Next Article in Journal
Influence of the Current Regime during Electrodeposition in a Cr(III)-Containing Fe-Cr-Ni Electrolyte on the Near-Surface pH, Alloy Composition, and Microcrack Behavior
Previous Article in Journal
Flow of Maxwell Fluid with Heat Transfer through Porous Medium with Thermophoresis Particle Deposition and Soret–Dufour Effects: Numerical Solution
Previous Article in Special Issue
SEM-EDX Analysis of Metal Particles Deposition from Surgical Burs after Implant Guided Surgery Procedures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nanoantioxidants: The Fourth Generation of Antioxidants—Recent Research Roadmap and Future Perspectives

by
Syed Tawab Shah
1,*,
Zaira Zaman Chowdhury
1,*,
Khanom Simarani
2,
Wan Jefrey Basirun
3,
Irfan Anjum Badruddin
4,
Mohamed Hussien
5,6,
Hussein Alrobei
7 and
Sarfaraz Kamangar
4
1
Nanotechnology & Catalysis Research Centre (NANOCAT), Institute of Postgraduate Studies Building, University of Malaya, Block A, Level 3, Kuala Lumpur 50603, Malaysia
2
Institute of Biological Sciences, Faculty of Science, University of Malaya, Kuala Lumpur 50603, Malaysia
3
Department of Chemistry, Faculty of Science, University of Malaya, Kuala Lumpur 50603, Malaysia
4
Mechanical Engineering Department, College of Engineering, King Khalid University, Abha 61413, Saudi Arabia
5
Department of Chemistry, Faculty of Science, King Khalid University, Abha 61413, Saudi Arabia
6
Agricultural Research Center, Central Agricultural Pesticide Laboratory, Pesticide Formulation Department, Dokki, Giza 12618, Egypt
7
Department of Mechanical Engineering, College of Engineering, Prince Sattam Bin Abdullaziz University, Al-Kharj 11942, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Coatings 2022, 12(10), 1568; https://doi.org/10.3390/coatings12101568
Submission received: 20 June 2022 / Revised: 8 September 2022 / Accepted: 20 September 2022 / Published: 17 October 2022
(This article belongs to the Collection Advanced Surface Coating of Nanoparticles)

Abstract

:
Antioxidants work by interacting with free radicals and converting them into harmless chemicals, interfering with the progression of potentially hazardous chain reactions. Antioxidants are useful in treating illnesses induced by free radicals because they help minimize oxidative stress. Antioxidants, whether natural or synthetic, have a limited effect on cellular health and function because of their low absorption, inability to traverse cellular membrane, and disintegration during delivery. The benefits of antioxidants, both natural and synthetic, are comparable. The use of antioxidants that are covalently attached to nanoparticles, or encased in particles with a hollow center, or feature the nanomaterial encapsulation of various origins has been employed to solve these challenges to provide improved stability, slow and slow sustained release, biocompatibility, and targeted administration. This review examines the importance of metal-based antioxidants and methods for enhancing antioxidant activities based on recent studies.

1. Introduction

Age-related disorders have emerged as a major public health concern worldwide in the twenty-first century. The social and economic burden of population aging is becoming increasingly substantial in the modern world. As people live longer, their health span does not necessarily rise at the same rate. Over the last few decades, the burden of age-related illnesses has grown progressively in most developed countries [1,2]. The process of aging is one that occurs throughout time and is a consequence of the degeneration of vital organ structures and functions. It increases the chance of developing a wide array of chronic diseases and contributes to a high death rate [3,4]. The free radical theory is unmatched among several ideas that explain and uncover the ageing process [5]. According to this view, reactive oxygen species (ROS) damage to cells and tissues is exacerbated with age because the body’s defense mechanisms repeatedly fail to heal it [6].
Degenerative aging is well understood to be a result of oxidative stress, which is a significant contributor to the overall process. When reactive oxygen species (ROS) are present, the pathophysiology of a number of different biological processes can be affected They have also been linked to heart disease, cancer, neurological illness, and pulmonary disease, as shown in Figure 1 [7]. ROS formation in cells has also been related to the progression of ageing; nevertheless, this cannot be considered the only factor responsible for the progression of aging. Furthermore, increasing levels of ROS are associated with mitochondrial dysfunction and cellular oxidative damage in age-related disorders [8,9].
Organic materials, such as biological components such as proteins and lipids, also include dietary supplements and cosmetics. An oxidative chain is established when atmospheric O2 reacts with alkyl oxidation products to form peroxyl (ROO) radicals. It is called autoxidation or peroxidation because hydroperoxides (alkyl hydroperoxides and H2O2) are the principal first-formed products of the autoxidation process [10]. Due to their unstable nature, hydroperoxides are capable of generating highly reactive hydroxyl (HO•) and alkoxyl (RO•) radicals, two types of radicals that can damage even generally stable molecules, such as DNA bases [11]. The production of reactive carbonyl species (such as 4-hydroxynonenal) due to alkyl hydro-peroxide cleavage exacerbates the oxidative damage [12]. When a cell’s ability to produce reactive oxygen species (ROS) exceeds its ability to mount an efficient antioxidant response, it experiences oxidative stress, which leads to cell death and mutation. Oxidative stress causes irreversible damage to proteins, lipids, and DNA, as well as mutation and cell death [13].
Research into bioactive antioxidants that lower oxidative stress and promote immunity is ongoing [14]. Researchers always search for new antioxidant molecules to promote healthy aging and prevent oxidative stress. This review covers the significance of metal-based antioxidants as well as strategies for boosting antioxidant activity based on current research.

2. Antioxidants

Antioxidants are substances or compounds that can prevent the oxidation of a valuable substrate even at low concentrations [10,15]. The discovery of antioxidants sparked a boom in the late 19th and early 20th centuries, thanks to their use in various industrial processes such as corrosion prevention, fuel polymerization, combustion engine fouling, and rubber vulcanization [16]. Antioxidants were formerly used only to protect unsaturated fats from oxidizing, resulting in a rancidity of fats [17]. Any compound’s antioxidant properties can be determined by measuring the rate at which fat consumes oxygen when stored oxygen-filled container within a sealed environment. Antioxidant vitamins A, C, and E were discovered and proved to be crucial in the biochemistry of living organisms, revolutionizing the field of nutrition [18,19]. The antioxidant is a stable molecule that donates an electron to harmful free radicals, neutralizing the resulting radical and reducing the amount of damage that it may induce. Generally speaking, these antioxidants either stop or delay cellular damage from occurring [20]. Antioxidants with a low molecular weight can swiftly and effectively interact with ROS, limiting damage to essential molecules from the chain reaction. Our bodies naturally produce antioxidant molecules such as glutathione, uric acid, and ubiquinol during regular metabolic activities [21]. Ascorbic acid (vitamin C), β-carotene, and α-tocopherol (vitamin E) are micronutrients and enzymes found in human bodies that can combat free radicals [22]. Because the body cannot produce these molecules, they must be taken from the food we consume. Endogenous antioxidants are those produced within, while exogenous antioxidants come from outside the body (found in foods such as glutathione and uric acid) [23]. Antioxidants found in nature may help protect against the damage caused by oxidative stress. Numerous fruits, vegetables, and byproducts of fruits contain antioxidants that have been shown to lower the amount of harmful free radicals in the body. These antioxidants include polyphenols (3,6-dihydroxyflavone), carotenoids (lutein), vitamins (ascorbic acid), and metabolic sensitizers (selenium methyl selenocysteine) [24,25,26,27]. Bioactive substances from various sources, such as antioxidants obtained from natural resources (such as Rosa rugosa dried flower extracts), have been found to exhibit scavenging and electrophile removal abilities and the ability to prevent the formation of reactive oxygen species [28,29]. On the other hand, natural antioxidants prone towards deterioration and have low bioavailability, poor absorption, and deterioration upon delivery are the main reasons [30]. Antioxidants include synthetic compounds such as butylated hydroxytoluene (BHT) and butylated hydroxyanisole (BHA), BHT analogues, GA esters, and other derivatives [24,31,32]. Nanotechnology has opened up new possibilities in a variety of fields. Nanoparticles are ideal for medication and drug delivery because of their small size and versatility. As a result of the integration of nanotechnology with material science and engineering, significant progress has been made in the reduction in free radical production in a variety of fields [31,32]. Antioxidant nanoparticles or nano-antioxidants are the nanoparticles designed for this purpose. Some of these nano-antioxidants have recently shown promise in nanomedicine studies [33,34].

3. Nanoantioxidants

Antioxidants have been proven to be helpful therapeutic and preventative agents for a variety of ailments. Most antioxidants have few drawbacks, such as low permeability, poor water solubility, instability during storage, and digestive tract degradability [35]. Free radical generation has been minimized in several fields by developing nanoparticles, and synthesized nanoparticles are known as nano-antioxidants due to the integration of material sciences and nanotechnology [31,36].
Nanoantioxidants are nanomaterials that capture chain-carrying radicals or reduce the number of initiation processes to lower the overall rate of autoxidation. Vitamin E, 3-carotene, selenium, glutathione, and polyphenols are the most studied dietary antioxidants as a potential pharmacological strategy to reduce ROS levels and combat disorders caused by oxidative stress. However, there have been no conclusive results from clinical trials on the benefits of α-tocopherol, selenium, or (3-carotene supplementation) for lowering cancer risks [13]. One of the most likely explanations for these surprising outcomes is that most antioxidants fail to reach biologically relevant targets. Nanoantioxidants are a unique possibility in this context since they may be tailored to have more extended stability than small molecules, quickly escape metabolic clearance, and target specific locations [13].
Nanomaterials can either act as passive delivery systems for small-molecule antioxidants or have their built-in antioxidant capabilities. Therefore, a critical step in creating new nano-antioxidants is to conduct in vitro chemical tests to assess their antioxidative activity.
Different forms of antioxidants have been discovered as a result of recent research trends. Nano-antioxidants have been made with a variety of materials until now (Figure 2). Nanoparticles of metal and metal oxide, CNTs and other carbonacious nanomaterials, and polymer-loaded antioxidant nanoparticles have all been reported to possess antioxidant capabilities. Different forms of antioxidants have been discovered as a result of recent research trends. Biodegradable nanoparticles have recently received a great deal of attention due to their high encapsulation efficiency, controlled release properties, and a lack of toxicity, among other factors. As a carrier for various forms of nano-antioxidants throughout history, biodegradable polymers have proven to be the most essential materials to be explored [37,38]. Metal-based nanoparticles with high antioxidant activity are being used to develop new antioxidant materials.
The antioxidant enzymes GSH-Px and thioredoxin reductase use selenium in selenoproteins to regulate hemostasis and the redox system [39]. Furthermore, as documented in several studies, selenium nanoparticles activate selenoproteins and display antioxidant activity in both in vivo and in vitro environments [40]. When administered to rats, nanoselenium has been found to significantly diminish nicotine-induced nephrotoxicity by inhibiting the production of reactive oxygen species and inflammation. This impact has been attributed to the small size of nanoselenium, which has a greater surface-to-volume ratio and, as a result, has better bioavailability than larger particles [41].
A dose-dependent antioxidant activity has been demonstrated for ZnO nanoparticles produced from Polygala tenuifolia root extracts [42]. Antibacterial and anticancer characteristics of silver nanoparticles have made them popular for medical applications. They are used in a variety of applications. The use of plant-based silver nanoparticles, on the other hand, has the advantage of having free radical scavenging action, which is becoming increasingly important in the treatment of oxidative stress-mediated toxicities [43].
Additionally, carbon-based nanoparticles have the potential to perform a radical cleaning up activity. Carbon nanotubes, for example, have antioxidant properties due to their strong affinity for electrons [44]. Specifically, single-walled carbon nanotubes exhibit significant antioxidant effects in the oxygen radical absorbance capacity assay, which has been validated [45].
A variety of nanoparticles coated with antioxidant CNTs and metal or metal-oxide nanoparticles have antioxidative properties [36]. Nano-oxidant synthesis has used various preparation strategies over the last few decades, including solvent/emulsion drying, solvent displacement, nanoprecipitation, templating, and solvent displacement methods [36].
Due to their intrinsic physicochemical features, reactive nitrogen and oxygen species can be cleaned up by using oxide nanoparticles, which mimic the antioxidant molecule [46]. There has been a great deal of interest in the biomedical field for cerium oxide nanoparticles (CONPs) because of their multi-enzymatic ROS scavenging abilities [47]. In addition to coexisting in both oxidation states, these CONPs exhibit a reduction potential of 1.52 V and can flip between both oxidation states [48]. However, Cerium dioxide in the form of a bulk crystal consists of Ce4+, whereas when it is reduced to nano-size, it considerably increases the relative proportion of Ce3+, resulting in improved catalytic activities. After three weeks of intravenous administration, researchers found that nanoceria dramatically reduced lipoperoxidation, demonstrating that CONPs effectively treat oxidative stress in an in vivo test on mice conducted by Hirst et al. (2013) [49]. Caputo et al. (2015) compared the antioxidant capacity of CONPs with NAC (N-acetyl-cysteine) and Trolox (soluble analogues of vitamin E) in a study [50].
Additionally, liposomes are used to deliver antioxidants to their intended areas. Since these liposomes are both amphiphilic and biocompatible, they can hold a variety of chemicals, including antioxidant enzymes that are both water-soluble and water insoluble [51]. Chitosan is the most commonly used substance to make nanoparticles, whether as a standalone substance or in combination with another [52]. Chitosan has mucoadhesive characteristics, rendering it superior for delivering drugs to mucosal surfaces, including the intestinal and nasal epithelium [53]. It has been found that curcumin coated with chitosan, which protects it from free radicals, is more effective in scavenging free radicals than curcumin alone [54]. By adjusting the pH and oxidative stress levels of inflamed tissues, Pu et al. (2014) found that curcumin-based nanocarriers can encapsulate curcumin antioxidant molecules and regulate their release, which in turn increases the generation of RNS/ROS by lipopolysaccharide-stimulated macrophages [55].

4. Classification of Antioxidant Action

4.1. Preventive Antioxidants

Preventive antioxidants work by lowering the rate of initiation [56]. This is a heterogeneous class of compounds that comprise scavengers [57], metal chelators [58], agents capable of degrading hydroperoxide [59], and both smaller molecules and enzymes such as superoxide dismutase (SOD) and glutathione peroxidase (GPX) or their analogs [60].

4.2. Chain-Breaking Antioxidants

An antioxidant (AH) and a peroxyl radical can create a hydroperoxide and a radical of the antioxidant (A•), which also trap a second ROO• and results in the production of non-radical end products if a formal H-atom transfer occurs between them [10]. Antioxidants that break chains are destroyed during the reaction. Therefore, they function stoichiometrically when regeneration mechanisms are absent. In this context, it is essential to keep in mind that chain-breaking antioxidants exclusively react with peroxyl radicals, but the interactions with other radical species produced during autoxidation are of lesser importance. Because R• interacts so quickly with oxygen, an antioxidant’s chances of catching it are minimal to none. With regard to the OH- and RO-initiating radicals, these radicals have extremely high reactivity towards all organic molecules, making it impossible for the antioxidant to compete with these reactions because the antioxidant, which, by definition, is only found in trace quantities, is unable to compete with these reactions [56]. Figure 3 shows characteristics of a molecule (AH) that makes it a chain-breaking antioxidant.
The most effective examples of this family of antioxidants may be phenols (such as resveratrol) and ascorbate (vitamin C), whereas synthetic equivalents include aromatic amines, BHT, BHA, and the nitroxide compound. Figure 4 despicts system for classifying nanoantioxidants based on structural and mode of action.
Besides direct antioxidants, substances that do not themselves have antioxidant activity but can promote and enhance the effectiveness of biological systems’ endogenous antioxidant defenses are usually classed as indirect antioxidants.

