging to the Ascomycota cluster or molds and are toxic secondary metabolites. Their characteristic feature is their low molecular weight. The source of mycotoxins is usually infected food products. They are often carcinogenic and mutagenic; among other aspects, they inhibit DNA synthesis and cause changes in RNA metabolism [130]. During various food processing technologies, including cooking, baking, frying, roasting and pasteurization, most mycotoxins remain chemically and thermally stable. Therefore, low-cost methods to eliminate these contaminants from food are still being sought [131]. Iron oxide and graphene oxide nanostructures can be cheaply produced and are readily available. Mycotoxins interact with the surface oxygen functional groups of the nanostructures. Surface-active maghemite nanoparticles (γ-Fe2O3) have shown chelating properties for citrulline and ochratoxin A in the presence of iron(III). Another promising solution is IONP@chitosan complexes as patulin adsorbents. As reported Horky et al. [132], after approximately 5 h, the patulin molecules were completely adsorbed. The adsorbent concentration was 400 μg, with a pH equal to the pH of apple juice. After adsorption, mycotoxins along with SPIONPs were removed in a magnetic field.

3.10. Anti-Allergic Effect

Histamine is found in some foods. In food, histamine is formed during storage as a result of the activity of bacteria—not only those intentionally added, but also those that contaminate it. Histamine content is considered one of the markers of food quality [133]. This compound is one of the most important intermediates in the course of allergic reactions. Consumption of histamine-rich foods can cause nausea, headaches, diarrhea and asthma [134]. According to Ghanbari, Adivi and Hashemi [134], magnetic Fe3O4@ agarose/IDA@silica nanoparticles can be used to remove histamine from biological fluids. The histamine removal efficiency of real plasma samples was 92%.

3.11. Control of the Process Flow

Bacillus subtilis is used for the industrial production of a wide range of compounds in the fermentation process. One of the main problems of industrial fermentation is the formation of biofilms by this bacterium, which leads to many process and operational complications. Biofilms are an unfavorable phenomenon in industrial settings. Periodic cleaning of equipment is necessary. The use of SPIONs coated with aminopropyltriethoxysilane is a promising approach to control biofilm formation without losing the viability of B. subtilis cells [135].

3.12. Preservation of Food in Supercooled State

Iron nanocrystals (Fe3O4) are a natural component of meat. They provide strong nucleation sites for heterogeneous ice crystallization. The ferromagnetism of magnetite nanoparticles allows an external oscillating magnetic field to induce magneto-mechanical motions. These movements can interfere with the formation of ice crystals in food by inducing a supercooled state. For this to occur, the magneto-mechanical rotation of the particles should be greater than the magnitude of Brownian motion [136]. A study by Kang at al. [137] found that fresh beef stored under refrigeration at −4 °C using an oscillating magnetic field was preserved for one week longer, without significant changes in weight loss, color or bacterial growth.

4. Effects of Iron Nanoparticles on Living Organisms

4.1. Antimicrobial Activity

A number of papers have shown that certain metal nanoparticles exhibit antimicrobial activity against a variety of Gram-negative and Gram-positive bacteria [8] (Table 6).
The most common action of NPs is to damage bacterial cells by destroying the cell membrane, damaging organelles, biomolecular distortion and by interfering with the biosynthesis of nucleic acids and proteins [8].
Damage to nucleic acids and proteins in a bacterial cell is caused by superoxide radicals (O2), hydroxyl radicals (-OH), hydrogen peroxide (H2O2) and singlet oxygen (O2). These are referred to as reactive oxygen species (ROS) [149]. Proteins including enzymes are inactivated or the reaction is caused to slow down by partial inhibition with metal ions capable of binding to mecapto (-SH), amine (-NH) and carboxyl (-COOH) groups. In addition, IONPs are able to interfere with F0F1-ATPase/ATP synthase function and reduce the rate of H+ flow across the membrane and the redox potential [150]. Lee et al. [10] showed that Fe0 nanoparticles bind to intracellular oxygen. This phenomenon causes oxidative stress and the breakdown of the cell membrane. In addition, they found that the strong antibacterial activity allows the use of nano-Fe0 as a biocide in many applications, also in combination with silver nanoparticles. The main problem is the strong oxidation of nZVI, which results in a reduction in their bactericidal activity, thus limiting their application. Therefore, work is needed to obtain nanoparticles that are resistant to oxidative corrosion.
Another means of exploiting the antibacterial properties of iron nanoparticles is through the catalytic reaction of Fe3O4 with H2O2. The reaction produces free radicals (intermediate products). These radicals are very toxic to microorganism cells, because they attack membranes, proteins and nucleic acids and ultimately cause the bacteria to malfunction [100]. Gabrielyan et al. [140] studied the impact of Fe3O4 on selected strains of Gram-positive and Gram-negative bacteria. They found that Fe3O4 was more effective against Gram-negative E. coli than Gram-positive E. hirae. The results prove that membrane permeability can be modified in the presence of Fe3O4.
Another use for iron nanoparticles is their ability to induce hyperthermia, a state of elevated temperature caused by an alternating magnetic field [16]. Superparamagnetic iron oxide is used to inactivate bacteria by creating elevated temperatures [142]. According to Rodrigues et al. [143], hyperthermia allowed the complete elimination of P. fluorescens at the temperature of 50–55 °C.

4.2. Antiviral Effect of IONPs

The antiviral activities of NPs occur both inside and outside cells. They involve NPs interacting with gp120-type proteins, competing with the virus for binding sites to the host cell, inferring viral attachment and preventing binding or blocking viral penetration into the host. Further mechanisms of antiviral action include acting on viral particles, disrupting the viral genome or binding to viral particles, and thereby preventing viral replication [151]. For example, Fe3O4 nanoparticles inhibit DEN-2 virus replication by inhibiting E protein expression [141]. Another action of iron NPs is to bind to the virus to prevent it from binding to cells.

4.3. Antifungal Activity

Iron oxide nanomaterials can potentially be used against many fungal pathogens (Table 6).
The small size of the nanoparticles and their greater surface area to volume ratio is the main mechanism of antimicrobial activity. This mechanism effectively weakens the microorganisms and reduces the supply of oxygen for respiration [146]. In addition, the same effects as in bacterial cells are displayed; i.e., IONPs are able to depolarize the cell membrane to induce the production of ROS and generate oxidative stress that disrupts cellular homeostasis [152].

4.4. Effects of Iron Nanoparticles on the Human Body

Iron nanoparticles can be toxic. Their intracellular and in vivo toxicity results from the production of excess reactive oxygen species (ROS), including free radicals such as superoxide anion, hydroxyl radical and non-radical hydrogen peroxide. High levels of free radicals can damage cells by peroxidizing lipids, damaging DNA, modulating gene transcription and altering proteins, and can cause impaired physiological functions and cell apoptosis. There are four main sources of oxidative stress in response to iron oxide nanoparticles (Figure 2):
  • Direct generation of ROS on the surfaces of nanoparticles;
  • Production of ROS via the leaching of iron particles;
  • Altered function of mitochondria and other organelles;
  • Induction of cellular signaling pathways [139].

4.5. In Vitro Evaluation

Cell viability analysis, proliferation, differentiation using atomic force microscopy and gene expression analysis are commonly used in vitro toxicity testing techniques [153,154]. An initial assessment of biocompatibility is achieved by examining all of these parameters. This is important in determining the interactions of nanoparticles with the cell membrane, aggregation and the presence of physiological effects. Evaluating the safety of nanoparticles in vitro is easy and inexpensive. Nanoparticles can affect the proper functioning of cellular structures (e.g., ROS generated by nanoparticles increase the activity of mitochondrial enzymes, reactions between some dyes used in experiments and nanoparticles), which should be taken into account when conducting experiments [155].

4.6. In Vivo Evaluation

The interactions of magnetic nanoparticles in biological systems are highly dynamic and complex. Investigation of the influence of nanoparticles on cellular function is important for many reasons. An important phenomenon is how IONPs affect physiological iron metabolism after they are degraded in the body. The accumulation of intracellular iron can affect the denaturation of proteins and nucleic acids [156].

4.7. Biocompatibility

Biocompatibility primarily depends on the degree of observed cytotoxicity. However, there are no clearly defined criteria for NP cytotoxicity. In recent studies, it has been observed that they can cause DNA damage, mitochondrial membrane dysfunction, oxidative stress and changes in gene expression. However, cytotoxicity associated with the above problems was not observed [157]. Biocompatibility is a requirement in classifying nanoparticles including iron for biomedical applications. Confirmation of biocompatibility is achieved by performing in vitro and in vivo analyses. Cytotoxicity investigations mainly include cell viability/proliferation/differentiation tests, microscopic analysis and intracellular localization, hemolysis and genotoxicity tests. In order to provide information on the effects of nanoparticles on living organisms, including blood compatibility, biodistribution, metabolism and clearance, studies should be performed both in vitro and in vivo [158]. It is widely believed that iron nanoparticles are the most biocompatible among other nanosubstances. However, in scientific works, one can find different opinions on this issue. In general, it has been shown that the toxicity of nanoparticles depends on their size, surface properties, concentration, shape and structure [159,160]. For this purpose, a number of in vitro biocompatibility studies of nanoparticles of various shapes (including nanooctahedra, nanorods, nanocubes, nanohexagons, nanodrills, nanotubes, nanobeads, nanobeads, nanorods, nanorings and nanocapsules) were carried out. Zhou et al. [161] conducted in vitro analyses of differentiated IONPs (nanooctahedrons, nanorods and nanocubes) in A549 lung cancer cells. More than 90% of the cells survived after 24 h of incubation. This proved that the nanoparticles were safe for cells in a given range (10–1000 μg mL−1). Chu et al. [162] studied the effects of IONPs of different shapes coated with poly(ethylene glycol)-phospholipid (DSPE-PEG) on ECA-109 cells. Laser irradiation of nanoparticles caused negligible toxicity to cells. On the other hand, Fan et al. [163] analyzed the cytotoxicity of magnetite nanoparticles. The viability results showed that ferrimagnetic IONP nanorods exhibited negligible toxicity at low IONP concentrations (50 μg mL−1 of Fe3O4), after 24 h of incubation. According to these authors, the shape of the nanoparticles is not as important as the concentration. The concentration of iron nanoparticles is a key factor in assessing cell viability during incubation. Ankamwar et al. [159] observed the cytotoxicity of IONPs at concentrations above 100 μg mL−1. Nanoparticle size is an important factor in biocompatibility. To increase the filtration time through the spleen and liver, nanoparticles should be small in size (<200 nm). On the other hand, to avoid rapid flow through the kidneys, the size should be larger than 10 nm [164]. An important property of nanoparticles that affects the human body is their surface area. Uncoated IONPs tend to be more toxic in vitro as well as in vivo compared to encapsulated nanoparticles. In medical applications, surface coating of IONPs is often used. In addition, it can protect against the agglomeration of IONPs and improve dispersion, support chemical drug combination modes and reduce non-specific effects on cells, providing lower toxicity and higher biocompatibility [165]. For obvious reasons, the coating materials used should have high biocompatibility and affinity for IONPs and should not affect the immune system or be antigenic. The most commonly used coatings are composed of various materials, both organic and inorganic [166]. One type of material used to protect iron nanoparticles is polymers, such as polyethylene glycol (PEG), which has been approved by the US Food and Drug Administration (FDA) and is considered a safe and biocompatible material. The advantage of this material that it is easily removed by the kidneys (for PEG < 30 kDa) or the digestive system (for PEG > 20 kDa). In addition, the adhesion of PEG particles to the surfaces of iron nanoparticles allows for an extended residence time in the bloodstream [167]. Other materials used to enhance the biocompatibility of IONPs include polyvinyl alcohol (PVA), polyethyleneimine (PEI), polyacrylic acid (PAA), poly(lactide-co-glycolide) (PLGA), dextran and protein molecules [138,168,169,170,171,172]. Disadvantages associated with the use of polymers include the in vivo aggregation of nanoparticles and the ease of separation of the colloidal suspension [159]. In a study by Han and Zhou [173], the enhanced biocompatibility of liposome-modified surface nanoparticles was demonstrated over an appropriate concentration range.

