Next Article in Journal
Implementing Defects for Ratiometric Luminescence Thermometry
Next Article in Special Issue
Advances in Gold Nanoparticle-Based Combined Cancer Therapy
Previous Article in Journal
Effects of 5 MeV Proton Irradiation on Nitrided SiO2/4H-SiC MOS Capacitors and the Related Mechanisms
Previous Article in Special Issue
Nanomedicine and Onco-Immunotherapy: From the Bench to Bedside to Biomarkers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Antioxidant Functionalized Nanoparticles: A Combat against Oxidative Stress

1
School of Bioengineering & Food Technology, Shoolini University of Biotechnology and Management Sciences, Solan 173229, Himachal Pradesh, India
2
School of Biological and Environmental Sciences, Shoolini University of Biotechnology and Management Sciences, Solan 173229, Himachal Pradesh, India
3
Department of Chemistry, Faculty of Science, University of Hradec Kralove, 50003 Hradec Kralove, Czech Republic
4
School of Biotechnology and Biosciences, Lovely Professional University, Phagwara 144411, Punjab, India
5
Biotransformation and Biomaterials Laboratory, Department of Microbial Engineering, Konkuk University, Seoul 05029, Korea
*
Authors to whom correspondence should be addressed.
Nanomaterials 2020, 10(7), 1334; https://doi.org/10.3390/nano10071334
Submission received: 7 June 2020 / Revised: 4 July 2020 / Accepted: 6 July 2020 / Published: 8 July 2020
(This article belongs to the Special Issue Advances in Nanomaterials in Biomedicine)

Abstract

:
Numerous abiotic stresses trigger the overproduction of reactive oxygen species (ROS) that are highly toxic and reactive. These ROS are known to cause damage to carbohydrates, DNA, lipids and proteins, and build the oxidative stress and results in the induction of various diseases. To resolve this issue, antioxidants molecules have gained significant attention to scavenge these free radicals and ROS. However, poor absorption ability, difficulty in crossing the cell membranes and degradation of these antioxidants during delivery are the few challenges associated with both natural and synthetic antioxidants that limit their bioavailability. Moreover, the use of nanoparticles as an antioxidant is overlooked, and is limited to a few nanomaterials. To address these issues, antioxidant functionalized nanoparticles derived from various biological origin have emerged as an important alternative, because of properties like biocompatibility, high stability and targeted delivery. Algae, bacteria, fungi, lichens and plants are known as the producers of diverse secondary metabolites and phenolic compounds with extraordinary antioxidant properties. Hence, these compounds could be used in amalgamation with biogenic derived nanoparticles (NPs) for better antioxidant potential. This review intends to increase our knowledge about the antioxidant functionalized nanoparticles and the mechanism by which antioxidants empower nanoparticles to combat oxidative stress.