5. Nanomaterials with Inherent Antioxidant Capabilities

There are variety of nanomaterials with inherent antioxidant capabilities that do not need to be functionalized with antioxidants; instead, their properties come from the surface features of the nanomaterials. Nanoparticles of inorganic metal are the most common type of such materials. On the other hand, organic nanoparticles are becoming increasingly essential, with a few examples below.
Metals such as silver, gold, platinum, and palladium, as well as metal oxide, are examples that can be found throughout the literature [46,61,62,63,64,65,66]. NPs act as catalase substitutes, but only at neutral or basic pH levels; at acidic pH values, on the other hand, a pro-oxidant impact starts to develop. The production of HO• radicals causes the prooxidant activity of the nanoparticle as a result of the occurrence of the Fenton reaction on the nanoparticle’s surface [67].
H2O2 + e → HO• + HO
Colorimetric glucose or DNA sensing uses this reaction for nanomaterials with peroxidase activity [68]. These nanoantioxidants have a two-sided activity, restricting their effectiveness and providing an exciting opportunity to build pH-responsive redox modulators of the cell’s oxidative stress [69].
There is still some debate about the mechanisms that underlie catalase activities in cerium oxide nanoparticles, although it appears that very stable surface peroxo/hydroperoxo species are at play [70]. The degree of oxidation of a nanomaterial may have an effect on its catalytic activity. Highly persistent surface peroxo/hydroperoxo species are implicated in the case of cerium oxide nanoparticles in catalase activity; however, the processes behind catalase activity are still a mystery [70,71]. The degree of oxidation of a nanomaterial determines whether or not it exhibits CAT-like behavior [70].
The material with a higher Mn3+/Mn2+ ratio, generated by oxidation with NaIO4, showed improved CAT activities in comparison to materials with a lower Mn3+/Mn2 ratio in Mn3O4 NPs with flower-like morphology (“nanoflowers”) [63].
When it comes to Co3O4 nanomaterials, nanoplates were the most common, followed by nanorods and nanocubes. For three distinct Co3O4 nanomaterials, the amount of CAT-like activity reduced inversely due to their redox potential. In order to explain this result, it was found that the rate-determining step in the CAT catalytic cycles is the cleavage of the O-O bond that occurs during the nanomaterial’s reduction of H2O2 [72]. In low pH environments, switching from Fenton chemistry to catalase activity is controlled by the pH of the environment (high pH). A possible explanation for this unusual behavior is the pH dependency of the interaction between HO• and HOOH (Equation (2)), which is predicted to be aided by the partial deprotonation of H2O2 (pKa = 11.6; see Equation (2)) [67].
H2O2 + HO• → O2•− + H2O + H+
Changes in pH in gold nanorods, core–shell gold–platinum nanorods, and gold–palladium nanorods are thought to result from the metal surface absorbing H+ or HO [73].
Unlike catalase activity, which has been found in many oxides, glutathione peroxidase (GPX) activity has only been found in vanadium [74] and manganese [63] oxides (see Table 1).
GPX Mimics: Until now, GPX activity has only been identified in vanadium [74] and manganese [63] oxides, in contrast to the ubiquitous catalase activity present in many metals and metal oxides.
Glutathione (GSH) is sacrificed when V2O5 nanowires are used in physiological settings to mediate the H2O2 to H2O reduction because of V2O5’s unique ability instead of producing polar peroxido species in comparison to HO• radicals (Equation (3)) [85].
2 GSH + H2O2 —* GS-SG + 2 H2O
Compounds with heavy chalcogen atoms, such as selenium and tellurium, are more likely to exhibit GPX-like activity. GPX-like activity was observed in graphene oxide-supported selenium nanoparticles, likely attributable to their high surface area and high rate of reaction [78].
SOD Mimics: Due to the fact that although the superoxide radical is a member of the peroxyl radical family, its unique chemistry distinguishes it from other radicals such as the alkylperoxyl radical, and antioxidants that trap the radical must be taken into consideration independently.
As the conjugated acid (HOO•) has a pKa of 4.5, O2• is the more common form of superoxide. Superoxide, in its protonated (neutral, HOO•) form, exhibits dualistic behaviour, as it can either abstract or donate a H atom to form HOOH and O2, respectively [89] when the superoxide is protonated (in the neutral state). On the other hand, deprotonated O2• has the majority of the properties of a reducing agent.
A diverse variety of nanomaterials, including noble metals (gold, platinum, and palladium) as well as metal oxides (cerium, cobalt, and manganese oxides), carbon clusters (carbon nanotubes, and fullerenes), and melanin, have been shown to have SOD-like activity.
SOD- and GPX-like activity were concurrently achieved by a “multi-nanozyme” based on MnO2 nano-nanoparticles deposited on V2O5 nanowires by polydopamine [90].
Instead of chain-breaking antioxidant activities, a given nanomaterial must be capable of quenching alkylperoxyl radicals by converting them to hydroperoxides [79,80,81,82]. Alkylperoxyl radicals (ROO•, where R is an alkyl, such as lipids) can be quenched by either an H-atom donating antioxidant AH or an electron-donating antioxidant (D) and a protic solvent SolvH, as shown in Figure 5.
In the same manner, nanoparticles with cleavable O-H groups, such as lignin and melanin, behave. The antioxidant activity of lignin nanoparticles has been observed in apolar polymers such as natural rubber and methanol, for example [91].

6. Antioxidant Functionalized Nanoparticles

Bioactive substances such as terpenoids, alkaloids, polyphenols, and phenolic acids have been found in bacteria, algae, fungi, lichens, and plants. These bioactive chemicals could lead to improved results. The antioxidative properties of these compounds have been demonstrated to decrease and stabilize metallicions.
Table 2 presents the various types of antioxidant nanoparticles with functionalized antioxidants from different biological sources.

7. Inert Scaffold with Antioxidant Functionalities

About fifteen years ago, researchers Liu and coworkers coupled Trolox, a synthetic a-tocopherol analogue, or salvianic acid [147] to gold nanoparticles encapsulated with thiol to demonstrate the efficacy of antioxidants grafted onto a nanomaterial [148].
The radical trapping ability of Au@Trolox was shown to be superior to that of Trolox itself, indicating that this technique will not decrease the antioxidant characteristics of the compound. The food-grade antioxidants caffeic acid [149] and gallic acid [150] were covalently attached to SiO2 NPs of different sizes by using aminopropyl-triethoxysilane (APTES) as the linker. Due to the fact that silica is regarded biochemically inert, it is used as a flowing aid in the production of nutraceutical and pharmaceutical products. Gallic acid coupled to SiO2 demonstrated excellent radical trapping capabilities and was capable of being reused without losing its activity. For the first time, however, the authors demonstrated that, following the reaction with free radicals, these nanoparticles exhibited considerable agglomeration, which they attributed to the presence of cross coupling between free radicals on the surface [151].
Similarly, the phenolic antioxidants, Trolox or curcumin, were covalently linked to graphite-coated cobalt magnetic nanoparticles using a similar technique [152]. Alternatively, on the outside of halloysite nanotubes [153,154]. The tubular hollow structure of inorganic alumino-silicate clay halloysite (HNTs) (up to 800 nm in length and 80 nm in external diameter) is made up of siloxane groups (moderate acidity) on the outer surface and aluminol (basicity) on the inner surface, allowing for a variable selectivity of molecules grafted onto either surface. Halloysite nanotubes were employed to covalently attach curcumin to the surface using a disulfide bridge, allowing for thiol-dependent release while preserving curcumin’s free radical trapping capabilities [153]. To create a bi-functional nanoantioxidant, a synthetic tocopherol (Trolox) was grafted onto the HNT’s external surface using APTES as a linker (HNT-Trolox). This was combined with quercetin, a naturally occurring polyphenolic antioxidant, to create a hybrid nanoantioxidant [154] (Figure 6). With regards to the model’s substrate peroxidation, this material demonstrated good antioxidant capabilities due to Trolox acting as the primary radical quencher and quercetin serving as a co-antioxidant.
Cerium nanoparticles can act as antioxidant enzyme mimics in addition to the scavenging ability of ROS and RNS. Nanoscale physical and chemical properties such as oxygen absorption/release capability and the relative thermodynamic efficacy of redox cycling between cerium nanoparticle surface Ce3+ and Ce4+ are responsible for this phenomenon [46,155]. Combining CNPs with curcumin in a single formulation may result in increased physiological activities due to curcumin’s anticancer properties. Curcumin-loaded nanoceria (CNP-Cur) and dextran-laden nanoceria (Dex-CNP-Cur) were studied in MYCN-amplified and non-amplified cells for their anticancer effects [155]. Antioxidant nanoparticle efficacy testing methods and mechanisms are summarized in Table 3.

7.1. Gold Nanoparticles (AuNPs)

Recently, interest in gold nanoparticles (AuNPs) has grown because of their size, shape, and optical properties, which are biocompatible [172]. Gold nanoparticles of various sizes and shapes have found medicinal applications in cancer detection and medicine administration (e.g., Paclitaxel) [172]. Table 2 reveals that antioxidant AuNPs are typically obtained from plant parts such as leaves and fruits and that they are particularly effective in combating free radicals. Markus et al. (2016) employed an internal membrane-bound technique to identify Kimchi-isolated probiotic strain Lactobacillus kimchicus DCY51T and then used this bacteria to produce antioxidant functionalized AuNPs [101]. When exposed to malignant cell types, such as human colorectal adenocarcinoma (HT29) and murine macrophages, AuNPs developed an amino acid capping layer, and the proteins attached to their surface rendered them harmless (RAW264.7). It has been discovered that biologically synthesized gold nanoparticles are superior scavengers of free radicals, notably DPPH, when compared to gold salts. In this study, techniques such as FT-IR and UV-visible spectra HRTEM were used to confirm the identification and usage of an Enterococcus species that lives in food for the synthesis of gold nanoparticles [103]. Extracellular proteins from Escherichia coli were employed to create anisotropic gold nanoparticles by Veeraapandian et al. (2012) [115]. When it comes to the size and structure of AuNPs, the amount of protein present has a significant impact. Extracellular proteins operate as a capping agent for nanoparticles, increasing their shelf life and improving their stability. When it comes to the size and structure of AuNPs, the amount of protein present has a significant impact. Extracellular proteins operate as a capping agent for nanoparticles, increasing their shelf life and improving their stability. Manjunath et al. (2017) produced gold nanoparticles using a fungal endophyte, Cladosporium cladosporioides, isolated from the seaweed Sargassum wightii [138]. The NADPH-dependent reductase enzyme was shown to be responsible for the utilization of phenolic compounds in the reduction of gold metal salts into gold nanoparticles. Using extracts from the mushroom Inonotus obliquus, Lee et al. (2015) produced AuNPs without the use of harmful chemicals [173]. Researchers made AuNPs from extracts of Gracilaria corticate, which is a marine red alga used as a reducing agent [174]. Sharma et al. (2014) reduced and stabilized AuNPs using the dried biomass of Lemanea fluviatilis [175]. Gold nanoparticles were synthesized by Debnath et al. (2016) from dried lichen biomass collected in the Eastern Himalayas at high elevations, without the need of any external stabilization or reducing chemicals [159]. To compare the antioxidant properties of the two gold nanoparticles, those produced from Acroscyphus sp. lichen possessed prismatic and multiplied twinned quasi-spherical shapes, while those made from Sticta sp. algae only had multiplied twinned quasi-spherical shapes.
DPPH-radical-induced free radical scavenging was investigated in both light and dark settings using a gold nanoparticle (AuNP) attached to cellulose fiber (UKP), which is an unbleached kraft paper [176]. The surface functionalization of AuNPs coated with polyethylene glycol PEG and the antioxidant salvianic acid A (Au@PEG3SA) (Figure 7) was successfully performed and their antioxidant activities were evaluated. Compared to pure salvianic acid, the radical scavenging activity was nine times higher with Au@PEG3SA added to the mix. Studying the radical scavenging activity in vitro and in vivo using stopped-flow analysis, laser scanning confocal microscopic inspection, and the thiobarbituric acid reactive substance assay allowed researchers to better understand how radical scavenging activity works. It has been found that AuNPs with antioxidant functions are more effective at scavenging ROS in living cells and could be employed in such applications [147].
Trolox, a vitamin E analogue, was used to produce functionalized AuNPs (Au@Trolox) via Au-S bonding aided by Trolox’s thiol (SH) ligand in combination with AuNPs (Figure 8) [148]. Additionally, the inclusion of AuNPs embedded in 3,6-dihydroxyflavone along with other nutrients such as lutein and selenium methly selenocysteine increased maximum inhibition by 87.13% for the reduction of DPPH, 85.11% for the reduction of OH, and 84.02% for the reduction of NO, with an overall increase in antioxidant activity of 29.53%, 26.61%, 25.45%, and 26.07%, respectively [24].

7.2. SiO2 Nanoparticles

Due to their mechanical stability, optical and chemical transparency, biocompatibility, and scalability, silica (SiO2) nanoparticles have numerous applications in chemistry, medicine, pharmaceuticals, and biomedical research [149,161,177]. Nanosized SiO2 particles can be immobilized with antioxidants to produce high-value hybrid nanocomposites. Natural antioxidants, such as GA-developed nanoantioxidants, were covalently grafted onto the surface of commercially available well-characterized SiO2NPs (8–30 nm in diameter). The ability of these nano-antioxidants to neutralise DPPH radicals was tested in a single assay (Figure 9). The SiO2-GA nanoparticles execute quick H-atom transfer (HAT) and secondary/slow radical–radical coupling processes, which are two different types of scavenging reactions. The rapid HAT reactions of all SiO2NPs can be compared to those of pure GA at stoichiometry ratios of 2 (nfast) [151]. Morin (2′,3, 4′,5,7-pentahydroxyflavone, a flavonoid)-functionalized mesoporous SiO2NPs (MSN) were also tested for their antioxidant properties as HO• scavengers and O2 quenchers. On the other hand, the nanoantioxidant composite showed a one order-of-magnitude lower inactivation of O2 in homogenous fluids and lipid membranes than morin. It was found that the nanoantioxidant composite had a synergistic effect on antioxidant property against HO• that was proportional to morin adsorbed in homogeneous solvents and lipid membranes, compared to morin’s one-order-of-magnitude lower inactivation of O2. Mesoporous poly(tannic acid) (TA) crosslinked SiO2NPs composites were examined in the same manner for their antioxidant properties, with TA concentrations ranging from 50 mg to 1000 mg and reaction times ranging from 2 to 24 h. Particle sizes varied from 237 to 445 nm as TA concentration and reaction time increased, leading to a decrease in thermal stability of the composite. The highest TA content (1000 mg) and the largest surface area (872 m2/g) were obtained after 12 and 2 h of reaction, respectively; this material also proved to be the most active antioxidant, with a total phenol content of 140.3 g mL−1 GA equivalent and 68 6 mm Trolox (a vitamin E analogue) equivalent [160].
However, MSN coated with caffeic acid (MSN-CAF) or rutin (quercetin 3-O-[-l-rhamnosyl-(1-6)-d-glucopyranoside]) was examined for its antioxidative stress properties. To prevent ROS formation and biological damage, these catechol antioxidants have ortho- and meta-hydroxyl groups in their catechol structure that can scavenge or donate hydrogen atoms to free radicals. Caco-2 and HaCaT cell lines were employed to evaluate ROS generation, the activation of the Keap1-Nrf2 pathway, and cell death induced by free and functionalized MSNs in two cellular models (Figure 2). The antiradical test (oxygen radical absorbance capacity (ORAC)) was also used to investigate the antiradical characteristics of free or grafted antioxidant compounds. MSN-RUT had a Trolox equivalent level at 30.3 g/mL, which is much higher than that of naked nanoparticles or amino-propyl functionalized silica nanoparticles (MSNNH2), and its antiradical function was determined to be 3.7 times more than that of MSN. However, MSN coated with caffeic acid (MSN-CAF) or rutin (quercetin 3-O-[-l-rhamnosyl-(1-6)-d-glucopyranoside]) was examined for its antioxidative stress properties. To prevent ROS generation and biological harm, these phenolic antioxidants can scavenge or donate hydrogen atom to free radicals because of the ortho- and meta-hydroxyl groups in the catechol structure they possess. Caco-2 and HaCaT cell lines were employed to evaluate ROS generation, the activation of the Keap1-Nrf2 pathway, and cell death induced by free and functionalized MSNs in two cellular models (Figure 10). The antiradical test (oxygen radical absorbance capacity (ORAC)) was also used to investigate the antiradical characteristics of free or grafted antioxidant compounds. At 30.3 g/mL, Antiradical activity of MSN-RUT was found to be 3.7 times greater than that of MSN-CAF, according to the ORAC test, which found that the Trolox equivalent level of MSN-CAF and MSN-RUT was significantly higher than that for naked nanoparticles or amino-propyl functionalized silica nanoparticles (MSNNH2) [149].