4.8. Effects on the Human Body

Iron nanoparticles have broad effects on the human body. Fe2O3 NPs have been found to protect the heart from ischemia at the tissue and cellular levels in an ischemia–reperfusion model. The therapeutic effect on iron deficiency occurs at doses of 3 mg to 6 mg per kilogram of body weight per day. Exceeding the recommended standard (10–20 mg/kg body weight) causes adverse effects [86]. One of the potential mechanisms of IONPs is to inhibit intracellular ROS and reduce peroxidative damage [174]. More interestingly, Zhang et al. [175] described that the consumption of IONPs can result in slower aging and prevent neurodegeneration. Feitosa et al. [176] found that oxidative stress causes aging associated with slowed cellular metabolism by measuring the clonogenic potential of LA-9 fibroblasts at day 7 after exposure to Fe3O4 NPs. Similarly, Fe3O4 nanoparticles modified with poly(L-lysine) can promote the proliferation of cancer stem cells from U251 glioblastoma multiforme via a reduction in the level of intracellular H2O2 [177]. In studies on aggregation into larger clusters of protein amyloid (Parkinson’s disease, Alzheimer’s disease and type II diabetes), it was shown that the amount of protein aggregates was decreased by the use of Fe3O4. IONPs inhibited the aggregation of lysozyme amyloid by blocking nucleation and caused the depolymerization of lysozyme clusters by interacting with adjacent protein sheets and breaking them down [178]. High concentrations of iron in the body can be toxic and lead to abnormalities in tissue homeostasis. It can cause oxidative stress, cytotoxicity, DNA damage, epigenetic events and inflammatory processes. Even if the high iron concentration of iron ions does not cause cytotoxicity, disturbances in cellular structures may occur, initiating carcinogenesis. Some studies have shown that iron may be responsible for the occurrence of cancer formation by, for example, generating ROS, which can damage DNA and proteins and cause lipid peroxidation [157]. However, numerous works confirm that IONPs are safe and non-cytotoxic at concentrations up to 100 μg mL−1 [156,159,179]. Before administering nanoparticles, there are different paths for their absorption, such as digestion, inhalation or dermal absorption.

5. Summary

Iron nanoparticles are increasingly used in food applications. This is influenced by their low manufacturing cost, the possibility of using their properties in a wide range and their biocompatibility. This review highlights the most important directions of nanoparticle use in the food industry. This is influenced by their high reactivity with oxygen, both in the presence of moisture and in anhydrous environments, and their ability to counteract the growth of microorganisms belonging to different groups in terms of morphological structure (Gram-negative and Gram-positive bacteria, yeasts and molds). The use of iron nanoparticles in the form of ZVI is particularly important. Iron nanoparticles, due to their high surface-area-to-size ratio, have become versatile tools in terms of creating carriers to stabilize enzymes. Enzymes that have been previously immobilized have greater resistance to changes in environmental conditions (pH, temperature) than enzymes not treated with this technique. In addition, they can be reused because of their ease of recovery. There are also disadvantages to this technique. The activity of some enzymes decreases, indicating the need for further research on this technique. The results obtained in the separation or purification of proteins from solutions through the use of SPIONs indicate that it is possible to transfer this technique to the removal of some proteins from food. Due to the magnetic properties of SPIONs, it is also possible to eliminate many adverse components, both pathogenic and allergenic, from food. Magnetic SPIONPs integrated with PCR, immunoenzymes, spectrometry and biosensors are fast and accurate methods for the detection of certain ingredients or contaminants in food, and they are helpful in identifying and quantifying microorganisms. The use of SPIONs increases the sensitivity and speed of analysis. It can also reduce the cost of manufacturing equipment. Another promising application of SPIONs is that they can be used to control the ice nucleation process in the freezing and storing of food.
Another direction of application for iron nanoparticle-based materials in the food industry is the fortification of food products. Iron at nano-size has high bioavailability, and problems such as an unacceptable color, taste, metallic aftertaste and rancidity in foods due to the use of water-soluble iron salts have been solved.
On the other hand, many papers have been written on the possible toxic effects of IONPs on the human body. The small size of the nanoparticles allows them to pass through physiological barriers, by which they can adversely affect cells or tissues, consequently leading to health risks. The coating of IONPs and their size are key to inducing toxicity mechanisms. Therefore, in order to safely use iron nanoparticle-based materials in the food industry, the safety and effects of IONPs on the human body must always be evaluated.
In conclusion, iron nanoparticles can be successfully used for the production of packaging that comes into contact with food, as well as the food itself. Smaller dimensions mean a larger external surface, which in turn improves water absorption, the release of aromas, the availability of bioactive substances and the speed of catalytic processes. Nanotechnology can be used in many areas of food production; for some, however, the methods are impractical and costly at an industrial scale. In addition, certain limitations (e.g., concentrations up to 100 μg mL−1) related to their use must be taken into account.

Funding

This research was funded by the state budget under the program of the Ministry of Education and Science (Republic of Poland) under the name Excellent Science—Support for Scientific Conferences, entitled “XXIII Polish Nationwide Scientific Conference: PROGRESS IN PRODUCTION ENGINEERING 2023”; project number: DNK/SP/546290/2022, amount of funding: 162,650.00 PLN, total value of the project: 238,650.00 PLN (Poland).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

NPs, nanoparticles; IONPs, iron oxide nanoparticles; USPIO, ultra-small superparamagnetic iron oxide; CLIO, cross-linked iron oxide; MIONs, monocrystalline iron oxide; TMAOH, tetramethylammonium hydroxide; MNPs, magnetic nanoparticles; SPIONPs, superparamagnetic iron oxide nanoparticles; IONzymes, iron oxide nanozymes; nZVI, zero-valent iron nanoparticles; PACFeNPs, proanthocyanidin iron nanoparticles.