1. Introduction

In the twenty-first century, age-related diseases have become a major health concern worldwide. Ageing is a natural and progressive process which involves the degeneration of the functioning and structure of vital organs and is one of the risk factors responsible for numerous chronic diseases and accounts for the high mortality rate [1,2,3,4]. Among various theories thatunveil and elucidate the ageing process, the free radical theory holds an exceptional rank [5]. This theory states that the ageing occurs due to successive failure of the defense mechanism to resort the damage induced by the reactive oxygen species (ROS), especially in the mitochondria [6].
It is well comprehended that oxidative stress plays a significant role in degenerative senescence. ROS have been found to be involved in the pathogenesis of various cellular processes, and is also associated with numerous diseases like cardiovascular, cancer, neurodegenerative and respiratory diseases, as depicted in Figure 1 [7]. The rise in ROS concentration in cells have also been associated with ageing, however, it cannot be considered as the only determining factor responsible for ageing. Moreover, in age-related diseases, the elevated concentration of ROS has been involved in the impairment of mitochondria and cellular oxidative damage [2,8].
The production of ROS generally relies on both enzymatic as well as non-enzymatic reactions. The enzymatic reactions involved in various cellular processes, like phagocytosis, prostaglandin synthesis and respiratory chain system, are known to generate ROS [9,10,11,12,13,14,15,16,17,18,19]. The superoxide radical (O2•−) is synthesized during the activity of enzymes, like NADPH oxidase, peroxidase and xanthine oxidase, in various cellular processes. It has also been found that various other ROS, like hydrogen peroxide (H2O2), hydroxyl radical (OH), hypochlorous acid (HOCl), peroxynitrite (ONOO), etc., are also formed during enzymatic reaction, and the action of enzymes like xanthine oxidase and amino acid oxidase leads to the formation of H2O2. Furthermore, OH is regarded as a highly reactive free radical species formed during the “Fenton reaction” between H2O2 and O2•−, in the presence Cu+ or Fe2+, which acts as the catalyst [11,12,13,14,15,16,17,18]. On the other hand, the non-enzymatic reactions between organic compounds and oxygen, or when cells are exposed to ionizing radiations and during mitochondrial respiration, have also been found to be involved in ROS formation [14,15,18]. At present, extensive research is being conducted to explore the natural compounds that can control oxidative stress and improve the immune system [20]. The search for novel molecules with antioxidant properties is an effective way to promote healthy ageing and counteract oxidative stress. Hence, this review focuses on highlighting the effectiveness of antioxidants functionalized nanoparticles. The first section of the review discusses synergism between ROS and age-related diseases, antioxidants and sources of antioxidants. The section following that discusses the role of nano-antioxidants; antioxidant functionalized nanoparticles and challenges associate with them.

2. Synergism between ROS and Age-Related Diseases

The overproductions of ROS have been found to be associated with numerous chronic diseases like cancer, cardiovascular, neurodegenerative and respiratory ailments. The synergism between ROS and chronic diseases is discussed in the following sections.

2.1. Cancer

Cancer, a fatal disease involves the malignant growth of tumors because of chromosomal alteration and lead to unregulated growth of the cells [21]. This deadly disease has a complex relationship with ROS and is involved at three different levels of cancer development, i.e., initiation, progression and promotion [22]. During the initiation stage, the ROS causes a mutation in DNA, which keeps on accumulating when the affected tissue does not get repaired [23]. The overproduction of ROS triggers the mutation in an oncogene, which potentially contributes to the initiation of cancer [24].
The cancer cells favor the excessive production of ROS in comparison to healthy cells, because of the alteration in the metabolic processes [25]. ROS-induced oxidative stresses in tumor triggers the cell signaling pathways and build resistance in tumor cells and elevate the supply of blood to tumor cells and promote their metastasis [26]. The elevated level of ROS plays a significant role in expanding tumor cells by altering the genes associated with apoptosis, cell proliferation and transcription factors [27]. Furthermore, ROS also downregulate the pro-apoptotic proteins by interfering with the Akt/PI3K and ERK cell signaling pathway, and upregulate the anti-apoptotic genes [28,29]. During the cancer progression stage, ROS interferes with cellular processes and upregulates the production of metalloproteinases by obstructing the angiogenesis process and, by anti-proteases, results in the metastasis of cancer cells [23,25,30].

2.2. Cardiovascular Disease

Cardiovascular disease, another fatal ailment, has a strong association with ROS during the development stage [31]. The overproduction of ROS in vascular cells during the reactions involving enzymes like NADPH oxidase, nitric oxide synthases causes the modification in low-density lipoproteins (LDL) [32]. Furthermore, ROS are also found to be involved in cardiac hypertrophy development, myocyte apoptosis and ischemia-reperfusion injury, which ultimately lead to cardiac arrest [33,34,35].

2.3. Neurodegenerative Diseases

Neurons are fundamental units of the brain and play a significant role in coordinating the actions and reactions of bodily functions. These neurons are highly vulnerable to ROS, as they weaken the antioxidants defense system, elevate the fatty acid (polyunsaturated) content in the cell membrane and increase the oxygen demand [36]. The significant research conducted in this direction revealed that ROS generation takes place through numerous mechanisms, and play a vital role in developing neurodegenerative diseases like Alzheimer’s, Huntington’s and Parkinson’s disease. These ROS considerably affects the neuron and other cellular processes, the controlling of ROS level may serve as a potential treatment to restrain neurodegenerative disorder and provide relief from its associated symptoms [37].