7.3. Silver Nanoparticles (AgNPs)

Silver nanoparticles (AgNPs) have a significantly greater surface area when compared to chemically identical particles, which is due to their outstanding biochemical and catalytic activity [172]. Colloidal silver nanoparticles are agglomerated to create oligomeric clusters, which are then stabilized in order to make AgNPs [172]. It is necessary to use biological catalysts (enzymes) in order to reduce the Ag+ ions, and various plant extracts can be used to create AgNPs with various antioxidant properties. It is necessary to use biological catalysts (enzymes) in order to reduce Ag+ ions, and various plant extracts can be used to create AgNPs with various antioxidant properties, as shown in Table 3. Patra et al. (2016) synthesized AgNPs at room temperature under light-exposed conditions using the aqueous watermelon rind component. Silver nanoparticles with potential antioxidant capabilities have been discovered in actinobacteria, and whole-cell biomass and cell free extract of Streptomyces naganishii MA7 and Streptomyces griseorubens AU2 have been used to create silver nanoparticles with potential antioxidant characteristics [102,117]. Antioxidant-functionalized AgNPs have been synthesized using fungal species such as Aspergillus versicolor ENT7, Cladosporium cladosporioides, and Pestalotiopsis microspore [139,141,148]. Researchers have also discovered that Ganoderma lucidium is an important plant source of antioxidant-functionalized AgNPs [178,179,180]. Venkatesan et al. (2016) synthesized silver nanoparticles using Ecklonia cava extracts [181]. Ecklonia cava is a marine alga that has been identified as a source of phenolic compounds capable of acting as capping and reducing agents. AgNPs were synthesized using extracts of Parmeliopsis ambigua, Punctelia subrudecta, Evernia mesomorpha, and Xanthoparmelia plitti [182]. AgNPs produced from Clerodendrum phlomidis leaf extracts have a ferric-reducing potential of 1.63 AU, which is higher than the leaves extract alone [183]. The AgNPs produced from Clerodendrum phlomidis leaves had a higher scavenging activity (55.86 g/mL) than the extract (which had a higher IC50 value of 1920 g/mL). The DPPH radical scavenging activity of AgNPs was likewise discovered to be dose-dependent, with the highest inhibition (85.74 percent) being higher than that of the extract alone. AgNPs had a lower IC50 value (9.12 g/mL) than the extract (388.4 g/mL) and standard ferulic acid (182.8 g/mL), showing that they had good antioxidant properties. Das and colleagues discovered that AgNPs produced from Morus alba leaves extract improved DPPH scavenging activity to 47.81 percent, compared to 56 percent for conventional ascorbic acid at the same dose, in a study published in 2019. AgNPs mediated by plant extracts, on the other hand, demonstrated 95.08 percent ABTS+ scavenging activity, which is equivalent to the 95.51 percent at 100g/mL reported for the BTH standard, showing that they have a high scavenging capacity. AgNPs scavenged 64.04 percent of nitric oxide at 100 g/mL, compared to 45.72 percent and 88.62 percent for plant extracts and gallic acid standards, respectively. On the other hand, when applied at a 100 g/mL concentration, AgNPs demonstrated a statistically significant superoxide scavenging activity of 81.92 percent, which was equivalent to 85.35 percent when applied to the tocopherol standard [184].
Due to their extensive properties, AgNPs offers a plethora of medical science applications. Synthesizing AgNPs, numerous methods, including new green approaches, can provide environmentally friendly capping agents that are not harmful and toxic [185]. Different green synthesis techniques were used, including lignin-capped silver nanoparticles (LCSN) (grafted on lignin with a spherical shape and a size range of 10–15 nm) [163]. In the presence of sunlight, Sida cordifolia leaf extracts stimulated the biogenic production of AgNPs [186], the aqueous extract of Clerodendrum phlomidis L. leaves [183], Seabuckthorn (SBT) leaf extracts employed by AgNPs (SBT@AgNPs) [187], and poly(vinyl alcohol)-embedded AgNPs (PVA-AgNPs) [188], which have been found to have antioxidant activities.

7.4. Copper Oxide Nanoparticles

For nanoparticle synthesis, copper has recently caught interest because of its availability and desirable properties such as catalysis, electrical, and optical properties [189,190,191]. In modern technology, as an essential inorganic substance, copper oxide is widely employed, especially for ceramics, catalysis, and superconducting applications. This material can also operate as an electrode to degrade nitrous oxide with ammonia and oxidize carbon monoxide, hydrocarbon, and phenol to produce supercritical water [192]. Cu2ONPs with antioxidant properties synthesized by plant-derived synthesis are summarized in Table 2. Using leaf extracts from Cissus arnotiana, Rajeshkumar and colleagues found that copper nanoparticles have similar radical scavenging properties as ascorbic acid [193].

7.5. Iron Nanoparticles (INPs)

The unique physiochemical properties of iron, such as its low toxicity, good magnetic properties, microwave absorption capabilities, and high catalytic activity, make it a crucial component in nanoparticle synthesis [194,195,196]. The three types of iron nanoparticles are as follows: iron oxide nanoparticles (IONPs), iron hydroxide nanoparticles (FeOOH), and zero-valent iron nanoparticles (INPs) [197,198,199,200]. These new technologies include hyperthermia, bio-separation and bioprocess intensification, ferrofluids for medication delivery and environmental remediation, food preservation and gene therapy, pigments, and thermal-ablation using lithium-ion batteries [201]. The antioxidant capacity of INPs mediated by plants is listed in Table 2. Amaranthine and phenolic chemicals found in the Amaranthus spinosus leaf extract mediate iron nanoparticles, which have high antioxidant activities [95] as a capping agent [93].
The antioxidant properties of Fe2O3 nanoparticles have already been studied, and the theory behind it is based on the neutralization of free radicals by the transfer of an electron [202]. However, numerous techniques, such as carbon coating, carboxymethyl-inulin, and poly (GA), modifying the surface with gallic acid, and the incorporation of curcumin in magnetic-silk core-shell nanoparticles, were successful in tailoring Fe2O3 nanoparticles. The composites showed improved dispersibility and stability. They were examined for their efficient antioxidant characteristics, antibacterial activities, targeted drug administration to specific organs, and their cytotoxicity and biocompatibility/hemocompatibility [164,165,166,167,168,203]. Magnetically separable ligand-functionalized magnetite with natural antioxidant gallic acid (GA) displayed antimicrobial and antioxidant activities (Figure 11).
Post-functionalization techniques successfully synthesized IONP functionalized with multipotent antioxidants (MPAO). PASS analysis and ADMET studies using the structure-based virtual screening technique were used to investigate and anticipate the likely bioactivities and safety profile of the MPAO molecule prior to the synthesis procedure. Magnetite was shown to retain its characteristics after being functionalized with MPAO. The free radical scavenging and antibacterial activities of MPAO functionalized IONP were promising [204,205].

7.6. Selenium Nanoparticles

As part of their research in 2013, Li and his colleagues developed 6-hydroxy-2,5,7,8-tetramethylchroman-2-carboxylic acid (Trolox)-coated surface-functionalized selenium nanoparticles (Se@Trolox) with antioxidant activities [206]. Furthermore, Se@Trolox was discovered to inhibit the activation of the AKT and MAPK signaling pathways, the buildup of reactive oxygen species (ROS) generated by cisplatin, and the phosphorylation of the p53 gene in HK-2 cells after DNA damage [206]. Spirulina and biogenic SeNPs have recently been the subject of research into their medicinal potential. A dose-dependent ABTS/DPPH radical removal was achieved by SeNPs, which are capable of scavenging them. The biological activities of Spirulina and SeNPs were found to be linked to their total phenolic content [130]. The antioxidant capability of plant-derived nickel oxide nanoparticles is demonstrated in Table 2.

7.7. Bimetallic Nanoantioxidants

The ability to scavenge ROS or RNS was studied in various combinations of bimetallic nanocomposites in addition to antioxidant-functionalized materials. The antioxidant activity of the stabilized, mono-dispersed Ag-Se bimetallic nanoparticles (Ag-Se) was examined in vitro using the azino-bis-ethyl benzthiazoline-sulfonic acid (ABTS), 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) and DPPH assays [171]. Gallic-acid-functionalized GA-SiO2@Ag particles demonstrated significantly increased proton-coupled electron transport at near-IR wavelengths (700–1100 nm) when observed at 700–1100 nm (PCET) [150].
It was found that silica-coated AuNPs (Si@AuNP) and PAPM (polyaspartic acid-based polymer micelles) were both glucose sensitive and efficient in targeting the cell. The nanocomposites were made in two steps: First, the primary amine-terminated nanoparticle (either Si@AuNP or PAPM) was functionalized with phenylboronic acid, and then vitamin C was chelated to the nanoparticle via chelation (Figure 12). The resulting Si@AuNP and PAPM have hydrodynamic sizes of 40–50 nm and 40–80 nm, respectively, and contain vitamin C concentrations of 4–8 weight percent and 10–13 weight percent. At micromolar concentrations, vitamin C protects the cell, whereas at millimolar concentrations, it induces oxidative stress and ultimately cell death through the production of H2O2.
Zingiber officinale (ginger) rhizome extract was used as a capping and reducing agent in the synthesis of copper and nickel hybrid nanoparticles. Nanoparticles at a concentration of 200 g/mL were tested against 2,2-diphenyl-1-picrylhydrazyl free radicals, and NPs showed a 42.1 ± 0.71% percent suppression of the free radicals [207].

8. Conclusions

Numerous treatments have been developed to counteract the negative impact that oxidative stress has on the human body. However, due to many factors, including their inability to cross the blood-brain barrier, conventional antioxidant therapy has been less effective. Therefore, using natural and synthetic antioxidants for the treatment of oxidative stress-induced diseases is no longer recommended. Recent nanoantioxidant research has demonstrated that inorganic nanoparticles can reduce oxidative stress more effectively than organic nanoparticles, with improved sensitivity, cellular antioxidant activity, and the lowest amount of cytotoxicity, as well as the targeted delivery of the nanoantioxidant to specific cells. For example, the covalent attachment or encapsulation of antioxidants with nanospheres derived from various sources has been investigated and is being considered for multiple applications. There are many different types of nanoparticles available, including inorganic, metal, and natural polymer-based nanoparticles; liposomes; and protein polysaccharide-based nanoparticles, to name a few. For nanoantioxidant composites to perform at their peak levels of catalytic and biological activity, it is vital to have a deeper understanding of their origin, physical and chemical properties, and their mechanism of action. Before any potential biomedical applications of nonbiodegradable and insoluble nanoparticles may be considered, extensive toxicity testing must be performed.
Nanoantioxidants synthesized in the lab must also be evaluated in vivo to ensure their safety, especially for long-term treatments, by identifying and analyzing their beneficial and harmful properties. The development of novel and effective delivery strategies for therapeutic nanoantioxidants is vital, as is the development of novel antioxidant activity tests that are accurate and dependable. Research into new compounds as well as the development of very well nanostructures and nanotechnology will play a key role in determining the future of nanoantioxidant-mediated therapy methods.

9. Recommendation for Future Research

The use of natural and synthetic antioxidants for the treatment of oxidative stress-induced disorders is now obsolete. Recently, nanoantioxidants have demonstrated the capacity to minimize oxidative stress with improved sensitivity, cellular antioxidant activity, lower cytotoxic effects, and tailored delivery in inorganic nanoparticles developed over the previous few decades. Antioxidants can be covalently attached or encapsulated with nanospheres of many types, such as inorganic nanoparticles, metal nanoparticles, and many more combinations, for various uses. However, a thorough understanding of the nanoantioxidant composites’ origin, physicochemical properties, and the mechanism of action is required to acquire the best catalytic and biological activity. In addition, comprehensive toxicity testing for nonbiodegradable and insoluble nanoparticles is required before any further biomedical application can be carried out. In addition, it is necessary to identify and assess the benefits and negative effects of the synthesised nanoantioxidants before they can be properly utilized in vivo, particularly for long-term treatment.

10. Limitation of This Study

There is a need for new nanoantioxidants and new methods for measuring their antioxidant activity that are more precise and dependable. Because of this, nanoantioxidant treatments in the future will be shaped by advances in molecular knowledge and the development of new nanostructures.