References

  1. Magro, M.; Vianello, F. Bare Iron Oxide Nanoparticles: Surface Tunability for Biomedical, Sensing and Environmental Applications. Nanomaterials 2019, 9, 1608. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Kamran, U.; Bhatti, H.N.; Iqbal, M.; Nazir, A. Green Synthesis of Metal Nanoparticles and their Applications in Different Fields: A Review. Z. Fur Phys. Chem. 2019, 233, 1325–1349. [Google Scholar] [CrossRef]
  3. Can, M.M.; Coşkun, M.; Fırat, T. A comparative study of nanosized iron oxide particles; magnetite (Fe3O4), maghemite (γ-Fe2O3) and hematite (α-Fe2O3), using ferromagnetic resonance. J. Alloy. Compd. 2012, 542, 241–247. [Google Scholar] [CrossRef]
  4. Nieuwoudt, M.K.; Comins, J.D.; Cukrowski, I. The growth of the passive film on iron in 0.05 M NaOH studied in situ by Raman microspectroscopy and electrochemical polarization. Part II: In situ Raman spectra of the passive film surface during growth by electrochemical polarization. J. Raman Spectrosc. 2011, 42, 1353–1365. [Google Scholar] [CrossRef]
  5. Tran, P.; Tran, T.T.-D.; Van Vo, T.; Lee, B.-J. Promising iron oxide-based magnetic nanoparticles in biomedical engineering. Arch. Pharmacal Res. 2012, 35, 2045–2061. [Google Scholar] [CrossRef] [PubMed]
  6. Colino, C.I.; Millán, C.G.; Lanao, J.M. Nanoparticles for Signaling in Biodiagnosis and Treatment of Infectious Diseases. Int. J. Mol. Sci. 2018, 19, 1627. [Google Scholar] [CrossRef] [Green Version]
  7. Senthil, M.; Ramesh, C. Biogenic Synthesis of Fe3O4 Nanoparticles Using Tridax Procumbens Leaf Extract and Its Antibacterial Activity on Pseudomonas Aeruginosa. Dig. J. Nanomater. Biostruct. 2012, 7, 1655–1661. [Google Scholar]
  8. Góral, D.; Góral-Kowalczyk, M. Application of Metal Nanoparticles for Production of Self-Sterilizing Coatings. Coatings 2022, 12, 480. [Google Scholar] [CrossRef]
  9. Huang, K.-C.; Ehrman, S.H. Synthesis of Iron Nanoparticles via Chemical Reduction with Palladium Ion Seeds. Langmuir 2007, 23, 1419–1426. [Google Scholar] [CrossRef]
  10. Lee, C.; Kim, J.Y.; Lee, W.I.; Nelson, K.L.; Yoon, J.; Sedlak, D.L. Bactericidal Effect of Zero-Valent Iron Nanoparticles on Escherichia coli. Environ. Sci. Technol. 2008, 42, 4927–4933. [Google Scholar] [CrossRef] [Green Version]
  11. Leszczyński, B. Nanocząstki Tlenków Żelaza w Hipertermii Magnetycznej; Uniwersytet Adama Mickiewicza w Poznaniu: Poznań, Poland, 2016. [Google Scholar]
  12. Kandpal, N.D.; Sah, N.; Loshali, R.; Joshi, R.; Prasad, J. Co-Precipitation Method of Synthesis and Characterization of Iron Oxide Nanoparticles. J. Sci. Ind. Res. 2014, 73, 87–90. [Google Scholar]
  13. Petcharoen, K.; Sirivat, A. Synthesis and characterization of magnetite nanoparticles via the chemical co-precipitation method. Mater. Sci. Eng. B 2012, 177, 421–427. [Google Scholar] [CrossRef]
  14. Hui, B.H.; Salimi, M.N. Production of Iron Oxide Nanoparticles by Co-Precipitation method with Optimization Studies of Processing Temperature, pH and Stirring Rate. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2020; Volume 743, p. 012036. [Google Scholar]
  15. Yoon, S. Preparation and Physical Characterizations of Superparamagnetic Maghemite Nanoparticles. J. Magn. 2014, 19, 323–326. [Google Scholar] [CrossRef] [Green Version]
  16. Winiarczyk, K.; Gac, W.; Góral-Kowalczyk, M.; Surowiec, Z. Magnetic properties of iron oxide nanoparticles with a DMSA-modified surface. Hyperfine Interact. 2021, 242, 5–10. [Google Scholar] [CrossRef]
  17. Chawla, P.; Najda, A.; Bains, A.; Nurzyńska-Wierdak, R.; Kaushik, R.; Tosif, M. Potential of Gum Arabic Functionalized Iron Hydroxide Nanoparticles Embedded Cellulose Paper for Packaging of Paneer. Nanomaterials 2021, 11, 1308. [Google Scholar] [CrossRef]
  18. Nayeem, J.; Al-Bari, A.A.; Mahiuddin; Rahman, A.; Mefford, O.T.; Ahmad, H.; Rahman, M. Silica coating of iron oxide magnetic nanoparticles by reverse microemulsion method and their functionalization with cationic polymer P(NIPAm-co-AMPTMA) for antibacterial vancomycin immobilization. Colloids Surf. A Physicochem. Eng. Asp. 2021, 611, 125857. [Google Scholar] [CrossRef]
  19. Chin, A.B.; Yaacob, I.I. Synthesis and characterization of magnetic iron oxide nanoparticles via w/o microemulsion and Massart’s procedure. J. Mater. Process. Technol. 2007, 191, 235–237. [Google Scholar] [CrossRef]
  20. Kekalo, K.; Koo, K.; Zeitchick, E.; Baker, I. Microemulsion Synthesis of Iron Core/Iron Oxide Shell Magnetic Nanoparticles and Their Physicochemical Properties. Mater. Res. Soc. Symp. Proc. 2012, 1416, 61–66. [Google Scholar] [CrossRef] [Green Version]
  21. Amara, D.; Felner, I.; Nowik, I.; Margel, S. Synthesis and characterization of Fe and Fe3O4 nanoparticles by thermal decomposition of triiron dodecacarbonyl. Colloids Surf. A Physicochem. Eng. Asp. 2009, 339, 106–110. [Google Scholar] [CrossRef]
  22. Hufschmid, R.; Arami, H.; Ferguson, R.M.; Gonzales, M.; Teeman, E.; Brush, L.N.; Browning, N.D.; Krishnan, K.M. Synthesis of phase-pure and monodisperse iron oxide nanoparticles by thermal decomposition. Nanoscale 2015, 7, 11142–11154. [Google Scholar] [CrossRef] [Green Version]
  23. Maity, D.; Kale, S.; Kaul-Ghanekar, R.; Xue, J.-M.; Ding, J. Studies of magnetite nanoparticles synthesized by thermal decomposition of iron (III) acetylacetonate in tri(ethylene glycol). J. Magn. Magn. Mater. 2009, 321, 3093–3098. [Google Scholar] [CrossRef]
  24. Cotin, G.; Kiefer, C.; Perton, F.; Ihiawakrim, D.; Blanco-Andujar, C.; Moldovan, S.; Lefevre, C.; Ersen, O.; Pichon, B.; Mertz, D.; et al. Unravelling the Thermal Decomposition Parameters for The Synthesis of Anisotropic Iron Oxide Nanoparticles. Nanomaterials 2018, 8, 881. [Google Scholar] [CrossRef] [PubMed]
  25. Takami, S.; Sato, T.; Mousavand, T.; Ohara, S.; Umetsu, M.; Adschiri, T. Hydrothermal synthesis of surface-modified iron oxide nanoparticles. Mater. Lett. 2007, 61, 4769–4772. [Google Scholar] [CrossRef]
  26. Xu, C.; Teja, A.S. Continuous hydrothermal synthesis of iron oxide and PVA-protected iron oxide nanoparticles. J. Supercrit. Fluids 2008, 44, 85–91. [Google Scholar] [CrossRef]
  27. Gurushankar, K.; Chinnaiah, K.; Kannan, K.; Gohulkumar, M.; Periyasamy, P. Synthesis and Characterization of FeO Nanoparticles by Hydrothermal Method. Rasayan J. Chem. 2021, 14, 1985–1989. [Google Scholar] [CrossRef]
  28. Hassanjani-Roshan, A.; Vaezi, M.R.; Shokuhfar, A.; Rajabali, Z. Synthesis of iron oxide nanoparticles via sonochemical method and their characterization. Particuology 2011, 9, 95–99. [Google Scholar] [CrossRef]
  29. Marchegiani, G.; Imperatori, P.; Mari, A.; Pilloni, L.; Chiolerio, A.; Allia, P.; Tiberto, P.; Suber, L. Sonochemical synthesis of versatile hydrophilic magnetite nanoparticles. Ultrason. Sonochem. 2012, 19, 877–882. [Google Scholar] [CrossRef] [PubMed]
  30. Jing, Z.; Wu, S. Synthesis, characterization and magnetic properties of γ-Fe2O3 nanoparticles via a non-aqueous medium. J. Solid State Chem. 2004, 177, 1213–1218. [Google Scholar] [CrossRef]
  31. Takai, Z.; Mustafa, M.K.; Asman, S.; Takai, Z.I.; Mustafa, M.K.; Sekak, K.A. Preparation and Characterization of Magnetite (Fe3O4) Nanoparticles by Sol-Gel Method. Int. J. Nanoelectron. Mater. 2019, 12, 37–46. [Google Scholar]
  32. Paulson, E.; Jothibas, M. Significance of thermal interfacing in hematite (α-Fe2O3) nanoparticles synthesized by sol-gel method and its characteristics properties. Surf. Interfaces 2021, 26, 101432. [Google Scholar] [CrossRef]
  33. Dave, S.R.; Gao, X. Monodisperse magnetic nanoparticles for biodetection, imaging, and drug delivery: A versatile and evolving technology. WIREs Nanomed. Nanobiotechnol. 2009, 1, 583–609. [Google Scholar] [CrossRef] [PubMed]
  34. Keenan, C.R.; Sedlak, D.L. Factors Affecting the Yield of Oxidants from the Reaction of Nanoparticulate Zero-Valent Iron and Oxygen. Environ. Sci. Technol. 2008, 42, 1262–1267. [Google Scholar] [CrossRef] [PubMed]
  35. Farrell, D.; Majetich, S.A.; Wilcoxon, J.P. Preparation and Characterization of Monodisperse Fe Nanoparticles. J. Phys. Chem. B 2003, 107, 11022–11030. [Google Scholar] [CrossRef]
  36. Vilardi, G.; Verdone, N.; Bubbico, R. Combined production of metallic-iron nanoparticles: Exergy and energy analysis of two alternative processes using Hydrazine and NaBH4 as reducing agents. J. Taiwan Inst. Chem. Eng. 2020, 118, 97–111. [Google Scholar] [CrossRef]
  37. Bharde, A.A.; Parikh, R.Y.; Baidakova, M.; Jouen, S.; Hannoyer, B.; Enoki, T.; Prasad, B.L.V.; Shouche, Y.S.; Ogale, S.; Sastry, M. Bacteria-Mediated Precursor-Dependent Biosynthesis of Superparamagnetic Iron Oxide and Iron Sulfide Nanoparticles. Langmuir 2008, 24, 5787–5794. [Google Scholar] [CrossRef]