2.4. Respiratory Disorders

Asthma and chronic obstructive pulmonary disease (COPD) are major respiratory disorders, accounting for high mortality worldwide [38]. Exposure to cigarette smoke and air pollutants significantly contributes to the overproduction of ROS in both asthma and COPD patients. The ROS primarily affects and damages the alveolar and connective tissues of the pulmonary system [39]. The overproduction of ROS also triggers the inflammatory cells, which, as a result, shows production of ROS in the pulmonary system. The ROS are predominantly observed during pathophysiology analysis for both asthma and COPD [40]. It is still unclear how increased ROS is a causative factor for these respiratory diseases [41]. Researchers are extensively working in this direction to decipher the role of these ROS in progression of these fatal diseases.
This synergism between ROS and these chronic diseases shows the major challenge associated with oxidative stress induced by ROS, and requires the effective solution to meet the challenges imposed by the overproduction of ROS.

3. Antioxidants

An antioxidant can be described as any substance or compound capable of inhibiting the oxidation of suitable substrate even when present in low concentrations [42]. During the late 19th and early 20th century, the exploration of antioxidants resulted in a boom, due to their involvement in various industrial processes like the prevention of corrosion, the polymerization of fuels, fouling in combustion engines and the vulcanization of rubber [43]. The application of antioxidants was limited for the prevention of the oxidation of unsaturated fats, as it resulted in the rancidity of fats [44]. The general procedure to determine the antioxidant potential of any compound involves the assessment of the rate of oxygen consumption when fat is kept in an enclosed container with oxygen. The identification of vitamins A, C and E, which act as antioxidant agents, has revolutionized the field and highlighted the significance of antioxidants in the biochemistry of living beings [45,46].
Another way to understand the antioxidant is that it is a stable molecule, which donates an electron to unwanted free radical species and neutralizes it, and curbs its ability to cause damage. In general, these antioxidants either inhibit or delay the cellular damage because of their scavenging properties [47]. The low molecular weight of these antioxidants allows them to interact with ROS (free radicals) easily and terminate their chain reaction before damaging vital molecules. Glutathione, uric acid and ubiquinol are few antioxidant molecules that are generated by our body during normal metabolic processes [48]. There are various enzymes are present in our body that can scavenge free radicals, and micronutrients like ascorbic acid (vitamin C), β-carotene and α-tocopherol (vitamin E) [49]. The body cannot produce these micronutrients on its own, therefore, these molecules are obtained from the consumed food.

4. Sources of Antioxidants

Dietary supplements are key source of antioxidants, which could aid in maintaining good health and prevent the onset of fatal diseases triggered by ROS. Even though some synthetic antioxidants have been developed, their carcinogenic and toxic nature has prompted the exploration for natural antioxidants like vitamins A, C and E [50]. Additionally, population studies have also revealed that the consumption of fruits, tea, vegetables and wine are a reliable source of natural antioxidants, and are effective in regulating the risk of cardiovascular diseases, which has intrigued researchers to exploring their potential as natural antioxidants [51]. Dietary supplements contain antioxidant compounds in the form of phytochemicals i.e., α-tocopherol, β-carotene, vitamin C, vitamin E and various phenolic compounds [52]. Numerous ethnomedicinal plants, fruits, vegetables, mushrooms and other spices have been well-documented as sources of natural antioxidants, which play a significant role in promoting healthy life and treating various fatal diseases [53].
Phenolic compounds obtained from natural sources are considered far better than synthesized chemicals. Microbes are also being explored for synthesizing organic compounds with antioxidant potential. Various fungal strains have been reported to produce compounds like ellagic acid, ferulic acid and gallic acid under solid-state fermentation and submerged fermentation conditions [54]. All these compounds are known to contain 2–4 reactive hydroxyl groups, which impart them the antioxidant potential. In addition, algae and lichens are active producers of different secondary metabolites, including antioxidants (Table 1).