Author Contributions

Conceptualization, S.T.S., Z.Z.C. and K.S.; writing, S.T.S., Z.Z.C., M.H. and W.J.B.; formal analysis, S.T.S., Z.Z.C. and K.S.; investigation, S.T.S., Z.Z.C. and W.J.B.; resources, I.A.B., S.K. and H.A.; data curation, S.T.S., Z.Z.C., M.H. and W.J.B.; writing—original draft preparation, S.T.S., Z.Z.C. and K.S.; writing—review and editing, S.T.S., Z.Z.C., S.K., H.A. and K.S.; visualization, S.T.S., Z.Z.C. and W.J.B.; supervision, Z.Z.C. and W.J.B.; project administration, Z.Z.C., W.J.B. and I.A.B.; funding acquisition, Z.Z.C., M.H., S.K., H.A. and I.A.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work was carried out under the Deanship of Scientific Research at King Khalid University for the project RGP. 2/101/43 and IF073-2019 and ST 077-2022 under University of Malaya, Malaysia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University for funding this work through the Large Groups Project under grant number RGP. 2/101/43. The authors also thank University of Malaya for supporting under the grant IF073-2019 and ST 077-2022.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Seals, D.R.; Justice, J.N.; LaRocca, T.J. Physiological geroscience: Targeting function to increase healthspan and achieve optimal longevity. J. Physiol. 2016, 594, 2001–2024. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Yabluchanskiy, A.; Ungvari, Z.; Csiszar, A.; Tarantini, S. Advances and challenges in geroscience research: An update. Physiol. Int. 2018, 105, 298–308. [Google Scholar] [CrossRef] [PubMed]
  3. Dabhade, P.; Kotwal, S. Tackling the aging process with bio-molecules: A possible role for caloric restriction, food-derived nutrients, vitamins, amino acids, peptides, and minerals. J. Nutr. Gerontol. Geriatr. 2013, 32, 24–40. [Google Scholar] [CrossRef] [PubMed]
  4. Chang, C.H.; Lee, K.-Y.; Shim, Y.H. Normal aging: Definition and physiologic changes. Jkma 2017, 60, 358–363. [Google Scholar] [CrossRef]
  5. Harman, D. Aging: A theory based on free radical and radiation chemistry. J. Gerontol. 1956, 11, 298–300. [Google Scholar] [CrossRef] [Green Version]
  6. Islam, M.T. Oxidative stress and mitochondrial dysfunction-linked neurodegenerative disorders. Neurol. Res. 2017, 39, 73–82. [Google Scholar] [CrossRef]
  7. Valentao, P.; Fernandes, E.; Carvalho, F.; Andrade, P.B.; Seabra, R.M.; de Lourdes Bastos, M. Antioxidant activity of hypericum androsaemum infusion: Scavenging activity against superoxide radical, hydroxyl radical and hypochlorous acid. Biol. Pharm. Bull. 2002, 25, 1320–1323. [Google Scholar] [CrossRef] [Green Version]
  8. Chang, K.-H.; Chen, C.-M. The role of oxidative stress in parkinson’s disease. Antioxidants 2020, 9, 597. [Google Scholar] [CrossRef]
  9. López-Otín, C.; Blasco, M.A.; Partridge, L.; Serrano, M.; Kroemer, G. The hallmarks of aging. Cell 2013, 153, 1194–1217. [Google Scholar] [CrossRef] [Green Version]
  10. Ingold, K.U.; Pratt, D.A. Advances in radical-trapping antioxidant chemistry in the 21st century: A kinetics and mechanisms perspective. Chem. Rev. 2014, 114, 9022–9046. [Google Scholar] [CrossRef]
  11. Cadet, J.; Wagner, J.R. Oxidatively generated base damage to cellular DNA by hydroxyl radical and one-electron oxidants: Similarities and differences. Arch. Biochem. Biophys. 2014, 557, 47–54. [Google Scholar] [CrossRef]
  12. Zhang, H.; Forman, H.J. 4-hydroxynonenal-mediated signaling and aging. Free Radic. Biol. Med. 2017, 111, 219–225. [Google Scholar] [CrossRef]
  13. Morry, J.; Ngamcherdtrakul, W.; Yantasee, W. Oxidative stress in cancer and fibrosis: Opportunity for therapeutic intervention with antioxidant compounds, enzymes, and nanoparticles. Redox Biol. 2017, 11, 240–253. [Google Scholar] [CrossRef]
  14. Ricordi, C.; Garcia-Contreras, M.; Farnetti, S. Diet and inflammation: Possible effects on immunity, chronic diseases, and life span. J. Am. Coll. Nutr. 2015, 34, 10–13. [Google Scholar] [CrossRef]
  15. Young, I.S.; Woodside, J.V. Antioxidants in health and disease. J. Clin. Pathol. 2001, 54, 176. [Google Scholar] [CrossRef] [Green Version]
  16. Mattill, H.A. Antioxidants. Annu. Rev. Biochem. 1947, 16, 177–192. [Google Scholar] [CrossRef]
  17. German, J.B. Food processing and lipid oxidation. Adv. Exp. Med. Biol. 1999, 459, 23–50. [Google Scholar]
  18. Jacob, R. Three eras of vitamin c discovery. Sub-Cell. Biochem. 1996, 25, 1–16. [Google Scholar]
  19. Knight, J.A. Free radicals: Their history and current status in aging and disease. Ann. Clin. Lab. Sci. 1998, 28, 331–346. [Google Scholar]
  20. Halliwell, B. How to Characterize an Antioxidant: An Update. Biochem. Soc. Symp. 1995, 61, 73–101. [Google Scholar]
  21. Shi, H.; Noguchi, N.; Niki, E. Comparative study on dynamics of antioxidative action of α-tocopheryl hydroquinone, ubiquinol, and α-tocopherol against lipid peroxidation. Free Radic. Biol. Med. 1999, 27, 334–346. [Google Scholar] [CrossRef]
  22. Levine, M.; Rumsey, S.C.; Daruwala, R.; Park, J.B.; Wang, Y. Criteria and recommendations for vitamin c intake. JAMA 1999, 281, 1415–1423. [Google Scholar] [CrossRef]
  23. Khan, M.I.; Giridhar, P. Dietary antioxidants: The insurer of health. Everyman’s Sci. 2011, 46, 214–218. [Google Scholar]
  24. Medhe, S.; Bansal, P.; Srivastava, M.M. Enhanced antioxidant activity of gold nanoparticle embedded 3, 6-dihydroxyflavone: A combinational study. Appl. Nanosci. 2014, 4, 153–161. [Google Scholar] [CrossRef] [Green Version]
  25. Fleuriet, A.; Macheix, J.-J. Phenolic acids in fruits and vegetables. Flavonoids Health Dis. 2003, 1. [Google Scholar]
  26. Lim, Y.; Lim, T.; Tee, J. Antioxidant properties of several tropical fruits: A comparative study. Food Chem. 2007, 103, 1003–1008. [Google Scholar] [CrossRef]
  27. Anagnostopoulou, M.A.; Kefalas, P.; Papageorgiou, V.P.; Assimopoulou, A.N.; Boskou, D. Radical scavenging activity of various extracts and fractions of sweet orange peel (citrus sinensis). Food Chem. 2006, 94, 19–25. [Google Scholar] [CrossRef]
  28. Li, L.; Ng, T.; Gao, W.; Li, W.; Fu, M.; Niu, S.; Zhao, L.; Chen, R.; Liu, F. Antioxidant activity of gallic acid from rose flowers in senescence accelerated mice. Life Sci. 2005, 77, 230–240. [Google Scholar] [CrossRef]
  29. Tułodziecka, A.; Szydłowska-Czerniak, A. Determination of total antioxidant capacity of rapeseed and its by-products by a novel cerium oxide nanoparticle-based spectrophotometric method. Food Anal. Methods 2016, 9, 3053–3062. [Google Scholar] [CrossRef] [Green Version]
  30. Verma, A.K. Anti-oxidant activities of biopolymeric nanoparticles: Boon or bane! J. Pharm. Res. 2014, 8, 871–876. [Google Scholar]
  31. Eftekhari, A.; Ahmadian, E.; Panahi-Azar, V.; Hosseini, H.; Tabibiazar, M.; Maleki Dizaj, S. Hepatoprotective and free radical scavenging actions of quercetin nanoparticles on aflatoxin b1-induced liver damage: In vitro/in vivo studies. Artif. Cells Nanomed. Biotechnol. 2018, 46, 411–420. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Yosri, N.; Khalifa, S.A.; Guo, Z.; Xu, B.; Zou, X.; El-Seedi, H.R. Marine organisms: Pioneer natural sources of polysaccharides/proteins for green synthesis of nanoparticles and their potential applications. Int. J. Biol. Macromol. 2021, 193, 1767–1798. [Google Scholar] [CrossRef] [PubMed]
  33. Hasanzadeh, M.; Mokhtari, F.; Shadjou, N.; Eftekhari, A.; Mokhtarzadeh, A.; Jouyban-Gharamaleki, V.; Mahboob, S. Poly arginine-graphene quantum dots as a biocompatible and non-toxic nanocomposite: Layer-by-layer electrochemical preparation, characterization and non-invasive malondialdehyde sensory application in exhaled breath condensate. Mater. Sci. Eng. C 2017, 75, 247–258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Hasanzadeh, M.; Tagi, S.; Solhi, E.; Mokhtarzadeh, A.; Shadjou, N.; Eftekhari, A.; Mahboob, S. An innovative immunosensor for ultrasensitive detection of breast cancer specific carbohydrate (ca 15-3) in unprocessed human plasma and mcf-7 breast cancer cell lysates using gold nanospear electrochemically assembled onto thiolated graphene quantum dots. Int. J. Biol. Macromol. 2018, 114, 1008–1017. [Google Scholar] [CrossRef]
  35. Hu, B.; Liu, X.; Zhang, C.; Zeng, X. Food macromolecule based nanodelivery systems for enhancing the bioavailability of polyphenols. J. Food Drug Anal. 2017, 25, 3–15. [Google Scholar] [CrossRef] [Green Version]
  36. Eftekhari, A.; Dizaj, S.M.; Chodari, L.; Sunar, S.; Hasanzadeh, A.; Ahmadian, E.; Hasanzadeh, M. The promising future of nano-antioxidant therapy against environmental pollutants induced-toxicities. Biomed. Pharmacother. 2018, 103, 1018–1027. [Google Scholar] [CrossRef]
  37. Eftekhari, A.; Ahmadian, E.; Azami, A.; Johari-Ahar, M.; Eghbal, M.A. Protective effects of coenzyme q10 nanoparticles on dichlorvos-induced hepatotoxicity and mitochondrial/lysosomal injury. Environ. Toxicol. 2018, 33, 167–177. [Google Scholar] [CrossRef]
  38. Pohlmann, A.R.; Schaffazick, S.R.; Creczynski-Pasa, T.B.; Guterres, S.S. Preparation of drug-loaded polymeric nanoparticles and evaluation of the antioxidant activity against lipid peroxidation. In Free Radicals and Antioxidant Protocols; Springer: Berlin/Heidelberg, Germany, 2010; pp. 109–121. [Google Scholar]
  39. Mohammed, E.; Safwat, G. Assessment of the ameliorative role of selenium nanoparticles on the oxidative stress of acetaminophen in some tissues of male albino rats. Beni-Suef Univ. J. Basic Appl. Sci. 2013, 2, 80–85. [Google Scholar] [CrossRef] [Green Version]
  40. Kumar, G.S.; Kulkarni, A.; Khurana, A.; Kaur, J.; Tikoo, K. Selenium nanoparticles involve hsp-70 and sirt1 in preventing the progression of type 1 diabetic nephropathy. Chem.-Biol. Interact. 2014, 223, 125–133. [Google Scholar] [CrossRef]
  41. Zahran, W.E.; Elsonbaty, S.M.; Moawed, F.S. Selenium nanoparticles with low-level ionizing radiation exposure ameliorate nicotine-induced inflammatory impairment in rat kidney. Environ. Sci. Pollut. Res. 2017, 24, 19980–19989. [Google Scholar] [CrossRef]
  42. Baranwal, A.; Mahato, K.; Srivastava, A.; Maurya, P.K.; Chandra, P. Phytofabricated metallic nanoparticles and their clinical applications. RSC Adv. 2016, 6, 105996–106010. [Google Scholar] [CrossRef]
  43. Abdel-Aziz, M.S.; Shaheen, M.S.; El-Nekeety, A.A.; Abdel-Wahhab, M.A. Antioxidant and antibacterial activity of silver nanoparticles biosynthesized using chenopodium murale leaf extract. J. Saudi Chem. Soc. 2014, 18, 356–363. [Google Scholar] [CrossRef] [Green Version]
  44. Watts, P.; Fearon, P.; Hsu, W.; Billingham, N.; Kroto, H.; Walton, D. Carbon nanotubes as polymer antioxidants. J. Mater. Chem. 2003, 13, 491–495. [Google Scholar] [CrossRef]
  45. Lucente-Schultz, R.M.; Moore, V.C.; Leonard, A.D.; Price, B.K.; Kosynkin, D.V.; Lu, M.; Partha, R.; Conyers, J.L.; Tour, J.M. Antioxidant single-walled carbon nanotubes. J. Am. Chem. Soc. 2009, 131, 3934–3941. [Google Scholar] [CrossRef]
  46. Nelson, B.C.; Johnson, M.E.; Walker, M.L.; Riley, K.R.; Sims, C.M. Antioxidant cerium oxide nanoparticles in biology and medicine. Antioxidants 2016, 5, 15. [Google Scholar] [CrossRef] [Green Version]
  47. Eriksson, P.; Tal, A.A.; Skallberg, A.; Brommesson, C.; Hu, Z.; Boyd, R.D.; Olovsson, W.; Fairley, N.; Abrikosov, I.A.; Zhang, X.; et al. Cerium oxide nanoparticles with antioxidant capabilities and gadolinium integration for mri contrast enhancement. Sci. Rep. 2018, 8, 6999. [Google Scholar] [CrossRef] [Green Version]
  48. Das, S.; Dowding, J.M.; Klump, K.E.; McGinnis, J.F.; Self, W.; Seal, S. Cerium oxide nanoparticles: Applications and prospects in nanomedicine. Nanomedicine 2013, 8, 1483–1508. [Google Scholar] [CrossRef]
  49. Hirst, S.M.; Karakoti, A.; Singh, S.; Self, W.; Tyler, R.; Seal, S.; Reilly, C.M. Bio-distribution and in vivo antioxidant effects of cerium oxide nanoparticles in mice. Environ. Toxicol. 2013, 28, 107–118. [Google Scholar] [CrossRef]
  50. Caputo, F.; De Nicola, M.; Sienkiewicz, A.; Giovanetti, A.; Bejarano, I.; Licoccia, S.; Traversa, E.; Ghibelli, L. Cerium oxide nanoparticles, combining antioxidant and uv shielding properties, prevent uv-induced cell damage and mutagenesis. Nanoscale 2015, 7, 15643–15656. [Google Scholar] [CrossRef]
  51. Du, L.; Li, J.; Chen, C.; Liu, Y. Nanocarrier: A potential tool for future antioxidant therapy. Free Radic. Res. 2014, 48, 1061–1069. [Google Scholar] [CrossRef]
  52. Hans, M.L.; Lowman, A.M. Biodegradable nanoparticles for drug delivery and targeting. Curr. Opin. Solid State Mater. Sci. 2002, 6, 319–327. [Google Scholar] [CrossRef]
  53. Vila, A.; Sánchez, A.; Tobío, M.; Calvo, P.; Alonso, M.J. Design of biodegradable particles for protein delivery. J. Control. Release 2002, 78, 15–24. [Google Scholar] [CrossRef]
  54. Shah, B.R.; Zhang, C.; Li, Y.; Li, B. Bioaccessibility and antioxidant activity of curcumin after encapsulated by nano and pickering emulsion based on chitosan-tripolyphosphate nanoparticles. Food Res. Int. 2016, 89, 399–407. [Google Scholar] [CrossRef]
  55. Pu, H.-L.; Chiang, W.-L.; Maiti, B.; Liao, Z.-X.; Ho, Y.-C.; Shim, M.S.; Chuang, E.-Y.; Xia, Y.; Sung, H.-W. Nanoparticles with dual responses to oxidative stress and reduced ph for drug release and anti-inflammatory applications. ACS Nano 2014, 8, 1213–1221. [Google Scholar] [CrossRef]
  56. Amorati, R.; Valgimigli, L. Advantages and limitations of common testing methods for antioxidants. Free Radic. Res. 2015, 49, 633–649. [Google Scholar] [CrossRef]
  57. Barajas-Carmona, J.G.; Francisco-Aldana, L.; Morales-Narváez, E. Wearable nanoplasmonic patch detecting sun/uv exposure. Anal. Chem. 2017, 89, 13589–13595. [Google Scholar] [CrossRef] [PubMed]
  58. Perron, N.R.; Brumaghim, J.L. A review of the antioxidant mechanisms of polyphenol compounds related to iron binding. Cell Biochem. Biophys. 2009, 53, 75–100. [Google Scholar] [CrossRef] [PubMed]
  59. Lu, J.; Holmgren, A. The thioredoxin antioxidant system. Free Radic. Biol. Med. 2014, 66, 75–87. [Google Scholar] [CrossRef] [PubMed]