  38. Bharde, A.; Wani, A.; Shouche, Y.; Joy, P.A.; Prasad, B.L.V.; Sastry, M. Bacterial Aerobic Synthesis of Nanocrystalline Magnetite. J. Am. Chem. Soc. 2005, 127, 9326–9327. [Google Scholar] [CrossRef]
  39. Moon, J.-W.; Rawn, C.J.; Rondinone, A.; Love, L.; Roh, Y.; Everett, S.; Lauf, R.J.; Phelps, T.J. Large-scale production of magnetic nanoparticles using bacterial fermentation. J. Ind. Microbiol. Biotechnol. 2010, 37, 1023–1031. [Google Scholar] [CrossRef]
  40. Sundaram, P.A.; Augustine, R.; Kannan, M. Extracellular biosynthesis of iron oxide nanoparticles by Bacillus subtilis strains isolated from rhizosphere soil. Biotechnol. Bioprocess Eng. 2012, 17, 835–840. [Google Scholar] [CrossRef]
  41. Elcey, C.D.; Kuruvilla, A.T.; Thomas, D. Synthesis of Magnetite Nanoparticles from Optimized Iron Reducing Bacteria Isolated from Iron Ore Mining Sites. Int. J. Curr. Microbiol. Appl. Sci. 2014, 3, 408–417. [Google Scholar]
  42. Kaul, R.K.; Kumar, P.; Burman, U.; Joshi, P.; Agrawal, A.; Raliya, R.; Tarafdar, J.C. Magnesium and iron nanoparticles production using microorganisms and various salts. Mater. Sci. 2012, 30, 254–258. [Google Scholar] [CrossRef]
  43. Mohamed, Y.M.; Azzam, A.M.; Amin, B.H.; Safwat, N.A. Mycosynthesis of iron nanoparticles by Alternaria alternata and its antibacterial activity. Afr. J. Biotechnol. 2015, 14, 1234–1241. [Google Scholar] [CrossRef] [Green Version]
  44. Mahanty, S.; Bakshi, M.; Ghosh, S.; Chatterjee, S.; Bhattacharyya, S.; Das, P.; Das, S.; Chaudhuri, P. Green Synthesis of Iron Oxide Nanoparticles Mediated by Filamentous Fungi Isolated from Sundarban Mangrove Ecosystem, India. Bionanoscience 2019, 9, 637–651. [Google Scholar] [CrossRef]
  45. Mahdavi, M.; Namvar, F.; Bin Ahmad, M.; Mohamad, R. Green Biosynthesis and Characterization of Magnetic Iron Oxide (Fe3O4) Nanoparticles Using Seaweed (Sargassum muticum) Aqueous Extract. Molecules 2013, 18, 5954–5964. [Google Scholar] [CrossRef] [PubMed]
  46. Subramaniyam, V.; Subashchandrabose, S.R.; Thavamani, P.; Megharaj, M.; Chen, Z.; Naidu, R. Chlorococcum sp. MM11—A novel phyco-nanofactory for the synthesis of iron nanoparticles. J. Appl. Phycol. 2015, 27, 1861–1869. [Google Scholar] [CrossRef]
  47. Machado, S.; Pinto, S.; Grosso, J.; Nouws, H.; Albergaria, J.; 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] [Green Version]
  48. Sathya, K.; Saravanathamizhan, R.; Baskar, G. Ultrasound assisted phytosynthesis of iron oxide nanoparticle. Ultrason. Sonochem. 2017, 39, 446–451. [Google Scholar] [CrossRef]
  49. 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]
  50. 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]
  51. Farshchi, H.K.; Azizi, M.; Jaafari, M.R.; Nemati, S.H.; Fotovat, A. Green synthesis of iron nanoparticles by Rosemary extract and cytotoxicity effect evaluation on cancer cell lines. Biocatal. Agric. Biotechnol. 2018, 16, 54–62. [Google Scholar] [CrossRef]
  52. 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 2018, 124, 363. [Google Scholar] [CrossRef]
  53. Noreen, S.; Ismail, S.; Ibrahim, S.M.; Kusuma, H.S.; Nazir, A.; Yaseen, M.; Khan, M.I.; Iqbal, M. ZnO, CuO and Fe2O3 green synthesis for the adsorptive removal of direct golden yellow dye adsorption: Kinetics, equilibrium and thermodynamics studies. Z. Für Phys. Chem. 2021, 235, 1055–1075. [Google Scholar] [CrossRef]
  54. 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. Z. Fur Phys. Chem. 2017, 7, 175–182. [Google Scholar] [CrossRef]
  55. Jegadeesan, G.B.; Srimathi, K.; Srinivas, N.S.S.; Manishkanna, S.; Vignesh, D. Green synthesis of iron oxide nanoparticles using Terminalia bellirica and Moringa oleifera fruit and leaf extracts: Antioxidant, antibacterial and thermoacoustic properties. Biocatal. Agric. Biotechnol. 2019, 21, 101354. [Google Scholar] [CrossRef]
  56. 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]
  57. Yusefi, M.; Shameli, K.; Yee, O.S.; Teow, S.-Y.; Hedayatnasab, Z.; Jahangirian, H.; Webster, T.J.; Kuča, K. Green Synthesis of Fe3O4 Nanoparticles Stabilized by a Garcinia mangostana Fruit Peel Extract for Hyperthermia and Anticancer Activities. Int. J. Nanomed. 2021, 16, 2515–2532. [Google Scholar] [CrossRef]
  58. 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 Biol. 2018, 180, 259–267. [Google Scholar] [CrossRef]
  59. Bibi, I.; Nazar, N.; Ata, S.; Sultan, M.; Ali, A.; Abbas, A.; Jilani, K.; Kamal, S.; Sarim, F.M.; Khan, M.I.; et al. Green synthesis of iron oxide nanoparticles using pomegranate seeds extract and photocatalytic activity evaluation for the degradation of textile dye. J. Mater. Res. Technol. 2019, 8, 6115–6124. [Google Scholar] [CrossRef]
  60. Ghafarzadegan, R.; Yaghoobi, M.; Momtaz, S.; Ashoory, N.; Yekta, M.G.; Hajiaghaee, R. Process optimization for green synthesis of iron nanoparticles by extract of fenugreek (Trigonella foenum-graecum L.) seeds. J. Med. Plants 2022, 21, 22–32. [Google Scholar] [CrossRef]
  61. Venkateswarlu, S.; Kumar, B.N.; Prasad, C.; Venkateswarlu, P.; Jyothi, N. Bio-inspired green synthesis of Fe3O4 spherical magnetic nanoparticles using Syzygium cumini seed extract. Phys. B Condens. Matter 2014, 449, 67–71. [Google Scholar] [CrossRef]
  62. Nnadozie, E.C.; Ajibade, P.A. Green synthesis and characterization of magnetite (Fe3O4) nanoparticles using Chromolaena odorata root extract for smart nanocomposite. Mater. Lett. 2020, 263, 127145. [Google Scholar] [CrossRef]
  63. Turakhia, B.; Chikkala, S.; Shah, S. Novelty of Bioengineered Iron Nanoparticles in Nanocoated Surgical Cotton: A Green Chemistry. Adv. Pharmacol. Sci. 2019, 2019, 9825969. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Njagi, E.C.; Huang, H.; Stafford, L.; Genuino, H.; Galindo, H.M.; Collins, J.B.; Hoag, G.E.; Suib, S.L. Biosynthesis of Iron and Silver Nanoparticles at Room Temperature Using Aqueous Sorghum Bran Extracts. Langmuir 2011, 27, 264–271. [Google Scholar] [CrossRef] [PubMed]
  65. Afsheen, S.; Tahir, M.B.; Iqbal, T.; Liaqat, A.; Abrar, M. Green synthesis and characterization of novel iron particles by using different extracts. J. Alloy. Compd. 2018, 732, 935–944. [Google Scholar] [CrossRef]
  66. Saif, S.; Tahir, A.; Chen, Y. Green Synthesis of Iron Nanoparticles and Their Environmental Applications and Implications. Nanomaterials 2016, 6, 209. [Google Scholar] [CrossRef]
  67. Bolade, O.P.; Williams, A.B.; Benson, N.U. Green synthesis of iron-based nanomaterials for environmental remediation: A review. Environ. Nanotechnol. Monit. Manag. 2020, 13, 100279. [Google Scholar] [CrossRef]
  68. Sharma, R.K.; Yadav, S.; Gupta, R.; Arora, G. Synthesis of Magnetic Nanoparticles Using Potato Extract for Dye Degradation: A Green Chemistry Experiment. J. Chem. Educ. 2019, 96, 3038–3044. [Google Scholar] [CrossRef]
  69. He, F.; Zhao, D. Preparation and Characterization of a New Class of Starch-Stabilized Bimetallic Nanoparticles for Degradation of Chlorinated Hydrocarbons in Water. Environ. Sci. Technol. 2005, 39, 3314–3320. [Google Scholar] [CrossRef]
  70. Shejawal, K.P.; Randive, D.S.; Bhinge, S.D.; Bhutkar, M.A.; Wadkar, G.H.; Jadhav, N.R. Green synthesis of silver and iron nanoparticles of isolated proanthocyanidin: Its characterization, antioxidant, antimicrobial, and cytotoxic activities against COLO320DM and HT29. J. Genet. Eng. Biotechnol. 2020, 18, 1–11. [Google Scholar] [CrossRef]
  71. Maleki, A.; Panahzadeh, M.; Eivazzadeh-Keihan, R. Agar: A natural and environmentally-friendly support composed of copper oxide nanoparticles for the green synthesis of 1,2,3–triazoles. Green Chem. Lett. Rev. 2019, 12, 395–406. [Google Scholar] [CrossRef]
  72. Sreeja, V.; Jayaprabha, K.N.; Joy, P.A. Water-dispersible ascorbic-acid-coated magnetite nanoparticles for contrast enhancement in MRI. Appl. Nanosci. 2015, 5, 435–441. [Google Scholar] [CrossRef] [Green Version]
  73. Răcuciu, M.; Barbu-Tudoran, L.; Oancea, S.; Drăghici, O.; Morosanu, C.; Grigoras, M.; Brînză, F.; Creangă, D.E. Aspartic Acid Stabilized Iron Oxide Nanoparticles for Biomedical Applications. Nanomaterials 2022, 12, 1151. [Google Scholar] [CrossRef] [PubMed]
  74. Cahyana, A.H.; Reza, A.I. Synthesis and characterization of Fe3O4 nanoparticles dispersed in paraffin as solvent. In AIP Conference Proceedings; AIP Publishing LLC: Melville, NY, USA, 2018; Volume 2049, p. 20010. [Google Scholar]
  75. Siskova, K.M.; Straska, J.; Krizek, M.; Tucek, J.; Machala, L.; Zboril, R. Formation of Zero-valent Iron Nanoparticles Mediated by Amino Acids. Procedia Environ. Sci. 2013, 18, 809–817. [Google Scholar] [CrossRef] [Green Version]