5. Nano-Antioxidants

Antioxidants have been accorded as effective therapeutic and prophylactic agents for various diseases. However, these antioxidants have received very limited success until now, as most of the antioxidants show low permeability, and are poorly soluble in water, demonstrate instability during storage and gastrointestinal degradation, which are some of their limitations [61]. The amalgamation of material sciences with nanotechnology has substantially improved and reduced the free radical synthesis during nanoparticle production in different areas and the nanoparticles synthesized for this purpose are regarded as nano-antioxidants [62,63]. Carbon nanotubes, metal and metal oxide nanoparticles and various types of polymer-loaded antioxidant nanoparticles, have been reported to exhibit antioxidant properties [63]. In the past few decades, various preparation protocols, such as emulsion/solvent evaporation, supercritical fluid technology, solvent displacement method, templating method and nanoprecipitation techniques, have been used for synthesizing nano-oxidants [63]. Some oxide nanoparticles can scavenge the reactive nitrogen and reactive oxygen species (RNS/ROS) and mimic the antioxidant molecule, due to their intrinsic physicochemical properties [64]. In the biomedical field, cerium oxide nanoparticles (CONPs) have gathered special attention for their multi-enzymatic scavenging of ROS and their regenerative abilities [65]. These CONPs have unique properties, like the coexistence in both oxidation states i.e., Ce3+ and Ce4+, the ability to reversibly switch between both oxidation states and the reduction potential of ~1.52 V [66]. Cerium dioxide as a bulk crystal primarily contains Ce4+, but during its reduction to nano-size, substantially enhances the relative amount of Ce3+, therefore, leading to higher catalytic activity, in contrast to various biological processes and biological antioxidants [67,68]. Hirst et al. (2013) conducted an in vivo test on mice to assess the antioxidant potential of nanoceria, which were injected intravenously in the subject, and the result of the study revealed that nanoceria significantly decreased the lipoperoxidation after the three weeks, which indicates that CONPs are effective in treating oxidative stress [69]. Caputo et al. (2015) conducted a comparative study to assess the antioxidant potential between CONPs and NAC (N-acetyl-cysteine) and Trolox (soluble analogues of vitamin E) [70]. The results of this study revealed that NAC and Trolox reduced the oxidative 2′-7′-Dichlorofluorescein (DCF) signal triggered by irradiated TiO2 nanoparticles, but the antioxidant potential was significantly lower in comparison to CONPs. This result also highlights the stability of CONPs because of their auto-regenerative redox cycle, which allows them to surpass the challenges related to the stability of the antioxidants molecules.
On the other hand, synthetic polymeric NPs have emerged as the promising nano-drug delivery system, as they can encapsulate the therapeutic agent and progressively release the therapeutic compound at the target site. Poly-D, L-lactide (PLA) and poly (lactic-co-glycolic acid) (PLGA) are some examples of synthetic biodegradable polymers that have been approved safe by the European Medicine Agency (EMA) and U.S. Food and Drug Administration (FDA) for administration (Table 2).
Liposomes are also used for delivering the antioxidant agents to the target site. The amphiphilic and biocompatible nature of these liposomes allows them to load both hydrophilic and lipophilic compounds and favor the encapsulation of the water-soluble and water-insoluble antioxidant enzymes [79].
Furthermore, chitosan is the material predominantly used for synthesizing nanoparticles as a sole material or in amalgamation with another [80]. Chitosan shows mucoadhesive properties, which improves the targeted delivery in mucosal surfaces such as intestinal and nasal epithelium [81]. Curcumin encapsulated in nanocarrier and covered and stabilized with chitosan has also been developed and evaluated for free radical scavenging in comparison with free curcumin, and showed the protective effect of chitosan on the antioxidant activity of curcumin [82]. Pu et al. (2014) reported the encapsulation of curcumin antioxidant compounds within the nanocarrier and regulation of release of antioxidant compounds, by changing the pH and oxidative stress of inflamed tissues to increase the overproduction of RNS/ROS synthesized by lipopolysaccharide (LPS)-stimulated macrophage [83].