  60. Brand, M.D.; Affourtit, C.; Esteves, T.C.; Green, K.; Lambert, A.J.; Miwa, S.; Pakay, J.L.; Parker, N. Mitochondrial superoxide: Production, biological effects, and activation of uncoupling proteins. Free Radic. Biol. Med. 2004, 37, 755–767. [Google Scholar] [CrossRef] [PubMed]
  61. Mu, J.; Zhang, L.; Zhao, M.; Wang, Y. Co3o4 nanoparticles as an efficient catalase mimic: Properties, mechanism and its electrocatalytic sensing application for hydrogen peroxide. J. Mol. Catal. A Chem. 2013, 378, 30–37. [Google Scholar] [CrossRef]
  62. Pirmohamed, T.; Dowding, J.M.; Singh, S.; Wasserman, B.; Heckert, E.; Karakoti, A.S.; King, J.E.; Seal, S.; Self, W.T. Nanoceria exhibit redox state-dependent catalase mimetic activity. Chem. Commun. 2010, 46, 2736–2738. [Google Scholar] [CrossRef] [Green Version]
  63. Singh, N.; Savanur, M.A.; Srivastava, S.; D’Silva, P.; Mugesh, G. A redox modulatory mn3o4 nanozyme with multi-enzyme activity provides efficient cytoprotection to human cells in a parkinson’s disease model. Angew. Chem. 2017, 129, 14455–14459. [Google Scholar] [CrossRef]
  64. Chen, Z.; Yin, J.-J.; Zhou, Y.-T.; Zhang, Y.; Song, L.; Song, M.; Hu, S.; Gu, N. Dual enzyme-like activities of iron oxide nanoparticles and their implication for diminishing cytotoxicity. ACS Nano 2012, 6, 4001–4012. [Google Scholar] [CrossRef]
  65. Silvestri, B.; Vitiello, G.; Luciani, G.; Calcagno, V.; Costantini, A.; Gallo, M.; Parisi, S.; Paladino, S.; Iacomino, M.; D’Errico, G. Probing the eumelanin–silica interface in chemically engineered bulk hybrid nanoparticles for targeted subcellular antioxidant protection. ACS Appl. Mater. Interfaces 2017, 9, 37615–37622. [Google Scholar] [CrossRef]
  66. Dong, J.; Song, L.; Yin, J.-J.; He, W.; Wu, Y.; Gu, N.; Zhang, Y. Co3o4 nanoparticles with multi-enzyme activities and their application in immunohistochemical assay. ACS Appl. Mater. Interfaces 2014, 6, 1959–1970. [Google Scholar] [CrossRef]
  67. Nosaka, Y.; Nosaka, A.Y. Generation and detection of reactive oxygen species in photocatalysis. Chem. Rev. 2017, 117, 11302–11336. [Google Scholar] [CrossRef]
  68. Wei, H.; Wang, E. Nanomaterials with enzyme-like characteristics (nanozymes): Next-generation artificial enzymes. Chem. Soc. Rev. 2013, 42, 6060–6093. [Google Scholar] [CrossRef]
  69. Wason, M.S.; Colon, J.; Das, S.; Seal, S.; Turkson, J.; Zhao, J.; Baker, C.H. Sensitization of pancreatic cancer cells to radiation by cerium oxide nanoparticle-induced ros production. Nanomed. Nanotechnol. Biol. Med. 2013, 9, 558–569. [Google Scholar] [CrossRef] [Green Version]
  70. Cafun, J.; Kvashnina, K.; Casals, E.; Puntes, V.; Glatzel, P. Absence of ce 3+ sites in chemically active colloidal ceria nanoparticles. ACS Nano 2013, 7, 10726–10732. [Google Scholar] [CrossRef] [Green Version]
  71. Celardo, I.; Pedersen, J.Z.; Traversa, E.; Ghibelli, L. Pharmacological potential of cerium oxide nanoparticles. Nanoscale 2011, 3, 1411–1420. [Google Scholar] [CrossRef]
  72. Mu, J.; Zhang, L.; Zhao, M.; Wang, Y. Catalase mimic property of co3o4 nanomaterials with different morphology and its application as a calcium sensor. ACS Appl. Mater. Interfaces 2014, 6, 7090–7098. [Google Scholar] [CrossRef]
  73. Li, J.; Liu, W.; Wu, X.; Gao, X. Mechanism of ph-switchable peroxidase and catalase-like activities of gold, silver, platinum and palladium. Biomaterials 2015, 48, 37–44. [Google Scholar] [CrossRef]
  74. Vernekar, A.A.; Sinha, D.; Srivastava, S.; Paramasivam, P.U.; D’Silva, P.; Mugesh, G. An antioxidant nanozyme that uncovers the cytoprotective potential of vanadia nanowires. Nat. Commun. 2014, 5, 5301. [Google Scholar] [CrossRef] [Green Version]
  75. He, W.; Zhou, Y.-T.; Wamer, W.G.; Hu, X.; Wu, X.; Zheng, Z.; Boudreau, M.D.; Yin, J.-J. Intrinsic catalytic activity of au nanoparticles with respect to hydrogen peroxide decomposition and superoxide scavenging. Biomaterials 2013, 34, 765–773. [Google Scholar] [CrossRef]
  76. Fan, J.; Yin, J.-J.; Ning, B.; Wu, X.; Hu, Y.; Ferrari, M.; Anderson, G.J.; Wei, J.; Zhao, Y.; Nie, G. Direct evidence for catalase and peroxidase activities of ferritin–platinum nanoparticles. Biomaterials 2011, 32, 1611–1618. [Google Scholar] [CrossRef]
  77. Liu, Y.; Wu, H.; Li, M.; Yin, J.-J.; Nie, Z. Ph dependent catalytic activities of platinum nanoparticles with respect to the decomposition of hydrogen peroxide and scavenging of superoxide and singlet oxygen. Nanoscale 2014, 6, 11904–11910. [Google Scholar] [CrossRef]
  78. Huang, Y.; Liu, C.; Pu, F.; Liu, Z.; Ren, J.; Qu, X. A go–se nanocomposite as an antioxidant nanozyme for cytoprotection. Chem. Commun. 2017, 53, 3082–3085. [Google Scholar] [CrossRef] [PubMed]
  79. Watanabe, A.; Kajita, M.; Kim, J.; Kanayama, A.; Takahashi, K.; Mashino, T.; Miyamoto, Y. In vitro free radical scavenging activity of platinum nanoparticles. Nanotechnology 2009, 20, 455105. [Google Scholar] [CrossRef] [PubMed]
  80. Lee, S.S.; Song, W.; Cho, M.; Puppala, H.L.; Nguyen, P.; Zhu, H.; Segatori, L.; Colvin, V.L. Antioxidant properties of cerium oxide nanocrystals as a function of nanocrystal diameter and surface coating. ACS Nano 2013, 7, 9693–9703. [Google Scholar] [CrossRef] [PubMed]
  81. Balaji, S.; Mandal, B.K.; Ranjan, S.; Dasgupta, N.; Chidambaram, R. Nano-zirconia–evaluation of its antioxidant and anticancer activity. J. Photochem. Photobiol. B Biol. 2017, 170, 125–133. [Google Scholar] [CrossRef]
  82. Ju, K.-Y.; Lee, Y.; Lee, S.; Park, S.B.; Lee, J.-K. Bioinspired polymerization of dopamine to generate melanin-like nanoparticles having an excellent free-radical-scavenging property. Biomacromolecules 2011, 12, 625–632. [Google Scholar] [CrossRef]
  83. Liu, Y.; Ai, K.; Ji, X.; Askhatova, D.; Du, R.; Lu, L.; Shi, J. Comprehensive insights into the multi-antioxidative mechanisms of melanin nanoparticles and their application to protect brain from injury in ischemic stroke. J. Am. Chem. Soc. 2017, 139, 856–862. [Google Scholar] [CrossRef] [Green Version]
  84. Korschelt, K.; Ragg, R.; Metzger, C.S.; Kluenker, M.; Oster, M.; Barton, B.; Panthöfer, M.; Strand, D.; Kolb, U.; Mondeshki, M. Glycine-functionalized copper (ii) hydroxide nanoparticles with high intrinsic superoxide dismutase activity. Nanoscale 2017, 9, 3952–3960. [Google Scholar] [CrossRef] [Green Version]
  85. Ragg, R.; Tahir, M.N.; Tremel, W. Solids go bio: Inorganic nanoparticles as enzyme mimics. Eur. J. Inorg. Chem. 2016, 2016, 1906–1915. [Google Scholar] [CrossRef]
  86. Samuel, E.L.; Marcano, D.C.; Berka, V.; Bitner, B.R.; Wu, G.; Potter, A.; Fabian, R.H.; Pautler, R.G.; Kent, T.A.; Tsai, A.-L. Highly efficient conversion of superoxide to oxygen using hydrophilic carbon clusters. Proc. Natl. Acad. Sci. USA 2015, 112, 2343–2348. [Google Scholar] [CrossRef] [Green Version]
  87. Ge, C.; Fang, G.; Shen, X.; Chong, Y.; Wamer, W.G.; Gao, X.; Chai, Z.; Chen, C.; Yin, J.-J. Facet energy versus enzyme-like activities: The unexpected protection of palladium nanocrystals against oxidative damage. ACS Nano 2016, 10, 10436–10445. [Google Scholar] [CrossRef]
  88. Liu, G.F.; Filipović, M.; Ivanović-Burmazović, I.; Beuerle, F.; Witte, P.; Hirsch, A. High catalytic activity of dendritic c60 monoadducts in metal-free superoxide dismutation. Angew. Chem. 2008, 120, 4055–4058. [Google Scholar] [CrossRef]
  89. Cedrowski, J.; Litwinienko, G.; Baschieri, A.; Amorati, R. Hydroperoxyl radicals (hoo.): Vitamin e regeneration and h-bond effects on the hydrogen atom transfer. Chem.–Eur. J. 2016, 22, 16441–16445. [Google Scholar] [CrossRef]
  90. Ju, E.; Dong, K.; Chen, Z.; Liu, Z.; Liu, C.; Huang, Y.; Wang, Z.; Pu, F.; Ren, J.; Qu, X. Copper (ii)–graphitic carbon nitride triggered synergy: Improved ros generation and reduced glutathione levels for enhanced photodynamic therapy. Angew. Chem. Int. Ed. 2016, 55, 11467–11471. [Google Scholar] [CrossRef]
  91. Barana, D.; Orlandi, M.; Zoia, L.; Castellani, L.; Hanel, T.; Bolck, C.; Gosselink, R. Lignin Based Functional Additives for Natural Rubber. ACS Sustain. Chem. Eng. 2018, 6, 11843–11852. [Google Scholar] [CrossRef]
  92. Harshiny, M.; Iswarya, C.N.; Matheswaran, M. Biogenic synthesis of iron nanoparticles using amaranthus dubius leaf extract as a reducing agent. Powder Technol. 2015, 286, 744–749. [Google Scholar] [CrossRef]
  93. Muthukumar, H.; Matheswaran, M. Amaranthus spinosus leaf extract mediated feo nanoparticles: Physicochemical traits, photocatalytic and antioxidant activity. ACS Sustain. Chem. Eng. 2015, 3, 3149–3156. [Google Scholar] [CrossRef]
  94. Tuzun, B.S.; Fafal, T.; Tastan, P.; Kivcak, B.; Yelken, B.O.; Kayabasi, C.; Susluer, S.Y.; Gunduz, C. Structural characterization, antioxidant and cytotoxic effects of iron nanoparticles synthesized using asphodelus aestivus brot. Aqueous extract. Green Process. Synth. 2020, 9, 153–163. [Google Scholar] [CrossRef]
  95. Periakaruppan, R.; Chen, X.; Thangaraj, K.; Jeyaraj, A.; Nguyen, H.H.; Yu, Y.; Hu, S.; Lu, L.; Li, X. Utilization of tea resources with the production of superparamagnetic biogenic iron oxide nanoparticles and an assessment of their antioxidant activities. J. Clean. Prod. 2021, 278, 123962. [Google Scholar] [CrossRef]
  96. Abdullah, J.A.A.; Salah Eddine, L.; Abderrhmane, B.; Alonso-González, M.; Guerrero, A.; Romero, A. Green synthesis and characterization of iron oxide nanoparticles by pheonix dactylifera leaf extract and evaluation of their antioxidant activity. Sustain. Chem. Pharm. 2020, 17, 100280. [Google Scholar] [CrossRef]
  97. Mohamed, N.; Hessen, O.E.A.; Mohammed, H.S. Thermal stability, paramagnetic properties, morphology and antioxidant activity of iron oxide nanoparticles synthesized by chemical and green methods. Inorg. Chem. Commun. 2021, 128, 108572. [Google Scholar] [CrossRef]
  98. Singh, K.; Chopra, D.S.; Singh, D.; Singh, N. Optimization and ecofriendly synthesis of iron oxide nanoparticles as potential antioxidant. Arab. J. Chem. 2020, 13, 9034–9046. [Google Scholar] [CrossRef]
  99. Chavan, R.R.; Bhinge, S.D.; Bhutkar, M.A.; Randive, D.S.; Wadkar, G.H.; Todkar, S.S.; Urade, M.N. Characterization, antioxidant, antimicrobial and cytotoxic activities of green synthesized silver and iron nanoparticles using alcoholic blumea eriantha dc plant extract. Mater. Today Commun. 2020, 24, 101320. [Google Scholar] [CrossRef]
  100. Srihasam, S.; Thyagarajan, K.; Korivi, M.; Lebaka, V.R.; Mallem, S.P.R. Phytogenic generation of nio nanoparticles using stevia leaf extract and evaluation of their in-vitro antioxidant and antimicrobial properties. Biomolecules 2020, 10, 89. [Google Scholar] [CrossRef] [Green Version]
  101. Markus, J.; Mathiyalagan, R.; Kim, Y.-J.; Abbai, R.; Singh, P.; Ahn, S.; Perez, Z.E.J.; Hurh, J.; Yang, D.C. Intracellular synthesis of gold nanoparticles with antioxidant activity by probiotic lactobacillus kimchicus dcy51t isolated from korean kimchi. Enzym. Microb. Technol. 2016, 95, 85–93. [Google Scholar] [CrossRef]
  102. Baygar, T.; Ugur, A. Biosynthesis of silver nanoparticles by streptomyces griseorubens isolated from soil and their antioxidant activity. IET Nanobiotechnology 2017, 11, 286–291. [Google Scholar] [CrossRef] [PubMed]
  103. Oladipo, I.C.; Lateef, A.; Elegbede, J.A.; Azeez, M.A.; Asafa, T.B.; Yekeen, T.A.; Akinboro, A.; Gueguim-Kana, E.B.; Beukes, L.S.; Oluyide, T.O.; et al. Enterococcus species for the one-pot biofabrication of gold nanoparticles: Characterization and nanobiotechnological applications. J. Photochem. Photobiol. B Biol. 2017, 173, 250–257. [Google Scholar] [CrossRef] [PubMed]
  104. Rizzi, V.; Gubitosa, J.; Fini, P.; Nuzzo, S.; Agostiano, A.; Cosma, P. Snail slime-based gold nanoparticles: An interesting potential ingredient in cosmetics as an antioxidant, sunscreen, and tyrosinase inhibitor. J. Photochem. Photobiol. B Biol. 2021, 224, 112309. [Google Scholar] [CrossRef] [PubMed]
  105. Mobaraki, F.; Momeni, M.; Taghavizadeh Yazdi, M.E.; Meshkat, Z.; Silanian Toosi, M.; Hosseini, S.M. Plant-derived synthesis and characterization of gold nanoparticles: Investigation of its antioxidant and anticancer activity against human testicular embryonic carcinoma stem cells. Process Biochem. 2021, 111, 167–177. [Google Scholar] [CrossRef]
  106. Abdoli, M.; Arkan, E.; Shekarbeygi, Z.; Khaledian, S. Green synthesis of gold nanoparticles using centaurea behen leaf aqueous extract and investigating their antioxidant and cytotoxic effects on acute leukemia cancer cell line (thp-1). Inorg. Chem. Commun. 2021, 129, 108649. [Google Scholar] [CrossRef]
  107. Chang, Y.; Zheng, C.; Chinnathambi, A.; Alahmadi, T.A.; Alharbi, S.A. Cytotoxicity, anti-acute leukemia, and antioxidant properties of gold nanoparticles green-synthesized using cannabis sativa l leaf aqueous extract. Arab. J. Chem. 2021, 14, 103060. [Google Scholar] [CrossRef]
  108. Chen, J.; Li, Y.; Fang, G.; Cao, Z.; Shang, Y.; Alfarraj, S.; Ali Alharbi, S.; Li, J.; Yang, S.; Duan, X. Green synthesis, characterization, cytotoxicity, antioxidant, and anti-human ovarian cancer activities of curcumae kwangsiensis leaf aqueous extract green-synthesized gold nanoparticles. Arab. J. Chem. 2021, 14, 103000. [Google Scholar] [CrossRef]
  109. Varghese, B.A.; Nair, R.V.R.; Jude, S.; Varma, K.; Amalraj, A.; Kuttappan, S. Green synthesis of gold nanoparticles using kaempferia parviflora rhizome extract and their characterization and application as an antimicrobial, antioxidant and catalytic degradation agent. J. Taiwan Inst. Chem. Eng. 2021, 126, 166–172. [Google Scholar] [CrossRef]
  110. Hosny, M.; Fawzy, M.; Abdelfatah, A.M.; Fawzy, E.E.; Eltaweil, A.S. Comparative study on the potentialities of two halophytic species in the green synthesis of gold nanoparticles and their anticancer, antioxidant and catalytic efficiencies. Adv. Powder Technol. 2021, 32, 3220–3233. [Google Scholar] [CrossRef]