  76. Pereira, A.; Silva, N.; Trindade, T.; Pereira, S. A Single-Source Route for the Synthesis of Metal Oxide Nanoparticles Using Vegetable Oil Solvents. J. Nanosci. Nanotechnol. 2012, 12, 8963–8968. [Google Scholar] [CrossRef] [PubMed]
  77. Keshk, S.M.A.S.; El-Zahhar, A.A.; Abu Haija, M.; Bondock, S. Synthesis of a Magnetic Nanoparticles/Dialdehyde Starch-Based Composite Film for Food Packaging. Starch/Staerke 2019, 71, 1800035. [Google Scholar] [CrossRef]
  78. Jafarzadeh, S.; Salehabadi, A.; Nafchi, A.M.; Oladzadabbasabadi, N.; Jafari, S.M. Cheese packaging by edible coatings and biodegradable nanocomposites; improvement in shelf life, physicochemical and sensory properties. Trends Food Sci. Technol. 2021, 116, 218–231. [Google Scholar] [CrossRef]
  79. Mu, H.; Gao, H.; Chen, H.; Tao, F.; Fang, X.; Ge, L. A nanosised oxygen scavenger: Preparation and antioxidant application to roasted sunflower seeds and walnuts. Food Chem. 2013, 136, 245–250. [Google Scholar] [CrossRef]
  80. Foltynowicz, Z.; Bardenshtein, A.; Sängerlaub, S.; Antvorskov, H.; Kozak, W. Nanoscale, zero valent iron particles for application as oxygen scavenger in food packaging. Food Packag. Shelf Life 2017, 11, 74–83. [Google Scholar] [CrossRef]
  81. Vilela, C.; Kurek, M.; Hayouka, Z.; Röcker, B.; Yildirim, S.; Antunes, M.D.C.; Nilsen-Nygaard, J.; Pettersen, M.K.; Freire, C.S.R. A concise guide to active agents for active food packaging. Trends Food Sci. Technol. 2018, 80, 212–222. [Google Scholar] [CrossRef]
  82. Busolo, M.A.; Lagaron, J.M. Oxygen scavenging polyolefin nanocomposite films containing an iron modified kaolinite of interest in active food packaging applications. Innov. Food Sci. Emerg. Technol. 2012, 16, 211–217. [Google Scholar] [CrossRef]
  83. EFSA Panel on Food Contact Materials, Enzymes, Flavourings and Processing Aids (CEF). Scientific Opinion on the safety assessment of the active substances iron, iron oxides, sodium chloride and calcium hydroxide for use in food contact materials. EFSA J. 2013, 11, 3387. [Google Scholar] [CrossRef] [Green Version]
  84. Khalaj, M.-J.; Ahmadi, H.; Lesankhosh, R.; Khalaj, G. Study of physical and mechanical properties of polypropylene nanocomposites for food packaging application: Nano-clay modified with iron nanoparticles. Trends Food Sci. Technol. 2016, 51, 41–48. [Google Scholar] [CrossRef]
  85. Mary, T.R.N.; Jayavel, R. Fabrication of chitosan/Cashew Nut Shell Liquid/plant extracts-based bio-formulated nanosheets with embedded iron oxide nanoparticles as multi-functional barrier resist eco-packaging material. Appl. Nanosci. 2022, 12, 1719–1730. [Google Scholar] [CrossRef]
  86. Ligaj, M.; Tichoniuk, M.; Cierpiszewski, R.; Foltynowicz, Z. Efficiency of Novel Antimicrobial Coating Based on Iron Nanoparticles for Dairy Products’ Packaging. Coatings 2020, 10, 156. [Google Scholar] [CrossRef] [Green Version]
  87. Song, Z.; Niu, C.; Wu, H.; Wei, J.; Zhang, Y.; Yue, T. Transcriptomic Analysis of the Molecular Mechanisms Underlying the Antibacterial Activity of IONPs@pDA-Nisin Composites toward Alicyclobacillus acidoterrestris. ACS Appl. Mater. Interfaces 2019, 11, 21874–21886. [Google Scholar] [CrossRef]
  88. Zhu, N.; Ji, H.; Yu, P.; Niu, J.; Farooq, M.U.; Akram, M.W.; Udego, I.O.; Li, H.; Niu, X. Surface Modification of Magnetic Iron Oxide Nanoparticles. Nanomaterials 2018, 8, 810. [Google Scholar] [CrossRef] [Green Version]
  89. Kumar, S.; Ye, F.; Dobretsov, S.; Dutta, J. Chitosan Nanocomposite Coatings for Food, Paints, and Water Treatment Applications. Appl. Sci. 2019, 9, 2409. [Google Scholar] [CrossRef] [Green Version]
  90. Bahrami, A.; Delshadi, R.; Assadpour, E.; Jafari, S.M.; Williams, L. Antimicrobial-loaded nanocarriers for food packaging applications. Adv. Colloid Interface Sci. 2020, 278, 102140. [Google Scholar] [CrossRef]
  91. Singh, T.P.; Chatli, M.K.; Sahoo, J. Development of chitosan based edible films: Process optimization using response surface methodology. J. Food Sci. Technol. 2015, 52, 2530–2543. [Google Scholar] [CrossRef]
  92. Nehra, P.; Chauhan, R.; Garg, N.; Verma, K. Antibacterial and antifungal activity of chitosan coated iron oxide nanoparticles. Br. J. Biomed. Sci. 2018, 75, 13–18. [Google Scholar] [CrossRef]
  93. Shrifian-Esfahni, A.; Salehi, M.T.; Nasr-Esfahni, M.; Ekramian, E. Chitosan-Modified Superparamgnetic Iron Oxide Nanoparticles: Design, Fabrication, Characterization and Antibacterial Activity. Chemik 2015, 69, 19–32. [Google Scholar]
  94. Niemirowicz, K.; Markiewicz, K.; Wilczewska, A.; Car, H. Magnetic nanoparticles as new diagnostic tools in medicine. Adv. Med. Sci. 2012, 57, 196–207. [Google Scholar] [CrossRef] [PubMed]
  95. Lattuada, M.; Hatton, T.A. Functionalization of Monodisperse Magnetic Nanoparticles. Langmuir 2007, 23, 2158–2168. [Google Scholar] [CrossRef] [PubMed]
  96. Bashir, N.; Sood, M.; Bandral, J.D. Enzyme immobilization and its applications in food processing: A review. Int. J. Chem. Stud. 2020, 8, 254–261. [Google Scholar] [CrossRef]
  97. Díaz-Hernández, A.; Gracida, J.; García-Almendárez, B.E.; Regalado, C.; Núñez, R.; Amaro-Reyes, A. Characterization of Magnetic Nanoparticles Coated with Chitosan: A Potential Approach for Enzyme Immobilization. J. Nanomater. 2018, 2018, 9468574. [Google Scholar] [CrossRef]
  98. Xu, J.; Sun, J.; Wang, Y.; Sheng, J.; Wang, F.; Sun, M. Application of Iron Magnetic Nanoparticles in Protein Immobilization. Molecules 2014, 19, 11465–11486. [Google Scholar] [CrossRef]
  99. Wu, J.; Wang, X.; Wang, Q.; Lou, Z.; Li, S.; Zhu, Y.; Qin, L.; Wei, H. Nanomaterials with enzyme-like characteristics (nanozymes): Next-generation artificial enzymes (II). Chem. Soc. Rev. 2019, 48, 1004–1076. [Google Scholar] [CrossRef]
  100. Gao, L.; Fan, K.; Yan, X. Iron Oxide Nanozyme: A Multifunctional Enzyme Mimetic for Biomedical Applications. Theranostics 2017, 7, 3207–3227. [Google Scholar] [CrossRef]
  101. Qin, T.; Ma, R.; Yin, Y.; Miao, X.; Chen, S.; Fan, K.; Xi, J.; Liu, Q.; Gu, Y.; Yin, Y.; et al. Catalytic inactivation of influenza virus by iron oxide nanozyme. Theranostics 2019, 9, 6920–6935. [Google Scholar] [CrossRef]
  102. You, S.-M.; Park, J.-S.; Luo, K.; Jeong, K.-B.; Adra, H.J.; Kim, Y.-R. Modulation of the peroxidase-like activity of iron oxide nanoparticles by surface functionalization with polysaccharides and its application for the detection of glutathione. Carbohydr. Polym. 2021, 267, 118164. [Google Scholar] [CrossRef]
  103. Hernández-Hernández, A.A.; Álvarez-Romero, G.A.; Contreras-López, E.; Aguilar-Arteaga, K.; Castañeda-Ovando, A. Food Analysis by Microextraction Methods Based on the Use of Magnetic Nanoparticles as Supports: Recent Advances. Food Anal. Methods 2017, 10, 2974–2993. [Google Scholar] [CrossRef]
  104. Speroni, F.; Elviri, L.; Careri, M.; Mangia, A. Magnetic Particles Functionalized with PAMAM-Dendrimers and Antibodies: A New System for an ELISA Method Able to Detect Ara H3/4 Peanut Allergen in Foods. In Proceedings of the Analytical and Bioanalytical Chemistry; Springer: Berlin/Heidelberg, Germany, 2010; Volume 397, pp. 3035–3042. [Google Scholar]
  105. Cao, M.; Li, Z.; Wang, J.; Ge, W.; Yue, T.; Li, R.; Colvin, V.L.; Yu, W.W. Food related applications of magnetic iron oxide nanoparticles: Enzyme immobilization, protein purification, and food analysis. Trends Food Sci. Technol. 2012, 27, 47–56. [Google Scholar] [CrossRef]
  106. Limchoowong, N.; Sricharoen, P.; Areerob, Y.; Nuengmatcha, P.; Sripakdee, T.; Techawongstien, S.; Chanthai, S. Preconcentration and trace determination of copper (II) in Thai food recipes using Fe3O4 @Chi–GQDs nanocomposites as a new magnetic adsorbent. Food Chem. 2017, 230, 388–397. [Google Scholar] [CrossRef] [PubMed]
  107. Zhao, Y.; Bi, C.; He, X.; Chen, L.; Zhang, Y. Preparation of molecularly imprinted polymers based on magnetic carbon nanotubes for determination of sulfamethoxazole in food samples. RSC Adv. 2015, 5, 70309–70318. [Google Scholar] [CrossRef]
  108. Abolhasani, J.; Khanmiri, R.H.; Babazadeh, M.; Ghorbani-Kalhor, E.; Edjlali, L.; Hassanpour, A. Determination of Hg(II) ions in sea food samples after extraction and preconcentration by novel Fe3O4@SiO2@polythiophene magnetic nanocomposite. Environ. Monit. Assess. 2015, 187, 1–11. [Google Scholar] [CrossRef] [PubMed]
  109. Bagheri, H.; Asgharinezhad, A.A.; Ebrahimzadeh, H. Determination of Trace Amounts of Cd(II), Cu(II), and Ni(II) in Food Samples Using a Novel Functionalized Magnetic Nanosorbent. Food Anal. Methods 2016, 9, 876–888. [Google Scholar] [CrossRef]
  110. He, D.; Zhang, X.; Gao, B.; Wang, L.; Zhao, Q.; Chen, H.; Wang, H.; Zhao, C. Preparation of magnetic molecularly imprinted polymer for the extraction of melamine from milk followed by liquid chromatography-tandem mass spectrometry. Food Control. 2014, 36, 36–41. [Google Scholar] [CrossRef]