  111. Aghamirzaei, M.; Khiabani, M.S.; Hamishehkar, H.; Mokarram, R.R.; Amjadi, M. Antioxidant, antimicrobial and cytotoxic activities of biosynthesized gold nanoparticles (aunps) from chinese lettuce (cl) leave extract (brassica rapa var. Pekinensis). Mater. Today Commun. 2021, 29, 102831. [Google Scholar] [CrossRef]
  112. Khuda, F.; Ul Haq, Z.; Ilahi, I.; Ullah, R.; Khan, A.; Fouad, H.; Ali Khan Khalil, A.; Ullah, Z.; Umar Khayam Sahibzada, M.; Shah, Y.; et al. Synthesis of gold nanoparticles using sambucus wightiana extract and investigation of its antimicrobial, anti-inflammatory, antioxidant and analgesic activities. Arab. J. Chem. 2021, 14, 103343. [Google Scholar] [CrossRef]
  113. Muniyappan, N.; Pandeeswaran, M.; Amalraj, A. Green synthesis of gold nanoparticles using curcuma pseudomontana isolated curcumin: Its characterization, antimicrobial, antioxidant and anti- inflammatory activities. Environ. Chem. Ecotoxicol. 2021, 3, 117–124. [Google Scholar] [CrossRef]
  114. Milanezi, F.G.; Meireles, L.M.; de Christo Scherer, M.M.; de Oliveira, J.P.; da Silva, A.R.; de Araujo, M.L.; Endringer, D.C.; Fronza, M.; Guimarães, M.C.C.; Scherer, R. Antioxidant, antimicrobial and cytotoxic activities of gold nanoparticles capped with quercetin. Saudi Pharm. J. 2019, 27, 968–974. [Google Scholar] [CrossRef]
  115. Veeraapandian, S.; Sawant, S.N.; Doble, M. Antibacterial and antioxidant activity of protein capped silver and gold nanoparticles synthesized with escherichia coli. J. Biomed. Nanotechnol. 2012, 8, 140–148. [Google Scholar] [CrossRef]
  116. Adewale, O.B.; Egbeyemi, K.A.; Onwuelu, J.O.; Potts-Johnson, S.S.; Anadozie, S.O.; Fadaka, A.O.; Osukoya, O.A.; Aluko, B.T.; Johnson, J.; Obafemi, T.O.; et al. Biological synthesis of gold and silver nanoparticles using leaf extracts of crassocephalum rubens and their comparative in vitro antioxidant activities. Heliyon 2020, 6, e05501. [Google Scholar] [CrossRef]
  117. Shanmugasundaram, T.; Radhakrishnan, M.; Gopikrishnan, V.; Pazhanimurugan, R.; Balagurunathan, R. A study of the bactericidal, anti-biofouling, cytotoxic and antioxidant properties of actinobacterially synthesised silver nanoparticles. Colloids Surf. B Biointerfaces 2013, 111, 680–687. [Google Scholar] [CrossRef]
  118. Prakash Patil, M.; Seong, Y.-A.; Kim, J.-O.; Bae Seo, Y.; Kim, G.-D. Synthesis of silver nanoparticles using aqueous extract of cuscuta japonica seeds and their antibacterial and antioxidant activities. Inorg. Chem. Commun. 2021, 134, 109035. [Google Scholar] [CrossRef]
  119. Raja, K.; Balamurugan, V.; Selvakumar, S.; Vasanth, K. Striga angustifolia mediated synthesis of silver nanoparticles: Anti-microbial, antioxidant and anti-proliferative activity in apoptotic p53 signalling pathway. J. Drug Deliv. Sci. Technol. 2021, 67, 102945. [Google Scholar] [CrossRef]
  120. Ragunathan, V.K.C. Sequential microwave-ultrasound-assisted silver nanoparticles synthesis: A swift approach, their antioxidant, antimicrobial, and in-silico studies. J. Mol. Liq. 2021, 347, 117954. [Google Scholar] [CrossRef]
  121. Chakraborty, B.; Kumar, R.S.; Almansour, A.I.; Kotresha, D.; Rudrappa, M.; Pallavi, S.S.; Hiremath, H.; Perumal, K.; Nayaka, S. Evaluation of antioxidant, antimicrobial and antiproliferative activity of silver nanoparticles derived from galphimia glauca leaf extract. J. King Saud Univ.—Sci. 2021, 33, 101660. [Google Scholar] [CrossRef]
  122. Kong, Y.; Paray, B.A.; Al-Sadoon, M.K.; Fahad Albeshr, M. Novel green synthesis, chemical characterization, toxicity, colorectal carcinoma, antioxidant, anti-diabetic, and anticholinergic properties of silver nanoparticles: A chemopharmacological study. Arab. J. Chem. 2021, 14, 103193. [Google Scholar] [CrossRef]
  123. Sun, X.; Zhang, H.; Wang, J.; Dong, M.; Jia, P.; Bu, T.; Wang, Q.; Wang, L. Sodium alginate-based nanocomposite films with strong antioxidant and antibacterial properties enhanced by polyphenol-rich kiwi peel extracts bio-reduced silver nanoparticles. Food Packag. Shelf Life 2021, 29, 100741. [Google Scholar] [CrossRef]
  124. Saygi, K.O.; Cacan, E. Antioxidant and cytotoxic activities of silver nanoparticles synthesized using tilia cordata flowers extract. Mater. Today Commun. 2021, 27, 102316. [Google Scholar] [CrossRef]
  125. Gauthami, R.; Vinitha, U.G.; Philip Anthony, S.; Sundaram Muthuraman, M. Cissampelous pairera mediated synthesis of silver nanoparticles and it’s invitro antioxidant, antibacterial and antidiabetic activities. Mater. Today Proc. 2021, 47, 853–857. [Google Scholar] [CrossRef]
  126. Khan, W.; Khan, N.; Jamila, N.; Masood, R.; Minhaz, A.; Amin, F.; Atlas, A.; Nishan, U. Antioxidant, antibacterial, and catalytic performance of biosynthesized silver nanoparticles of rhus javanica, rumex hastatus, and callistemon viminalis. Saudi J. Biol. Sci. 2021, 29, 894–904. [Google Scholar] [CrossRef]
  127. Badmus, J.A.; Oyemomi, S.A.; Adedosu, O.T.; Yekeen, T.A.; Azeez, M.A.; Adebayo, E.A.; Lateef, A.; Badeggi, U.M.; Botha, S.; Hussein, A.A.; et al. Photo-assisted bio-fabrication of silver nanoparticles using annona muricata leaf extract: Exploring the antioxidant, anti-diabetic, antimicrobial, and cytotoxic activities. Heliyon 2020, 6, e05413. [Google Scholar] [CrossRef]
  128. Ramya, S.; Shanmugasundaram, T.; Balagurunathan, R. Biomedical potential of actinobacterially synthesized selenium nanoparticles with special reference to anti-biofilm, anti-oxidant, wound healing, cytotoxic and anti-viral activities. J. Trace Elem. Med. Biol. 2015, 32, 30–39. [Google Scholar] [CrossRef]
  129. Torres, S.K.; Campos, V.L.; León, C.G.; Rodríguez-Llamazares, S.M.; Rojas, S.M.; González, M.; Smith, C.; Mondaca, M.A. Biosynthesis of selenium nanoparticles by pantoea agglomerans and their antioxidant activity. J. Nanoparticle Res. 2012, 14, 1236. [Google Scholar] [CrossRef]
  130. Abdel-Moneim, A.-M.E.; El-Saadony, M.T.; Shehata, A.M.; Saad, A.M.; Aldhumri, S.A.; Ouda, S.M.; Mesalam, N.M. Antioxidant and antimicrobial activities of spirulina platensis extracts and biogenic selenium nanoparticles against selected pathogenic bacteria and fungi. Saudi J. Biol. Sci. 2022, 29, 1197–1209. [Google Scholar] [CrossRef]
  131. Nagalingam, M.; Rajeshkumar, S.; Balu, S.K.; Tharani, M.; Arunachalam, K. Anticancer and antioxidant activity of morinda citrifolia leaf mediated selenium nanoparticles. J. Nanomater. 2022, 2022, 2155772. [Google Scholar] [CrossRef]
  132. Chen, W.; Cheng, H.; Xia, W. Construction of polygonatum sibiricum polysaccharide functionalized selenium nanoparticles for the enhancement of stability and antioxidant activity. Antioxidants 2022, 11, 240. [Google Scholar] [CrossRef] [PubMed]
  133. Sowani, H.; Mohite, P.; Munot, H.; Shouche, Y.; Bapat, T.; Kumar, A.R.; Kulkarni, M.; Zinjarde, S. Green synthesis of gold and silver nanoparticles by an actinomycete gordonia amicalis hs-11: Mechanistic aspects and biological application. Process Biochem. 2016, 51, 374–383. [Google Scholar] [CrossRef]
  134. Sivasankar, P.; Seedevi, P.; Poongodi, S.; Sivakumar, M.; Murugan, T.; Sivakumar, L.; Sivakumar, K.; Balasubramanian, T. Characterization, antimicrobial and antioxidant property of exopolysaccharide mediated silver nanoparticles synthesized by streptomyces violaceus mm72. Carbohydr. Polym. 2018, 181, 752–759. [Google Scholar] [CrossRef] [PubMed]
  135. Shreema, K.; Mathammal, R.; Kalaiselvi, V.; Vijayakumar, S.; Selvakumar, K.; Senthil, K. Green synthesis of silver doped zinc oxide nanoparticles using fresh leaf extract morinda citrifoliaand its antioxidant potential. Mater. Today Proc. 2021, 47, 2126–2131. [Google Scholar] [CrossRef]
  136. Moghaddam, A.B.; Moniri, M.; Azizi, S.; Rahim, R.A.; Ariff, A.B.; Saad, W.Z.; Namvar, F.; Navaderi, M.; Mohamad, R. Biosynthesis of zno nanoparticles by a new pichia kudriavzevii yeast strain and evaluation of their antimicrobial and antioxidant activities. Molecules 2017, 22, 872. [Google Scholar] [CrossRef] [Green Version]
  137. Netala, V.R.; Bethu, M.S.; Pushpalatha, B.; Baki, V.B.; Aishwarya, S.; Rao, J.V.; Tartte, V. Biogenesis of silver nanoparticles using endophytic fungus pestalotiopsis microspora and evaluation of their antioxidant and anticancer activities. Int. J. Nanomed. 2016, 11, 5683. [Google Scholar] [CrossRef] [Green Version]
  138. Joshi, C.G.; Danagoudar, A.; Poyya, J.; Kudva, A.K.; Dhananjaya, B.L. Biogenic synthesis of gold nanoparticles by marine endophytic fungus-cladosporium cladosporioides isolated from seaweed and evaluation of their antioxidant and antimicrobial properties. Process Biochem. 2017, 63, 137–144. [Google Scholar]
  139. Manjunath Hulikere, M.; Joshi, C.G. Characterization, antioxidant and antimicrobial activity of silver nanoparticles synthesized using marine endophytic fungus- cladosporium cladosporioides. Process Biochem. 2019, 82, 199–204. [Google Scholar] [CrossRef]
  140. Rajeshkumar, S.; Nandhini, N.T.; Manjunath, K.; Sivaperumal, P.; Krishna Prasad, G.; Alotaibi, S.S.; Roopan, S.M. Environment friendly synthesis copper oxide nanoparticles and its antioxidant, antibacterial activities using seaweed (sargassum longifolium) extract. J. Mol. Struct. 2021, 1242, 130724. [Google Scholar] [CrossRef]
  141. Ginting, B.; Maulana, I.; Karnila, I. Biosynthesis copper nanoparticles using blumea balsamifera leaf extracts: Characterization of its antioxidant and cytotoxicity activities. Surf. Interfaces 2020, 21, 100799. [Google Scholar] [CrossRef]
  142. Venugopalan, R.; Pitchai, S.; Devarayan, K.; Swaminathan, V.C. Biogenic synthesis of copper nanoparticles using borreria hispida (linn.) extract and its antioxidant activity. Mater. Today Proc. 2020, 33, 4023–4025. [Google Scholar] [CrossRef]
  143. Hemmati, S.; Ahmeda, A.; Salehabadi, Y.; Zangeneh, A.; Zangeneh, M.M. Synthesis, characterization, and evaluation of cytotoxicity, antioxidant, antifungal, antibacterial, and cutaneous wound healing effects of copper nanoparticles using the aqueous extract of strawberry fruit and l-ascorbic acid. Polyhedron 2020, 180, 114425. [Google Scholar] [CrossRef]
  144. Rajeshkumar, S.; Rinitha, G. Nanostructural characterization of antimicrobial and antioxidant copper nanoparticles synthesized using novel persea americana seeds. OpenNano 2018, 3, 18–27. [Google Scholar] [CrossRef]
  145. Zangeneh, M.M.; Ghaneialvar, H.; Akbaribazm, M.; Ghanimatdan, M.; Abbasi, N.; Goorani, S.; Pirabbasi, E.; Zangeneh, A. Novel synthesis of falcaria vulgaris leaf extract conjugated copper nanoparticles with potent cytotoxicity, antioxidant, antifungal, antibacterial, and cutaneous wound healing activities under in vitro and in vivo condition. J. Photochem. Photobiol. B Biol. 2019, 197, 111556. [Google Scholar] [CrossRef]
  146. Netala, V.R.; Kotakadi, V.S.; Bobbu, P.; Gaddam, S.A.; Tartte, V. Endophytic fungal isolate mediated biosynthesis of silver nanoparticles and their free radical scavenging activity and anti microbial studies. 3 Biotech 2016, 6, 132. [Google Scholar] [CrossRef] [Green Version]
  147. Du, L.; Suo, S.; Wang, G.; Jia, H.; Liu, K.J.; Zhao, B.; Liu, Y. Mechanism and cellular kinetic studies of the enhancement of antioxidant activity by using surface-functionalized gold nanoparticles. Chem.–Eur. J. 2013, 19, 1281–1287. [Google Scholar] [CrossRef]
  148. Nie, Z.; Liu, K.J.; Zhong, C.-J.; Wang, L.-F.; Yang, Y.; Tian, Q.; Liu, Y. Enhanced radical scavenging activity by antioxidant-functionalized gold nanoparticles: A novel inspiration for development of new artificial antioxidants. Free Radic. Biol. Med. 2007, 43, 1243–1254. [Google Scholar] [CrossRef]
  149. Ebabe Elle, R.; Rahmani, S.; Lauret, C.; Morena, M.; Bidel, L.P.R.; Boulahtouf, A.; Balaguer, P.; Cristol, J.-P.; Durand, J.-O.; Charnay, C. Functionalized mesoporous silica nanoparticle with antioxidants as a new carrier that generates lower oxidative stress impact on cells. Mol. Pharm. 2016, 13, 2647–2660. [Google Scholar] [CrossRef]
  150. Sotiriou, G.A.; Blattmann, C.O.; Deligiannakis, Y. Nanoantioxidant-driven plasmon enhanced proton-coupled electron transfer. Nanoscale 2016, 8, 796–803. [Google Scholar] [CrossRef]
  151. Deligiannakis, Y.; Sotiriou, G.A.; Pratsinis, S.E. Antioxidant and antiradical sio2 nanoparticles covalently functionalized with gallic acid. ACS Appl. Mater. Interfaces 2012, 4, 6609–6617. [Google Scholar] [CrossRef]
  152. Viglianisi, C.; Di Pilla, V.; Menichetti, S.; Rotello, V.M.; Candiani, G.; Malloggi, C.; Amorati, R. Linking an α-tocopherol derivative to cobalt (0) nanomagnets: Magnetically responsive antioxidants with superior radical trapping activity and reduced cytotoxicity. Chem.–Eur. J. 2014, 20, 6857–6860. [Google Scholar] [CrossRef] [Green Version]
  153. Massaro, M.; Amorati, R.; Cavallaro, G.; Guernelli, S.; Lazzara, G.; Milioto, S.; Noto, R.; Poma, P.; Riela, S. Direct chemical grafted curcumin on halloysite nanotubes as dual-responsive prodrug for pharmacological applications. Colloids Surf. B Biointerfaces 2016, 140, 505–513. [Google Scholar] [CrossRef]
  154. Massaro, M.; Riela, S.; Guernelli, S.; Parisi, F.; Lazzara, G.; Baschieri, A.; Valgimigli, L.; Amorati, R. A synergic nanoantioxidant based on covalently modified halloysite–trolox nanotubes with intra-lumen loaded quercetin. J. Mater. Chem. B 2016, 4, 2229–2241. [Google Scholar] [CrossRef] [Green Version]
  155. Kalashnikova, I.; Mazar, J.; Neal, C.J.; Rosado, A.L.; Das, S.; Westmoreland, T.J.; Seal, S. Nanoparticle delivery of curcumin induces cellular hypoxia and ros-mediated apoptosis via modulation of bcl-2/bax in human neuroblastoma. Nanoscale 2017, 9, 10375–10387. [Google Scholar] [CrossRef]
  156. Enes, R.F.; Tomé, A.C.; Cavaleiro, J.A.; Amorati, R.; Fumo, M.G.; Pedulli, G.F.; Valgimigli, L. Synthesis and antioxidant activity of [60]fullerene–bht conjugates. Chem.–Eur. J. 2006, 12, 4646–4653. [Google Scholar] [CrossRef]
  157. Czochara, R.; Kusio, J.; Litwinienko, G. Fullerene c 60 conjugated with phenols as new hybrid antioxidants to improve the oxidative stability of polymers at elevated temperatures. RSC Adv. 2017, 7, 44021–44025. [Google Scholar] [CrossRef] [Green Version]
  158. Gil, D.; Rodriguez, J.; Ward, B.; Vertegel, A.; Ivanov, V.; Reukov, V. Antioxidant activity of sod and catalase conjugated with nanocrystalline ceria. Bioengineering 2017, 4, 18. [Google Scholar] [CrossRef] [Green Version]
  159. Arriagada, F.; Correa, O.; Günther, G.; Nonell, S.; Mura, F.; Olea-Azar, C.; Morales, J. Morin flavonoid adsorbed on mesoporous silica, a novel antioxidant nanomaterial. PLoS ONE 2016, 11, e0164507. [Google Scholar] [CrossRef]
  160. Sahiner, N.; Sagbas, S.; Aktas, N. Preparation and characterization of monodisperse, mesoporous natural poly(tannic acid)–silica nanoparticle composites with antioxidant properties. Microporous Mesoporous Mater. 2016, 226, 316–324. [Google Scholar] [CrossRef]
  161. Liu, P.; Tang, H.; Lu, M.; Gao, C.; Wang, F.; Ding, Y.; Zhang, S.; Yang, M. Preparation of nanosilica-immobilized antioxidant and the antioxidative behavior in low density polyethylene. Polym. Degrad. Stab. 2017, 135, 1–7. [Google Scholar] [CrossRef]
  162. Kotcherlakota, R.; Barui, A.K.; Prashar, S.; Fajardo, M.; Briones, D.; Rodríguez-Diéguez, A.; Patra, C.R.; Gómez-Ruiz, S. Curcumin loaded mesoporous silica: An effective drug delivery system for cancer treatment. Biomater. Sci. 2016, 4, 448–459. [Google Scholar] [CrossRef] [PubMed]
  163. Marulasiddeshwara, M.; Dakshayani, S.; Kumar, M.S.; Chethana, R.; Kumar, P.R.; Devaraja, S. Facile-one pot-green synthesis, antibacterial, antifungal, antioxidant and antiplatelet activities of lignin capped silver nanoparticles: A promising therapeutic agent. Mater. Sci. Eng. C 2017, 81, 182–190. [Google Scholar] [CrossRef] [PubMed]
  164. Shah, S.T.; Yehye, W.A.; Saad, O.; Simarani, K.; Chowdhury, Z.Z.; Alhadi, A.A.; Al-Ani, L.A. Surface functionalization of iron oxide nanoparticles with gallic acid as potential antioxidant and antimicrobial agents. Nanomaterials 2017, 7, 306. [Google Scholar] [CrossRef] [PubMed]
  165. Santiago-Rodríguez, L.; Lafontaine, M.M.; Castro, C.; Méndez-Vega, J.; Latorre-Esteves, M.; Juan, E.J.; Mora, E.; Torres-Lugo, M.; Rinaldi, C. Synthesis, stability, cellular uptake, and blood circulation time of carboxymethyl-inulin coated magnetic nanoparticles. J. Mater. Chem. B 2013, 1, 2807–2817. [Google Scholar] [CrossRef] [Green Version]
  166. Bhattacharya, K.; Gogoi, B.; Buragohain, A.K.; Deb, P. Fe2o3/c nanocomposites having distinctive antioxidant activity and hemolysis prevention efficiency. Mater. Sci. Eng. C 2014, 42, 595–600. [Google Scholar] [CrossRef]
  167. Szekeres, M.; Illés, E.; Janko, C.; Farkas, K.; Tóth, I.Y.; Nesztor, D.; Zupkó, I.; Földesi, I.; Alexiou, C.; Tombácz, E. Hemocompatibility and biomedical potential of poly(gallic acid) coated iron oxide nanoparticles for theranostic use. J. Nanomed. Nanotechnol. 2015, 6. [Google Scholar]
  168. Song, W.; Muthana, M.; Mukherjee, J.; Falconer, R.J.; Biggs, C.A.; Zhao, X. Magnetic-silk core–shell nanoparticles as potential carriers for targeted delivery of curcumin into human breast cancer cells. ACS Biomater. Sci. Eng. 2017, 3, 1027–1038. [Google Scholar] [CrossRef]
  169. Liu, K.; Xiao, X.; Wang, J.; Chen, C.Y.O.; Hu, H. Polyphenolic composition and antioxidant, antiproliferative, and antimicrobial activities of mushroom inonotus sanghuang. LWT—Food Sci. Technol. 2017, 82, 154–161. [Google Scholar] [CrossRef] [Green Version]
  170. Joseph, A.; Wood, T.; Chen, C.-C.; Corry, K.; Snyder, J.M.; Juul, S.E.; Parikh, P.; Nance, E. Curcumin-loaded polymeric nanoparticles for neuroprotection in neonatal rats with hypoxic-ischemic encephalopathy. Nano Res. 2018, 11, 5670–5688. [Google Scholar] [CrossRef]
  171. Mittal, A.K.; Kumar, S.; Banerjee, U.C. Quercetin and gallic acid mediated synthesis of bimetallic (silver and selenium) nanoparticles and their antitumor and antimicrobial potential. J. Colloid Interface Sci. 2014, 431, 194–199. [Google Scholar] [CrossRef]
  172. Kumar, H.; Bhardwaj, K.; Kuča, K.; Kalia, A.; Nepovimova, E.; Verma, R.; Kumar, D. Flower-based green synthesis of metallic nanoparticles: Applications beyond fragrance. Nanomaterials 2020, 10, 766. [Google Scholar] [CrossRef] [Green Version]
  173. Lee, K.D.; Nagajyothi, P.C.; Sreekanth, T.V.M.; Park, S. Eco-friendly synthesis of gold nanoparticles (aunps) using inonotus obliquus and their antibacterial, antioxidant and cytotoxic activities. J. Ind. Eng. Chem. 2015, 26, 67–72. [Google Scholar] [CrossRef]
  174. Naveena, B.E.; Prakash, S. Biological synthesis of gold nanoparticles using marine algae gracilaria corticata and its application as a potent antimicrobial and antioxidant agent. Asian J. Pharm. Clin. Res. 2013, 6, 179–182. [Google Scholar]
  175. Sharma, B.; Purkayastha, D.D.; Hazra, S.; Thajamanbi, M.; Bhattacharjee, C.R.; Ghosh, N.N.; Rout, J. Biosynthesis of fluorescent gold nanoparticles using an edible freshwater red alga, lemanea fluviatilis (l.) c.Ag. And antioxidant activity of biomatrix loaded nanoparticles. Bioprocess Biosyst. Eng. 2014, 37, 2559–2565. [Google Scholar] [CrossRef]
  176. Bumbudsanpharoke, N.; Choi, J.; Park, I.; Ko, S. Facile biosynthesis and antioxidant property of nanogold-cellulose fiber composite. J. Nanomater. 2015, 16, 195. [Google Scholar] [CrossRef] [Green Version]
  177. Liberman, A.; Mendez, N.; Trogler, W.C.; Kummel, A.C. Synthesis and surface functionalization of silica nanoparticles for nanomedicine. Surf. Sci. Rep. 2014, 69, 132–158. [Google Scholar] [CrossRef] [Green Version]
  178. Sriramulu, M.; Sumathi, S. Photocatalytic, antioxidant, antibacterial and anti-inflammatory activity of silver nanoparticles synthesised using forest and edible mushroom. Adv. Nat. Sci. Nanosci. Nanotechnol. 2017, 8, 045012. [Google Scholar] [CrossRef] [Green Version]
  179. Aygün, A.; Özdemir, S.; Gülcan, M.; Cellat, K.; Şen, F. Synthesis and characterization of reishi mushroom-mediated green synthesis of silver nanoparticles for the biochemical applications. J. Pharm. Biomed. Anal. 2020, 178, 112970. [Google Scholar] [CrossRef]
  180. Poudel, M.; Pokharel, R.; Sudip, K.C.; Awal, S.C.; Pradhananga, R. Biosynthesis of silver nanoparticles using ganoderma lucidum and assessment of antioxidant and antibacterial activity. Int. J. Appl. Sci. Biotechnol. 2017, 5, 523–531. [Google Scholar] [CrossRef]
  181. Venkatesan, J.; Kim, S.-K.; Shim, M.S. Antimicrobial, antioxidant, and anticancer activities of biosynthesized silver nanoparticles using marine algae ecklonia cava. Nanomaterials 2016, 6, 235. [Google Scholar] [CrossRef]
  182. Dasari, S.; Suresh, K.A.; Rajesh, M.; Siva Reddy, C.S.; Hemalatha, C.S.; Wudayagiri, R.; Valluru, L. Biosynthesis, characterization, antibacterial and antioxidant activity of silver nanoparticles produced by lichens. J. Bionanosci. 2013, 7, 237–244. [Google Scholar] [CrossRef]
  183. Sriranjani, R.; Srinithya, B.; Vellingiri, V.; Brindha, P.; Anthony, S.P.; Sivasubramanian, A.; Muthuraman, M.S. Silver nanoparticle synthesis using clerodendrum phlomidis leaf extract and preliminary investigation of its antioxidant and anticancer activities. J. Mol. Liq. 2016, 220, 926–930. [Google Scholar] [CrossRef]
  184. Das, D.; Ghosh, R.; Mandal, P. Biogenic synthesis of silver nanoparticles using s1 genotype of morus alba leaf extract: Characterization, antimicrobial and antioxidant potential assessment. SN Appl. Sci. 2019, 1, 498. [Google Scholar] [CrossRef] [Green Version]
  185. Vilas, V.; Philip, D.; Mathew, J. Essential oil mediated synthesis of silver nanocrystals for environmental, anti-microbial and antioxidant applications. Mater. Sci. Eng. C 2016, 61, 429–436. [Google Scholar] [CrossRef]
  186. Srinithya, B.; Kumar, V.V.; Vadivel, V.; Pemaiah, B.; Anthony, S.P.; Muthuraman, M.S. Synthesis of biofunctionalized agnps using medicinally important sida cordifolia leaf extract for enhanced antioxidant and anticancer activities. Mater. Lett. 2016, 170, 101–104. [Google Scholar] [CrossRef]
  187. Kalaiyarasan, T.; Bharti, V.K.; Chaurasia, O. One pot green preparation of seabuckthorn silver nanoparticles (sbt@ agnps) featuring high stability and longevity, antibacterial, antioxidant potential: A nano disinfectant future perspective. RSC Adv. 2017, 7, 51130–51141. [Google Scholar] [CrossRef] [Green Version]
  188. Teerasong, S.; Jinnarak, A.; Chaneam, S.; Wilairat, P.; Nacapricha, D. Poly (vinyl alcohol) capped silver nanoparticles for antioxidant assay based on seed-mediated nanoparticle growth. Talanta 2017, 170, 193–198. [Google Scholar] [CrossRef]
  189. Guajardo-Pacheco, M.J.; Morales-Sánchez, J.E.; González-Hernández, J.; Ruiz, F. Synthesis of copper nanoparticles using soybeans as a chelant agent. Mater. Lett. 2010, 64, 1361–1364. [Google Scholar] [CrossRef]
  190. Xi, Y.; Hu, C.; Gao, P.; Yang, R.; He, X.; Wang, X.; Wan, B. Morphology and phase selective synthesis of cuxo (x=1, 2) nanostructures and their catalytic degradation activity. Mater. Sci. Eng. B 2010, 166, 113–117. [Google Scholar] [CrossRef]
  191. He, Y. A novel solid-stabilized emulsion approach to cuo nanostructured microspheres. Mater. Res. Bull. 2007, 42, 190–195. [Google Scholar] [CrossRef]
  192. Motoyoshi, R.; Oku, T.; Suzuki, A.; Kikuchi, K.; Kikuchi, S.; Jeyadevan, B.; Cuya, J. Fabrication and characterization of cuprous oxide: Fullerene solar cells. Synth. Met. 2010, 160, 1219–1222. [Google Scholar] [CrossRef]
  193. Rajeshkumar, S.; Menon, S.; Venkat Kumar, S.; Tambuwala, M.M.; Bakshi, H.A.; Mehta, M.; Satija, S.; Gupta, G.; Chellappan, D.K.; Thangavelu, L.; et al. Antibacterial and antioxidant potential of biosynthesized copper nanoparticles mediated through cissus arnotiana plant extract. J. Photochem. Photobiol. B Biol. 2019, 197, 111531. [Google Scholar] [CrossRef] [PubMed]
  194. Herlekar, M.; Barve, S.; Kumar, R. Plant-mediated green synthesis of iron nanoparticles. J. Nanoparticles 2014, 2014, 140614. [Google Scholar] [CrossRef] [Green Version]
  195. Huber, D.L. Synthesis, properties, and applications of iron nanoparticles. Small 2005, 1, 482–501. [Google Scholar] [CrossRef]
  196. Guo, J.; Wang, R.; Tjiu, W.W.; Pan, J.; Liu, T. Synthesis of fe nanoparticles@graphene composites for environmental applications. J. Hazard. Mater. 2012, 225–226, 63–73. [Google Scholar] [CrossRef]
  197. Babay, S.; Mhiri, T.; Toumi, M. Synthesis, structural and spectroscopic characterizations of maghemite γ-fe2o3 prepared by one-step coprecipitation route. J. Mol. Struct. 2015, 1085, 286–293. [Google Scholar] [CrossRef]
  198. Saleh, N.; Kim, H.-J.; Phenrat, T.; Matyjaszewski, K.; Tilton, R.D.; Lowry, G.V. Ionic strength and composition affect the mobility of surface-modified fe0 nanoparticles in water-saturated sand columns. Environ. Sci. Technol. 2008, 42, 3349–3355. [Google Scholar] [CrossRef]
  199. Kim, H.J.; Kim, D.-G.; Yoon, H.; Choi, Y.-S.; Yoon, J.; Lee, J.-C. Polyphenol/feiii complex coated membranes having multifunctional properties prepared by a one-step fast assembly. Adv. Mater. Interfaces 2015, 2, 1500298. [Google Scholar] [CrossRef]
  200. Yang, L.; Cao, Z.; Sajja, H.K.; Mao, H.; Wang, L.; Geng, H.; Xu, H.; Jiang, T.; Wood, W.C.; Nie, S.; et al. Development of receptor targeted magnetic iron oxide nanoparticles for efficient drug delivery and tumor imaging. J. Biomed. Nanotechnol. 2008, 4, 439–449. [Google Scholar] [CrossRef] [Green Version]
  201. Ebrahiminezhad, A.; Zare-Hoseinabadi, A.; Sarmah, A.K.; Taghizadeh, S.; Ghasemi, Y.; Berenjian, A. Plant-mediated synthesis and applications of iron nanoparticles. Mol. Biotechnol. 2018, 60, 154–168. [Google Scholar] [CrossRef]
  202. Alves, A.d.C.S.; Mainardes, R.M.; Khalil, N.M. Nanoencapsulation of gallic acid and evaluation of its cytotoxicity and antioxidant activity. Mater. Sci. Eng. C 2016, 60, 126–134. [Google Scholar] [CrossRef]
  203. Dorniani, D.; Hussein, M.Z.B.; Kura, A.U.; Fakurazi, S.; Shaari, A.H.; Ahmad, Z. Preparation of fe3o4 magnetic nanoparticles coated with gallic acid for drug delivery. Int. J. Nanomed. 2012, 7, 5745. [Google Scholar] [CrossRef] [Green Version]
  204. Shah, S.T.; Chowdhury, Z.Z.; Johan, M.R.B.; Badruddin, I.A.; Khaleed, H.M.T.; Kamangar, S.; Alrobei, H. Surface functionalization of magnetite nanoparticles with multipotent antioxidant as potential magnetic nanoantioxidants and antimicrobial agents. Molecules 2022, 27, 789. [Google Scholar] [CrossRef]
  205. Shah, S.T.; Chowdhury, Z.Z.; Johan, M.R.; Badruddin, I.A.; Alrobei, H.; Kamangar, S. Design and synthesis of multipotent antioxidants for functionalization of iron oxide nanoparticles. Coatings 2022, 12, 517. [Google Scholar] [CrossRef]
  206. Li, Y.; Li, X.; Zheng, W.; Fan, C.; Zhang, Y.; Chen, T. Functionalized selenium nanoparticles with nephroprotective activity, the important roles of ros-mediated signaling pathways. J. Mater. Chem. B 2013, 1, 6365–6372. [Google Scholar] [CrossRef]
  207. Abdullah; Hussain, T.; Faisal, S.; Rizwan, M.; Saira; Zaman, N.; Iqbal, M.; Iqbal, A.; Ali, Z. Green synthesis and characterization of copper and nickel hybrid nanomaterials: Investigation of their biological and photocatalytic potential for the removal of organic crystal violet dye. J. Saudi Chem. Soc. 2022, 26, 101486. [Google Scholar] [CrossRef]
Figure 1. Oxidative stress Adverse Impacts on Human Health.
Figure 1. Oxidative stress Adverse Impacts on Human Health.
Coatings 12 01568 g001
Figure 2. Nanoparticles with Antioxidant Activity: “Nanoantioxidants”.
Figure 2. Nanoparticles with Antioxidant Activity: “Nanoantioxidants”.
Coatings 12 01568 g002
Figure 3. Characteristics of a chain-breaking antioxidant.
Figure 3. Characteristics of a chain-breaking antioxidant.
Coatings 12 01568 g003
Figure 4. System for Classifying Nanoantioxidants Based on Structural and Mode of Action.
Figure 4. System for Classifying Nanoantioxidants Based on Structural and Mode of Action.