  111. Yu, X.; Ang, H.C.; Yang, H.; Zheng, C.; Zhang, Y. Low temperature cleanup combined with magnetic nanoparticle extraction to determine pyrethroids residue in vegetables oils. Food Control. 2017, 74, 112–120. [Google Scholar] [CrossRef]
  112. Pirouz, M.J.; Beyki, M.H.; Shemirani, F. Anhydride functionalised calcium ferrite nanoparticles: A new selective magnetic material for enrichment of lead ions from water and food samples. Food Chem. 2015, 170, 131–137. [Google Scholar] [CrossRef]
  113. Chen, H.; Deng, X.; Ding, G.; Qiao, Y. The synthesis, adsorption mechanism and application of polyethyleneimine functionalized magnetic nanoparticles for the analysis of synthetic colorants in candies and beverages. Food Chem. 2019, 293, 340–347. [Google Scholar] [CrossRef]
  114. Yu, C.; Guo, J.; Gu, H. Electrocatalytical Oxidation of Nitrite and Its Determination Based on Au@Fe3O4 Nanoparticles. Electroanalysis 2010, 22, 1005–1011. [Google Scholar] [CrossRef]
  115. Wen, C.-Y.; Hu, J.; Zhang, Z.-L.; Tian, Z.-Q.; Ou, G.-P.; Liao, Y.-L.; Li, Y.; Xie, M.; Sun, Z.-Y.; Pang, D.-W. One-Step Sensitive Detection of Salmonella typhimurium by Coupling Magnetic Capture and Fluorescence Identification with Functional Nanospheres. Anal. Chem. 2013, 85, 1223–1230. [Google Scholar] [CrossRef] [PubMed]
  116. Yang, H.; Qu, L.; Wimbrow, A.N.; Jiang, X.; Sun, Y. Rapid detection of Listeria monocytogenes by nanoparticle-based immunomagnetic separation and real-time PCR. Int. J. Food Microbiol. 2007, 118, 132–138. [Google Scholar] [CrossRef] [PubMed]
  117. Yang, K.; Jenkins, D.M.; Su, W.W. Rapid concentration of bacteria using submicron magnetic anion exchangers for improving PCR-based multiplex pathogen detection. J. Microbiol. Methods 2011, 86, 69–77. [Google Scholar] [CrossRef]
  118. Chen, X.; Wu, X.; Gan, M.; Xu, F.; He, L.; Yang, D.; Xu, H.; Shah, N.P.; Wei, H. Rapid detection of Staphylococcus aureus in dairy and meat foods by combination of capture with silica-coated magnetic nanoparticles and thermophilic helicase-dependent isothermal amplification. J. Dairy Sci. 2015, 98, 1563–1570. [Google Scholar] [CrossRef] [Green Version]
  119. Gravel, A.; Doyen, A. The use of edible insect proteins in food: Challenges and issues related to their functional properties. Innov. Food Sci. Emerg. Technol. 2020, 59, 102272. [Google Scholar] [CrossRef]
  120. Khan, S.; Roser, D.; Davies, C.; Peters, G.; Stuetz, R.; Tucker, R.; Ashbolt, N. Chemical contaminants in feedlot wastes: Concentrations, effects and attenuation. Environ. Int. 2008, 34, 839–859. [Google Scholar] [CrossRef] [PubMed]
  121. Gao, R.; Cui, X.; Hao, Y.; Zhang, L.; Liu, D.; Tang, Y. A highly-efficient imprinted magnetic nanoparticle for selective separation and detection of 17β-estradiol in milk. Food Chem. 2016, 194, 1040–1047. [Google Scholar] [CrossRef]
  122. Luo, Y.-B.; Yu, Q.-W.; Yuan, B.-F.; Feng, Y.-Q. Fast microextraction of phthalate acid esters from beverage, environmental water and perfume samples by magnetic multi-walled carbon nanotubes. Talanta 2012, 90, 123–131. [Google Scholar] [CrossRef] [PubMed]
  123. Kumari, A.; Chauhan, A.K. Iron nanoparticles as a promising compound for food fortification in iron deficiency anemia: A review. J. Food Sci. Technol. 2021, 59, 3319–3335. [Google Scholar] [CrossRef]
  124. Von Moos, L.M.; Schneider, M.; Hilty, F.M.; Hilbe, M.; Arnold, M.; Ziegler, N.; Mato, D.S.; Winkler, H.; Tarik, M.; Ludwig, C.; et al. Iron phosphate nanoparticles for food fortification: Biological effects in rats and human cell lines. Nanotoxicology 2017, 11, 496–506. [Google Scholar] [CrossRef]
  125. Arami, H.; Khandhar, A.; Liggitt, D.; Krishnan, K.M. In vivo delivery, pharmacokinetics, biodistribution and toxicity of iron oxide nanoparticles. Chem. Soc. Rev. 2015, 44, 8576–8607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Perfecto, A.; Elgy, C.; Valsami-Jones, E.; Sharp, P.; Hilty, F.; Fairweather-Tait, S. Mechanisms of Iron Uptake from Ferric Phosphate Nanoparticles in Human Intestinal Caco-2 Cells. Nutrients 2017, 9, 359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Hilty, F.M.; Knijnenburg, J.T.; Teleki, A.; Krumeich, F.; Hurrell, R.F.; Pratsinis, S.E.; Zimmermann, M.B. Incorporation of Mg and Ca into Nanostructured Fe2O3 Improves Fe Solubility in Dilute Acid and Sensory Characteristics in Foods. J. Food Sci. 2011, 76, N2–N10. [Google Scholar] [CrossRef] [PubMed]
  128. Mufti, N.; Atma, T.; Fuad, A.; Sutadji, E. Synthesis and Characterization of Black, Red and Yellow Nanoparticles Pigments from the Iron Sand. In AIP Conference Proceedings; AIP Publishing LLC: Melville, NY, USA, 2014; Volume 1617, pp. 165–169. [Google Scholar]
  129. Voss, L.; Hsiao, I.-L.; Ebisch, M.; Vidmar, J.; Dreiack, N.; Böhmert, L.; Stock, V.; Braeuning, A.; Loeschner, K.; Laux, P.; et al. The presence of iron oxide nanoparticles in the food pigment E172. Food Chem. 2020, 327, 127000. [Google Scholar] [CrossRef] [PubMed]
  130. Alshannaq, A.; Yu, J.-H. Occurrence, Toxicity, and Analysis of Major Mycotoxins in Food. Int. J. Environ. Res. Public Health 2017, 14, 632. [Google Scholar] [CrossRef]
  131. Agriopoulou, S.; Stamatelopoulou, E.; Varzakas, T. Advances in Occurrence, Importance, and Mycotoxin Control Strategies: Prevention and Detoxification in Foods. Foods 2020, 9, 137. [Google Scholar] [CrossRef]
  132. Horky, P.; Skalickova, S.; Baholet, D.; Skladanka, J. Nanoparticles as a Solution for Eliminating the Risk of Mycotoxins. Nanomaterials 2018, 8, 727. [Google Scholar] [CrossRef] [Green Version]
  133. Castro, S.S.L.; de Oliveira, M.F.; Stradiotto, N.R. Study of the Electrochemical Behavior of Histamine Using a Nafion®-Copper (II) Hexacyanoferrate Film-Modified Electrode. Int. J. Electrochem. Sci. 2010, 5, 1447–1456. [Google Scholar]
  134. Adivi, F.G.; Hashemi, P. Removal of histamine from biological samples by functionalized Fe3O4@Agarose@Silica nanoparticles and its fast determination by ion mobility spectrometry. Colloids Surf. B Biointerfaces 2021, 203, 111717. [Google Scholar] [CrossRef]
  135. Ranmadugala, D.; Ebrahiminezhad, A.; Manley-Harris, M.; Ghasemi, Y.; Berenjian, A. The effect of iron oxide nanoparticles on Bacillus subtilis biofilm, growth and viability. Process. Biochem. 2017, 62, 231–240. [Google Scholar] [CrossRef]
  136. Kobayashi, A.; Horikawa, M.; Kirschvink, J.L.; Golash, H.N. Magnetic control of heterogeneous ice nucleation with nanophase magnetite: Biophysical and agricultural implications. Proc. Natl. Acad. Sci. USA 2018, 115, 5383–5388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Kang, T.; Hoptowit, R.; Jun, S. Effects of an oscillating magnetic field on ice nucleation in aqueous iron-oxide nanoparticle dispersions during supercooling and preservation of beef as a food application. J. Food Process. Eng. 2020, 43, e13525. [Google Scholar] [CrossRef]
  138. Liu, G.; Gao, J.; Ai, H.; Chen, X. Applications and Potential Toxicity of Magnetic Iron Oxide Nanoparticles. Small 2013, 9, 1533–1545. [Google Scholar] [CrossRef] [PubMed]
  139. Diao, M.; Yao, M. Use of zero-valent iron nanoparticles in inactivating microbes. Water Res. 2009, 43, 5243–5251. [Google Scholar] [CrossRef] [PubMed]
  140. Gabrielyan, L.; Hovhannisyan, A.; Gevorgyan, V.; Ananyan, M.; Trchounian, A. Antibacterial effects of iron oxide (Fe3O4) nanoparticles: Distinguishing concentration-dependent effects with different bacterial cells growth and membrane-associated mechanisms. Appl. Microbiol. Biotechnol. 2019, 103, 2773–2782. [Google Scholar] [CrossRef] [PubMed]
  141. Murugan, K.; Wei, J.; Alsalhi, M.S.; Nicoletti, M.; Paulpandi, M.; Samidoss, C.M.; Dinesh, D.; Chandramohan, B.; Paneerselvam, C.; Subramaniam, J.; et al. Magnetic nanoparticles are highly toxic to chloroquine-resistant Plasmodium falciparum, dengue virus (DEN-2), and their mosquito vectors. Parasitol. Res. 2017, 116, 495–502. [Google Scholar] [CrossRef]
  142. Park, H.; Park, H.-J.; Kim, J.A.; Lee, S.H.; Kim, J.H.; Yoon, J.; Park, T.H. Inactivation of Pseudomonas aeruginosa PA01 biofilms by hyperthermia using superparamagnetic nanoparticles. J. Microbiol. Methods 2011, 84, 41–45. [Google Scholar] [CrossRef]
  143. Rodrigues, D.; Bañobre-López, M.; Espiña, B.; Rivas, J.; Azeredo, J. Effect of magnetic hyperthermia on the structure of biofilm and cellular viability of a food spoilage bacterium. Biofouling 2013, 29, 1225–1232. [Google Scholar] [CrossRef]
  144. Zhu, Q.; Li, N.; Wang, C.; Zhang, Q.; Sun, H. Effect of interactions between various humic acid fractions and iron nanoparticles on the toxicity to white rot fungus. Chemosphere 2020, 247, 125895. [Google Scholar] [CrossRef]
  145. Majeed, S.; Danish, M.; Ibrahim, M.N.M.; Sekeri, S.H.; Ansari, M.T.; Nanda, A.; Ahmad, G. Bacteria Mediated Synthesis of Iron Oxide Nanoparticles and Their Antibacterial, Antioxidant, Cytocompatibility Properties. J. Clust. Sci. 2021, 32, 1083–1094. [Google Scholar] [CrossRef]
  146. Parveen, S.; Wani, A.H.; Shah, M.A.; Devi, H.S.; Bhat, M.Y.; Koka, J.A. Preparation, characterization and antifungal activity of iron oxide nanoparticles. Microb. Pathog. 2018, 115, 287–292. [Google Scholar] [CrossRef]
  147. Vignesh, V.; Sathiyanarayanan, G.; Sathishkumar, G.; Parthiban, K.; Sathish-Kumar, K.; Thirumurugan, R. Formulation of iron oxide nanoparticles using exopolysaccharide: Evaluation of their antibacterial and anticancer activities. RSC Adv. 2015, 5, 27794–27804. [Google Scholar] [CrossRef]
  148. Aisida, S.O.; Madubuonu, N.; Alnasir, M.H.; Ahmad, I.; Botha, S.; Maaza, M.; Ezema, F.I. Biogenic synthesis of iron oxide nanorods using Moringa oleifera leaf extract for antibacterial applications. Appl. Nanosci. 2020, 10, 305–315. [Google Scholar] [CrossRef]
  149. Nadeem, M.; Khan, R.; Shah, N.; Bangash, I.R.; Abbasi, B.H.; Hano, C.; Liu, C.; Ullah, S.; Hashmi, S.S.; Nadhman, A.; et al. A Review of Microbial Mediated Iron Nanoparticles (IONPs) and Its Biomedical Applications. Nanomaterials 2022, 12, 130. [Google Scholar] [CrossRef] [PubMed]
  150. Gudkov, S.; Burmistrov, D.; Serov, D.; Rebezov, M.; Semenova, A.; Lisitsyn, A. Do Iron Oxide Nanoparticles Have Significant Antibacterial Properties? Antibiotics 2021, 10, 884. [Google Scholar] [CrossRef]
  151. Maduray, K.; Parboosing, R. Metal Nanoparticles: A Promising Treatment for Viral and Arboviral Infections. Biol. Trace Elem. Res. 2020, 199, 3159–3176. [Google Scholar] [CrossRef] [PubMed]
  152. Caldeirão, A.; Araujo, H.; Tomasella, C.; Sampaio, C.; Oliveira, M.D.S.; Ramage, G.; Pessan, J.; Monteiro, D. Effects of Antifungal Carriers Based on Chitosan-Coated Iron Oxide Nanoparticles on Microcosm Biofilms. Antibiotics 2021, 10, 588. [Google Scholar] [CrossRef] [PubMed]
  153. Fischer, H.C.; Chan, W.C. Nanotoxicity: The growing need for in vivo study. Curr. Opin. Biotechnol. 2007, 18, 565–571. [Google Scholar] [CrossRef]
  154. Kedziorek, D.A.; Muja, N.; Walczak, P.; Ruiz-Cabello, J.; Gilad, A.A.; Jie, C.C.; Bulte, J.W.M. Gene expression profiling reveals early cellular responses to intracellular magnetic labeling with superparamagnetic iron oxide nanoparticles. Magn. Reson. Med. 2010, 63, 1031–1043. [Google Scholar] [CrossRef] [Green Version]
  155. Monteiro-Riviere, N.; Inman, A.; Zhang, L. Limitations and relative utility of screening assays to assess engineered nanoparticle toxicity in a human cell line. Toxicol. Appl. Pharmacol. 2009, 234, 222–235. [Google Scholar] [CrossRef]
  156. Malhotra, N.; Lee, J.S.; Liman, R.A.D.; Ruallo, J.M.S.; Villaflore, O.B.; Ger, T.R.; Hsiao, C. Der Potential Toxicity of Iron Oxide Magnetic Nanoparticles: A Review. Molecules 2020, 25, 3159. [Google Scholar] [CrossRef] [PubMed]
  157. Singh, N.; Jenkins, G.J.; Asadi, R.; Doak, S.H. Potential toxicity of superparamagnetic iron oxide nanoparticles (SPION). Nano Rev. 2010, 1, 5358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Xie, W.; Guo, Z.; Gao, F.; Gao, Q.; Wang, D.; Liaw, B.-S.; Cai, Q.; Sun, X.; Wang, X.; Zhao, L. Shape-, size- and structure-controlled synthesis and biocompatibility of iron oxide nanoparticles for magnetic theranostics. Theranostics 2018, 8, 3284–3307. [Google Scholar] [CrossRef] [PubMed]
  159. Ankamwar, B.; Lai, T.C.; Huang, J.H.; Liu, R.S.; Hsiao, M.; Chen, C.H.; Hwu, Y.K. Biocompatibility of Fe3O4nanoparticles evaluated by in vitro cytotoxicity assays using normal, glia and breast cancer cells. Nanotechnology 2010, 21, 9. [Google Scholar] [CrossRef] [PubMed]
  160. Liu, Y.; Chen, Z.; Wang, J. Systematic evaluation of biocompatibility of magnetic Fe3O4 nanoparticles with six different mammalian cell lines. J. Nanoparticle Res. 2011, 13, 199–212. [Google Scholar] [CrossRef]
  161. Zhou, X.; Shi, Y.; Ren, L.; Bao, S.; Han, Y.; Wu, S.; Zhang, H.; Zhong, L.; Zhang, Q. Controllable synthesis, magnetic and biocompatible properties of Fe3O4 and α-Fe2O3 nanocrystals. J. Solid State Chem. 2012, 196, 138–144. [Google Scholar] [CrossRef]
  162. Chu, M.; Shao, Y.; Peng, J.; Dai, X.; Li, H.; Wu, Q.; Shi, D. Near-infrared laser light mediated cancer therapy by photothermal effect of Fe3O4 magnetic nanoparticles. Biomaterials 2013, 34, 4078–4088. [Google Scholar] [CrossRef]
  163. Fan, H.-M.; Olivo, M.; Shuter, B.; Yi, J.-B.; Bhuvaneswari, R.; Tan, H.-R.; Xing, G.-C.; Ng, C.-T.; Liu, L.; Lucky, S.S.; et al. Quantum Dot Capped Magnetite Nanorings as High Performance Nanoprobe for Multiphoton Fluorescence and Magnetic Resonance Imaging. J. Am. Chem. Soc. 2010, 132, 14803–14811. [Google Scholar] [CrossRef]
  164. Weissleder, R.; Nahrendorf, M.; Pittet, M.J. Imaging macrophages with nanoparticles. Nat. Mater. 2014, 13, 125–138. [Google Scholar] [CrossRef]
  165. Veiseh, O.; Gunn, J.W.; Zhang, M. Design and fabrication of magnetic nanoparticles for targeted drug delivery and imaging. Adv. Drug Deliv. Rev. 2010, 62, 284–304. [Google Scholar] [CrossRef] [Green Version]
  166. Wu, W.; He, Q.; Jiang, C. Magnetic Iron Oxide Nanoparticles: Synthesis and Surface Functionalization Strategies. Nanoscale Res. Lett. 2008, 3, 397–415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Harris, J.M.; Chess, R.B. Effect of pegylation on pharmaceuticals. Nat. Rev. Drug Discov. 2003, 2, 214–221. [Google Scholar] [CrossRef] [PubMed]
  168. Mulens-Arias, V.; Rojas, J.M.; Sanz-Ortega, L.; Portilla, Y.; Pérez-Yagüe, S.; Barber, D.F. Polyethylenimine-coated superparamagnetic iron oxide nanoparticles impair in vitro and in vivo angiogenesis. Nanomed. Nanotechnol. Biol. Med. 2019, 21, 102063. [Google Scholar] [CrossRef] [PubMed]
  169. Nie, L.; Chang, P.; Ji, C.; Zhang, F.; Zhou, Q.; Sun, M.; Sun, Y.; Politis, C.; Shavandi, A. Poly(acrylic acid) capped iron oxide nanoparticles via ligand exchange with antibacterial properties for biofilm applications. Colloids Surf. B Biointerfaces 2021, 197, 111385. [Google Scholar] [CrossRef]
  170. Okassa, L.N.; Marchais, H.; Douziech-Eyrolles, L.; Hervé, K.; Cohen-Jonathan, S.; Munnier, E.; Soucé, M.; Linassier, C.; Dubois, P.; Chourpa, I. Optimization of iron oxide nanoparticles encapsulation within poly(d,l-lactide-co-glycolide) sub-micron particles. Eur. J. Pharm. Biopharm. 2007, 67, 31–38. [Google Scholar] [CrossRef]
  171. Predescu, A.M.; Matei, E.; Berbecaru, A.C.; Pantilimon, C.; Drăgan, C.; Vidu, R.; Predescu, C.; Kuncser, V. Synthesis and characterization of dextran-coated iron oxide nanoparticles. R. Soc. Open Sci. 2018, 5, 171525. [Google Scholar] [CrossRef]
  172. Nadeem, M.; Ahmad, M.; Akhtar, M.S.; Shaari, A.; Riaz, S.; Naseem, S.; Masood, M.; Saeed, M.A. Magnetic Properties of Polyvinyl Alcohol and Doxorubicine Loaded Iron Oxide Nanoparticles for Anticancer Drug Delivery Applications. PLoS ONE 2016, 11, e0158084. [Google Scholar] [CrossRef] [Green Version]
  173. Han, L.; Zhou, X. Synthesis and characterization of liposomes nano-composite-particles with hydrophobic magnetite as a MRI probe. Appl. Surf. Sci. 2016, 376, 252–260. [Google Scholar] [CrossRef]
  174. Xiong, F.; Wang, H.; Feng, Y.; Li, Y.; Hua, X.; Pang, X.; Zhang, S.; Song, L.; Zhang, Y.; Gu, N. Cardioprotective activity of iron oxide nanoparticles. Sci. Rep. 2015, 5, srep08579. [Google Scholar] [CrossRef] [Green Version]
  175. Zhang, Y.; Wang, Z.; Li, X.; Wang, L.; Yin, M.; Wang, L.; Chen, N.; Fan, C.; Song, H. Dietary Iron Oxide Nanoparticles Delay Aging and Ameliorate Neurodegeneration in Drosophila. Adv. Mater. 2016, 28, 1387–1393. [Google Scholar] [CrossRef]
  176. Feitosa, K.A.; Correia, R.D.O.; Fattori, A.C.M.; Albuquerque, Y.R.; Brassolatti, P.; Luna, G.F.; Rodolpho, J.M.D.A.; Nogueira, C.T.; Bernardi, J.C.; Speglich, C.; et al. Toxicological effects of the mixed iron oxide nanoparticle (Fe3O4 NP) on murine fibroblasts LA-9. J. Toxicol. Environ. Health Part A 2022, 85, 649–670. [Google Scholar] [CrossRef] [PubMed]
  177. Wang, X.; Tu, Q.; Zhao, B.; An, Y.; Wang, J.-C.; Liu, W.; Yuan, M.-S.; Ahmed, S.M.; Xu, J.; Liu, R.; et al. Effects of poly(l-lysine)-modified Fe3O4 nanoparticles on endogenous reactive oxygen species in cancer stem cells. Biomaterials 2013, 34, 1155–1169. [Google Scholar] [CrossRef] [PubMed]
  178. Bellova, A.; Bystrenova, E.; Koneracka, M.; Kopcansky, P.; Valle, F.; Tomasovicova, N.; Timko, M.; Bagelova, J.; Biscarini, F.; Gazova, Z. Effect of Fe3O4magnetic nanoparticles on lysozyme amyloid aggregation. Nanotechnology 2010, 21, 065103. [Google Scholar] [CrossRef] [PubMed]