Coatings 12 01568 g004
Figure 5. Comparisons and contrasts between sacrificial (A), catalytic, and (B) SOD-like nanoantioxidants (C). Nanoparticles as the electron donor or accepter.
Figure 5. Comparisons and contrasts between sacrificial (A), catalytic, and (B) SOD-like nanoantioxidants (C). Nanoparticles as the electron donor or accepter.
Coatings 12 01568 g005
Figure 6. Trolox grafting procedure on halloysite nanotubes; the incorporation of the antioxidant quercetin in the nanotubes’ interior.
Figure 6. Trolox grafting procedure on halloysite nanotubes; the incorporation of the antioxidant quercetin in the nanotubes’ interior.
Coatings 12 01568 g006
Figure 7. AuNPs (Au@- PEG3SA) Functionalized with Salvianic Acid A.
Figure 7. AuNPs (Au@- PEG3SA) Functionalized with Salvianic Acid A.
Coatings 12 01568 g007
Figure 8. The molecular architecture of (a) alpha-tocopherol, (b) trolox, (c) trolox-functionalized AuNPs, and (d) reactions of Au@Trolox with DPPH• radicals.
Figure 8. The molecular architecture of (a) alpha-tocopherol, (b) trolox, (c) trolox-functionalized AuNPs, and (d) reactions of Au@Trolox with DPPH• radicals.
Coatings 12 01568 g008
Figure 9. (A) Through the H atom transfer (HAT) process from the GA (gallic acid) molecule, one DPPH radical is scavenged by two SiO2-GA nanoparticles, resulting in the formation of a transitory GA radical. (B) HAT from the GA semiquinone results in the formation of a nonradical GA quinone, which SiO2-GA nanoparticles scavenge.
Figure 9. (A) Through the H atom transfer (HAT) process from the GA (gallic acid) molecule, one DPPH radical is scavenged by two SiO2-GA nanoparticles, resulting in the formation of a transitory GA radical. (B) HAT from the GA semiquinone results in the formation of a nonradical GA quinone, which SiO2-GA nanoparticles scavenge.
Coatings 12 01568 g009
Figure 10. ROS generation, Nrf2 activation, and cell death are all affected by free catechol antioxidant (Rutin), naked MSNs, and MSNs-RUT.
Figure 10. ROS generation, Nrf2 activation, and cell death are all affected by free catechol antioxidant (Rutin), naked MSNs, and MSNs-RUT.
Coatings 12 01568 g010
Figure 11. Surface Functionalization of IONP with Gallic Acid.
Figure 11. Surface Functionalization of IONP with Gallic Acid.
Coatings 12 01568 g011
Figure 12. (A) Vitamin C-coated nanoparticles (Si@AuNP or PAPM) and (B) cellular oxidative Stress at concentrations of micro- and millimolar vitamin C.
Figure 12. (A) Vitamin C-coated nanoparticles (Si@AuNP or PAPM) and (B) cellular oxidative Stress at concentrations of micro- and millimolar vitamin C.
Coatings 12 01568 g012
Table 1. Mechanisms of Action and Chemical Assays for Evaluating the Activity of Antioxidant Nanoparticles.
Table 1. Mechanisms of Action and Chemical Assays for Evaluating the Activity of Antioxidant Nanoparticles.
NanoantioxidantAssayRef.
Catalase mimic (pH > 7)
PVDF coated Au NPsSpectrophotometric H2O2 decrease Oxygen evolution[75]
-(EPR)-
Cerium Oxide NanoparticlesSpectrophotometric H2O2 disappearance Oxygen evolution[62]
Cobalt Oxide nanoparticlesClark electrod Oxygen evolution from H2O2[61]
Gold Plantium Nanorods; Core shellH2O2 decrease (spectrophotometric), O2 evolution[73]
Gold Nanorods; Core shell(Clark electrode)-
Gold Palladium Nanorods--
Apoferritin-coated Pt NPsH2O2 consumption, evolution of O2 bubbles[76]
Platinium NanopowderO2 detection by EPR line broadening[77]
Trimanganese tetraoxide nanoflowersSpectrophotometric H2O2 decrease[63]
Dimercaptosuccinic acid coated Fe3O4 NPsO2 formation (Clark electrode)[64]
Eumelanin-silica NPsFerrous Xylenol orange assay decrease in H2O2 concentration.[65]
GPX mimic
Vanadium NanowiresChange in coenzyme NADPH absorbance at 340 nm in a glutathione reductase coupled assay (spectrophotometric)[74]
GO supported Selenium nanoparticlesSpectrophotometric Glutathione reductase coupled assay[78]
Trimanganese tetraoxide nanoflowersSpectrophotometric Glutathione reductase coupled assay[63]
Chain-breaking
Polyacrylic acid (PAA)- protected Pt NPsDPPH (spectrophotometric), linoleic acid peroxidation, and AAPH-derived radical scavenging are inhibited (EPR detection)[79]
Oleic acid coated cerium oxide NPsRadical scavenging from AAPH (ORAC assay)[80]
ZrO2 NPsDPPH (spectrophotometric)[81]
PEG coated melanin NPsDPPH (spectrophotometric and EPR)[82]
SOD mimic
PEG-coated melanin NPs5-Diethoxyphosphoryl-5-Methyl-1-Pyrroline N Oxide Reaction with O2•− and EPR Study of O2 Evolution (Clark electrode)[83]
PVDF coated Gold nanoparticlesspin-trap with EPR detection and xanthine/xanthine oxidase[75]
Glycine-coated Cu(OH)2 NPsIodonitrotetrazolium chloride and Xanthine/Xanthine Oxidase (spectrophotometric)[84]
PEG coated MnO NPscytochrome C and xanthine/xanthine oxidase (spectrophotometric)[85]
PEG-coated carbon nanoclustersEPR analysis of the O2 reaction[86]
Pd nanocrystalsspin-trap with EPR detection and xanthine/xanthine oxidase[87]
Platinium NanopwderXanthine/Xanthine Oxidase with an EPR-detection spin-trap[77]
Trimanganese tetraoxide nanoflowersIodonitrotetrazolium chloride and Xanthine/Xanthine Oxidase (spectrophotometric)[63]
Dimercapto succinic acid coated Co3O4 NPsspin-trap with EPR detection and xanthine/xanthine oxidase[66]
MWCNTscytochrome C and xanthine/xanthine oxidase (spectrophotometric)[88]
Functionalized fullereneReaction with 1O2 (spectrophotometric)[88]
Table 2. Nanoparticles with Antioxidant Functionalization.
Table 2. Nanoparticles with Antioxidant Functionalization.
Types of NanoparticlesBiological ExtractMorphologySizeAntioxidant ActivityRef.
IronAmaranthus dubiusleaf extractSpherical43–220 nmDPPH[92]
IronAmaranthus spinosus leaf extractSpherical-DPPH[93]
IronAsphodelus aestivus Brot. ExtractSpherical20–25 nmDPPH, ABTS[94]
Iron oxidetea-pruning wasteSpherical20–35 nmDPPH[95]
Iron oxidePhoenix dactylifera L.Spherical2–30 nmDPPH[96]
Iron oxidecurcuminSpherical-DPPH[97]
Iron oxideCoriandrum sativum L. (cilantro)Spherical-DPPH[98]
Iron oxideBlumea eriantha DCspherical and irregular shapes10–60DPPH, ABTS, Hydrogen peroxide scavenging activity and Reducing power assay activity[99]
Nickel OxideStevia rebaudiana Bertoni leaf extractSpherical20–50 nmDPPH[100]
GoldLactobacillus kimchicus DCY51T biomassSpherical13 nmDPPH[101]
SilverStreptomyces griseorubens AU2 cell free supernatantSpherical5–20 nmDPPH[102]
GoldEnterococcus species cell free extractSphere8–50 nmDPPH[103]
GoldSnail slimeSpherical, hexagonal, trapezoidal, and rod shape14 ± 6 nmDPPH, ABTS[104]
GoldAchillea biebersteiniiSpherical2–30 nmDPPH[105]
GoldCentaurea behenSpherical50 nmDPPH[106]
Types of NanoparticlesBiological ExtractMorphologySizeAntioxidant ActivityRef.
GoldCannabis sativa L.Spherical-DPPH[107]
GoldCurcumae Kwangsiensis foliumSpherical8–25 nmDPPH[108]
GoldKaempferia parviflora rhizomeSpherical20–60 nmDPPH[109]
GoldAtriplex halimus and Chenopodium amperosidiesSpherical2–10 nmDPPH[110]
GoldBrassica rapa var. pekinensisSpherical25 nmDPPH, ABTS[111]
GoldSambucus wightianatrigonal, cubic, hexagonal, and polygonal15.96 nmDPPH[112]
GoldCurcuma pseudomontanaSpherical20 nmDPPH, RP, H2O2, NO and Cupric ion reducing antioxidant capacity (CUPRAC) assay[113]
GoldQuercetinSpherical100 nmABTS, DPPH and NO assay[114]
Gold and SilverEscherichia coli cell proteinTriangular, circular, hexagonal (AuNPs), Sphere (AgNPs)10–100 nm (AuNPs), 10–50 nm (AgNPs)EC75[115]
Gold and SilverCrassocephalum rubensspherical15–25 nm for AgNPs, 10–20 nm for AuNPslipid peroxidation and DPPH[116]
SilverStreptomyces griseorubens AU2 cell free supernatant biomassSphere5–50 nmDPPH[117]
SilverCuscuta japonicaoval-spherical30–50 nmDPPH[118]
SilverStriga angustifoliananoflakes106.40 nmDPPH[119]
SilverOlea europaeaSpherical10.47 ± 9.19 nmDPPH[120]
SilverGalphimia glaucaSpherical19–37 nmDPPH[121]
SilverCannabis sativaSpherical10–50 nmDPPH[122]
Silverpolyphenol-rich kiwi peelSpherical200–300 nmABTS[123]
SilverTilia cordataSpherical50 nmFRAP[124]
SilverCissampelous paireraSpherical60 nm to 118 nmDPPH[125]
SilverRhus javanica, Rumex hastatus, and Callistemon viminalis-1–100 nmDPPH, ABTS[126]
SilverAnnona muricataSpherical35 nmDPPH, ABTS[127]
SeleniumStreptomycesminutiscleroticus RT M10A62 biomassSpherical10–250 nmDPPH[128]
SeleniumPantoea agglomerans UC-32Spherical100 nmHigh antioxidant activity in human umbilical[129]
SeleniumSpirulina platensisSpherical32–86 nmDPPH, ABTS[130]
SeleniumMorinda citrifoliaSpherical12–160 nmDPPH[131]
-Polygonatum sibiricum polysaccharideSpherical105 nmDPPH, ABTS[132]
Gold and SilverGordonia amicalis HS-11 cell free supernatantGrain5–25 nmCFS synthesized AgNPs and AuNPs[133]
SilverStreptomyces violaceus MM72 exopolysaccharides-30 nmDPPH[134]
Silver doped zinc oxideMorinda citrifoliaSpherical-DPPH[135]
Zinc OxidePichia kudriavzevii cell free extractHexagonal10–61 nmDPPH[136]
SilverPestalotiopsis microspora filtrateSphere2–10 nm-[137]
GoldCladosporium cladosporioides filtrateCubic100 nmDPPH[138]
SilverCladosporiumSphere100 nm−DPPH[139]
CopperCissus arnotianaSphere60–90 nmDPPH-
Copper oxideSargassum longifoliumSphere40 and 60 nmDPPH[140]
CopperBlumea balsamiferaSphere30–55 nmDPPH[141]
CopperBorreria hispidaSphere121 ± 37 nmDPPH[142]
CopperFragaria ananassaSphere10–30 nmDPPH[143]
CopperPersea americanaSphere42–90 nmDPPH[144]
CopperFalcaria vulgarisSphere20–25 nmDPPH[145]
Copper OxideCissus arnotiana leaf extractSpherical80–90 nmDPPH[146]
Table 3. Antioxidant nanoparticle efficacy testing methods and mechanisms.
Table 3. Antioxidant nanoparticle efficacy testing methods and mechanisms.
NanomaterialAntioxidantMethod to Determine ActivityRef.
Halloysite nanotubesCurcuminPressure sensor method (inhibited autoxidation oxygen consumption), spectrophotometric method (DPPH assay)[153]
Halloysite nanotubesTroloxPressure sensor method (inhibited autoxidation oxygen consumption), spectrophotometric method (DPPH assay)[154]
Graphite coated cobalt nanoparticlesTroloxPressure sensor method (inhibited autoxidation oxygen consumption)[152]
SilicaRutin or Caffeic acidORAC[149]
Gold NanoparticlesTroloxDPPH• (EPR detection)[148]
Gold NanoparticlesSalvianic acidDPPH• (EPR detection)[147]
Silica coated Silver nanoparticlesGallic acidDPPH• (spectrophotometric and EPR)[150]
Silica nanoparticlesGallic acidDPPH• (spectrophotometric and EPR)[151]
Single walled CNTsBHTORAC[45]
FullereneBHTInhibited autoxidation, O2 consumption (pressure sensor)[156]
FullereneFlavonoidsInhibited autoxidation, O2 consumption (pressure sensor)[156]
double hydroxide NPsSODXanthine and xanthine oxidase, nitroblue tetrazolium (spectrophotometric)[157]
Cerium nanoparticlesSODXanthine and xanthine oxidase, hydrosoluble tetrazolium salt (spectrophotometric)[158]
Cerium nanoparticlesCATAmplex red assay[158]
MSNSurface functionalization; morin(2’,3,4’,5,7-pentahydroxyflavone)An effective scavenger of HO• and quencher for 1O2[159]
MSNPoly Tannic acid; crossing linkingTrolox equivalency antioxidant capacity (TEAC)[160]
MSNCaffeic acid and Rutin; covalent graftingFunctions of antiradicals, cell toxicity alleviation, and oxidative stress prevention[149]
MSN3,5-di-tert-butyl-4-hydroxybenzoic acid; graftingOxidation induction time (OIT) determined by DSC[161]
MSNCurcumin; loadedQuantification of cellular ROS[162]
Polyethylene glycol coated AuNPsSalvianic acid; Surface
functionalization
Thiobarbituric acid reactive substance (TBARS)
DPPH(spectrophotometric)
[147]
AuNPsTrolox; Self-assemblyDPPH• radical scavenging assay (spectrophotometric)[148]
AuNPs3,6-dihydroxyflavone, lutein and
selenium methyl selenocysteine;
embedded
Radical scavenging tests using hydroxyl, hydrogen peroxide, nitric oxide, and DPPH (2,2-diphenyl-l-picrylhydrazyl)[23]
AgNPsCoated with LigninHuman pathogens S. aureus, E. coli, and A. niger are combated, antioxidant assay DPPH (spectrophotometric) and antifungal and antibacterial agent.[163]
Fe3O4NPsFunctionalized with Gallic acidSuperior antibacterial and antifungal activity; antioxidant properties and are magnetically separable[164]
Fe2O3NPsCoated with Carboxymethyl inulinNon-toxic to cancer cell lines that have been immortalised.[165]
Fe2O3NPsCoated with carbonPotential antioxidant; compatible with peripheral blood mononuclear cells[166]
Fe2O3NPsCoated with Poly GAReduces oxidative stress significantly; biocompatible and bioactive[167]
Magnetic-silk
core-shell
nanoparticle
Loaded with CurcuminCellular uptake and cytotoxicity in breast cancer cell lines from human breast cancer patients[168]
Ceria nanoparticlesdextran-coated and loaded with curcumininherently anti-cancer[155]
Ceria nanoparticlesPEG-Phospholipid coatedA less toxic, biocompatible, oxidative stress-reducing, and effective treatment for patients with intracerebral hemorrhage.[169]
PLGA-PEGCoated with CurcuminNeonatal hypoxic-ischemic encephalopathy can be protected by this therapy[170]
Se-Ag BimetallicGallic acid and QuercetinAntimicrobial and antioxidant activities[171]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shah, S.T.; Chowdhury, Z.Z.; Simarani, K.; Basirun, W.J.; Badruddin, I.A.; Hussien, M.; Alrobei, H.; Kamangar, S. Nanoantioxidants: The Fourth Generation of Antioxidants—Recent Research Roadmap and Future Perspectives. Coatings 2022, 12, 1568. https://doi.org/10.3390/coatings12101568

AMA Style

Shah ST, Chowdhury ZZ, Simarani K, Basirun WJ, Badruddin IA, Hussien M, Alrobei H, Kamangar S. Nanoantioxidants: The Fourth Generation of Antioxidants—Recent Research Roadmap and Future Perspectives. Coatings. 2022; 12(10):1568. https://doi.org/10.3390/coatings12101568

Chicago/Turabian Style

Shah, Syed Tawab, Zaira Zaman Chowdhury, Khanom Simarani, Wan Jefrey Basirun, Irfan Anjum Badruddin, Mohamed Hussien, Hussein Alrobei, and Sarfaraz Kamangar. 2022. "Nanoantioxidants: The Fourth Generation of Antioxidants—Recent Research Roadmap and Future Perspectives" Coatings 12, no. 10: 1568. https://doi.org/10.3390/coatings12101568

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