  179. Karlsson, H.L.; Cronholm, P.; Gustafsson, J.; Möller, L. Copper Oxide Nanoparticles Are Highly Toxic: A Comparison between Metal Oxide Nanoparticles and Carbon Nanotubes. Chem. Res. Toxicol. 2008, 21, 1726–1732. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of a protein purification system based on magnetic nanoparticles.
Figure 1. Schematic illustration of a protein purification system based on magnetic nanoparticles.
Materials 16 00780 g001
Figure 2. Criteria for evaluation of toxicity of IONPs.
Figure 2. Criteria for evaluation of toxicity of IONPs.
Materials 16 00780 g002
Table 1. Some typical synthesis methods for iron nanoparticles.
Table 1. Some typical synthesis methods for iron nanoparticles.
MethodTypeShapeSizeReferences
Chemical reduction methodα-Fespherical21 nm[9]
Fe0spherical10–80 nm[10]
Co-precipitationFe3O4 and FeOdifferent8.5 nm[11]
Fe3O4-20–22 nm[12]
Fe3O4spherical15 nm[13]
γ-Fe2O3spherical11.40 and 7.3657 nm[14]
γ-Fe2O3sphericalfrom 5 to 8 nm[15]
Fe3O4-13.9 nm[16]
FeOHpolygonal disc6.60 nm[17]
Microemulsion synthesisFe3O4spherical~12 nm[18]
γ-Fe2O3spherical>10 nm[19]
Fe3O4-8–16 nm[20]
Thermal decompositionFe and Fe3O4spherical5.41 and 7.38 nm[21]
Superparamagnetic iron oxide 2–30 nm[22]
Fe3O4 11 nm[23]
Fe1−xO@Fe3−xO4cubes17.5 nm length 6.3, nm thickness[24]
Hydrothermal synthesisα-Fe2O3 or Fe3O4cubic and spherical25 or 14 nm, respectively[25]
α-Fe2O3spherical15.6 nm[26]
FeOrhombohedral22, 14, 8 nm[27]
Sonochemical processα-Fe2O3rhombohedralFrom ~12 nm to ~19 nm[28]
Fe3O4-20 nm[29]
Non-aqueous methodγ-Fe2O3acicular shapedmajor axis: 17 nm; minor axis: 1.7 nm[30]
Sol–gel methodFe3O4spherical79.04 nm[31]
α-Fe2O3spherical and spheroid-shaped12.7 nm[32]
Table 2. Examples of microorganisms used to synthesize iron nanoparticles.
Table 2. Examples of microorganisms used to synthesize iron nanoparticles.
MicroorganismSpeciesNPsPrecursorReferences
BacteriaActinobacter sp.γ-Fe2O3 maghemite
Fe3O4 magnetite
aqueous potassium ferricyanide/ferrocyanide
K3[Fe(CN)6]/
K4[Fe(CN)6]
[38]
Actinobacter sp.maghemite (γ-Fe2O3) and greigite (Fe3S4)aqueous solution ferrous salts[37]
Thermoanaerobacter sp.Fe3O4 magnetiteFeOOH[39]
Bacillus subtilisFe3O4Fe2O3[40]
Thiobacillus thioparusFe3O4 magnetiteFeSO4[41]
Alcaligenes faecalisFe or their oxidesFe2O3FeSO4[42]
FungiPochonia chlamydosporiumAspergillus fumigatus
Aspergillus wentii
Curvularia lunata Chaetmium globosum
Fe or their oxidesFe2O3
FeSO4
[42]
Alternaria alternataγ-Fe2O3 and α-Fe2O3FeCl3[43]
Trichoderma asperellum
Phialemoniopsis ocularis
Fusarium incarnatum
Fe oxidesFeCl3
FeCl2
[44]
AlgaeSargassum muticumFe3O4ferric chloride solution[45]
Chlorococcum sp.Feiron chloride[46]
Table 3. Examples of plants used to synthesize iron nanoparticles.
Table 3. Examples of plants used to synthesize iron nanoparticles.
Part of PlantSpeciesPrecursorReferences
LeavesQuercus spp.iron(III) chloride solution[47]
Coriandrum sativumFerric Chloride[48]
Eucalyptus spp.FeSO4 and Ni(NO3)2[49]
Moringa oleiferaIron nitrate(III) (Fe(NO3)3·9H2O)[50]
Rosmarinus officinalisFeSO4[51]
Daphne mezereumFeCl3·6H2O[52]
Cassia fistulaFe2O3[53]
FruitsPiper spp.FeCl2 and K2PdCl4[54]
Terminalia bellirica
Moringa oleifera
ferric chloride heptahydrate (FeCl2·7H2O) and potassium ferricyanide (K3Fe(CN)6)[55]
PeelArtocarpus heterophyllusFeCl2[56]
Garcinia mangostanairon(II) chloride tetrahydrate (FeCl2·4H2O ≥ 99%) and iron(III) chloride hexahydrate (FeCl3·6H2O, 97%)[57]
FlowerHibiscus sabdariffaFeCl3[58]
SeedsPunica granatumiron chloride[59]
Trigonella foenum-graecumFeCl3·6H2O[60]
Syzygium cuminiFeCl3·6H2O[61]
RootsChromolaena odoratasolution of Fe(II) and Fe(III)[62]
Zingiber officinaleferric chloride FeCl3[63]
BransSorghum moenchFeCl3[64]
BudsSyzygium aromaticumiron chloride tetra-hydrate[65]
Table 4. Examples of synthesis of iron nanoparticles using green reagents.
Table 4. Examples of synthesis of iron nanoparticles using green reagents.
Type of ReagentMaterialPrecursorNanoparticlesReferences
BiopolymerStarchFeSO4·7H2OFe3O4[68]
StarchFeCl3Fe-Pd nanoparticles[69]
Sodium alginateFeCl3Fe3O4[70]
AgarFeCl2·4H2O
FeCl3·6H2O
Fe3O4[71]
PolyphenolProanthocyanidinferric chloride solutionPACFeNPs[70]
AcidAscorbic acidFeCl2·4H2O and 2 mM of FeCl3·6H2OSPION[72]
Aspartic acidferrous chloride tetra-hydrate (FeCl2·H2O), ferric chloride hexa-hydrate
(FeCl3·6H2O)
A-IONPs[73]
Palmitic
acid
rustFe3O4[74]
L-glutamic acid
L-glutamine
L-arginine
L-cysteine
FeSO4·7H2OnZVI[75]
OilSunflower oilFeCl3·6H2OnZVI[76]
Table 5. Examples of the use of IONPs in food analysis.
Table 5. Examples of the use of IONPs in food analysis.
Food TypeAnalyzed Substance/OrganismMaterial for Functionalized IONPsReferences
Thai foodCu(II)chitosan–graphene quantum dots[106]
milk and honeysulfamethoxazolemolecularly imprinted polymers[107]
fish, shrimp, canned tunaHg(II)SiO2@polythiophene[108]
cantaloupe, apple, nectarineCd(II), Cu(II), Ni(II)2-aminobenzothiazole[109]
milkmelamineSiO2@MIPs[110]
vegetable oilpesticidepolystyrene[111]
canned tuna fish, canned tomato paste, parsley, milkPb(II)3-aminopropyl-trie-
thoxysilane@ phthalic anhydride
[112]
candies and beveragessynthetic colorantspolyethyleneimine[113]
sausagenitrateAu@l-cysteine[114]
milkSalmonella typhimuriumCdSe/ZnS QDs with
NH 2−PEG−CM(MW 3400) as spacers to be conjugated withantibodies
[115]
milkListeria monocytogenescarboxyl with rabbit anti-Listeria monocytogenes[116]
Table 6. The ways in which IONPs interact with microorganisms.
Table 6. The ways in which IONPs interact with microorganisms.
IONPsMicroorganismSpeciesEffectReferences
Fe3O4/SiO2
vancomycin
bacteriaBacillus cereus
Staphylococcus aureus Shigella boydii
Escherichia coli
inhibition of bacterial growth[18]
Fe0bacteriaEscherichia coliinactivation[10]
Bacillus subtilis
Escherichia coli
Staphylococcus epidermidis
3% provided full inhibition of microbial growth[86]
Bacillus subtilis
var. niger
Pseudomonas fluorescens
inactivation[138]
fungusAspergillus versicolorinactivation
Geotrichum candidum
Rhodotorula rubra
3% provided full inhibition of microbial growth[86]
S-Fe0bacteriaPseudomonas spp. HLS-6
Escherichia coli DH5α
oxidative stress in the bacteria, destroys the cell structure and damages the intracellular DNA[139]
IONzymyvirusesInfluenza Ainduces peroxidation of membrane lipids in synthesized liposomes, inactivates viruses[101]
Fe3O4bacteriaEscherichia coli BW 25113
Enterococcus hirae ATCC 9790
change in membrane permeability, bactericidal effect on E. coli[140]
protozoanPlasmodium falciparuminhibited replication by inhibiting E protein expression[141]
virusserotype DEN-2
Superparamagnetic iron oxidebacteriaPseudomonas aeruginosa PA01disintegration of the bacterial cell membrane by hyperthermia[142]
Pseudomonas fluorescens[143]
γ-Fe2O3fungusPhanerochaete chrysosporiumconcentration and exposure time affected malondialdehyde content, reactive oxygen species production and lactate dehydrogenase (LDH) activity[144]
γ-Fe2O3 and α-
Fe2O3
bacteriaEscherichia coli
Pseudomonas aeruginosa
Bacillus subtilis
Staphylococcus
aureus
inhibition of bacterial growth[43]
IONPsbacteriaStaphylococcus aureus
Staphylococcus epidermidis
Escherichia coli
Salmonella typhi
Vibrio cholera
[145]
fungusTrichothecium roseum, Cladosporium herbarum, Penicillium chrysogenum
Alternaria alternate
Aspergillus niger
inhibitory effect on growth[146]
Aeromonas hydrophila (ATCC 49140)
Aeromonas hydrophila (MTCC 1739)
Aeromonas sobria (MTCC 3613)
Aeromonas hydrophila
[147]
FeOHbacteriaStaphylococcus aureusexcellent antimicrobial activity[17]
FeO-NPs/FeO-NRsbacteriaStaphylococcus aureus
Escherichia coli
Pseudomonas aeruginosa, Shigella
Salmonella typhi
Pasteurella
at 10 μg mL−1, it showed a zone of growth inhibition[148]
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

Góral, D.; Marczuk, A.; Góral-Kowalczyk, M.; Koval, I.; Andrejko, D. Application of Iron Nanoparticle-Based Materials in the Food Industry. Materials 2023, 16, 780. https://0-doi-org.brum.beds.ac.uk/10.3390/ma16020780

AMA Style

Góral D, Marczuk A, Góral-Kowalczyk M, Koval I, Andrejko D. Application of Iron Nanoparticle-Based Materials in the Food Industry. Materials. 2023; 16(2):780. https://0-doi-org.brum.beds.ac.uk/10.3390/ma16020780

Chicago/Turabian Style

Góral, Dariusz, Andrzej Marczuk, Małgorzata Góral-Kowalczyk, Iryna Koval, and Dariusz Andrejko. 2023. "Application of Iron Nanoparticle-Based Materials in the Food Industry" Materials 16, no. 2: 780. https://0-doi-org.brum.beds.ac.uk/10.3390/ma16020780

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