6. Antioxidant Functionalized Nanoparticles

Bacteria, algae, fungi, lichens and plants are known to contain diverse bioactive compounds like terpenoids, alkaloids, polyphenols, phenolic acids etc. These bioactive compounds show potential antioxidant activity, and are known to reduce and stabilize the metallic ions. The diverse types of antioxidant functionalized nanoparticles derived from various biological extracts (Table 3) are discussed in the following sections.

6.1. Silver Nanoparticles (AgNPs)

The silver nanoparticles (AgNPs), show a substantial biochemical and catalytic activity, which can be attributed to their significantly large surface area, as compared with other particles with similar chemical structures [134]. The AgNPs synthesis occurs via two steps, at first, Ag+ ions are reduced to Ag°, followed by the agglomeration of colloidal silver nanoparticles, to form the oligomeric clusters which are finally stabilized [134]. Biological catalysts (enzymes) are required to reduce the Ag+ ions, and the production of AgNPs with characteristic antioxidant properties can be achieved with a variety of plant extracts, as shown in Table 3. Patra et al. (2016) extracted the aqueous portion of watermelon rind (WRA), and used it to synthesize the AgNPs under light-exposed conditions at room temperature [92].
In recent studies, actinobacteria have been identified from various ecosystems that are potential natural producers of AgNPs and the whole-cell biomass and cell free extract of Streptomyces naganishii MA7 and Streptomyces griseorubens AU2, respectively, were used to synthesize silver nanoparticles with potential antioxidant properties [108,111]. On the other hand, fungal species like Aspergillus versicolor ENT7, Cladosporium cladosporioides and Pestalotiopsis microspore have also been used for the production of antioxidant functionalized AgNPs [117,119,120]. Various researchers have also reported Ganoderma lucidium as a potential source of antioxidant functionalized AgNPs [126,128,129]. Venkatesan et al. (2016) prepared the extracts of Ecklonia cava, a marine alga known to be a reservoir of phenolic compounds that can act as capping and reducing agents, and used it for the synthesis of silver nanoparticles [59]. Several lichens have also been identified with potential antioxidant properties, and extracts of Parmeliopsis ambigua, Punctelia subrudecta, Evernia mesomorpha, and Xanthoparmelia plitti were used to synthesize AgNPs [132].
The AgNPs synthesized using the leaves extract of Clerodendrum phlomidis, are found to have ferric reducing power of 1.63 AU, which is higher than the leaves extract used alone [87]. Whereas, on comparing the IC50 value of the extract (i.e., 1920 μg/mL), the AgNPs synthesized using the leaves extract of Clerodendrum phlomidis showed higher scavenging activity as confirmed by IC50 value (i.e.,55.86 μg/mL). Furthermore, the DPPH radical scavenging activity of AgNPs was found to be dose-dependent, showing the maximum inhibition (85.74%) greater than that of the extract alone. The AgNPs IC50 value (9.12 μg/mL) was found to be significantly lower, as compared to the extract (IC50 value 388.4 μg/mL) and standard ferulic acid (182.8 μg/mL), which confirmed the AgNPs with high antioxidant potential. In 2019, Das and his colleagues reported that AgNPs synthesized using Morus alba leaves extract increased the DPPH scavenging activity to 47.81% against the ~56% of the activity shown by standard ascorbic acid at the same concentration. However, plant extract mediated AgNPs showed 95.08% ABTS+ scavenging activity, which is the equivalent to that which was shown by the BTH standard i.e., 95.51% at 100μg/mL, therefore indicating its strong scavenging potential. On the other hand, for nitric oxide scavenging activity, AgNPs displayed 64.04% scavenging, as compared to 45.72% and 88.62% in plant extract and gallic acid standard, respectively, at 100 μg/mL. Similarly, AgNPs at 100μg/mL concentration have shown a significant superoxide scavenging activity of 81.92%, as compared to the 85.35% present in the tocopherol standard [90].

6.2. Gold Nanoparticles (AuNPs)

Gold nanoparticles (AuNPs) have attained significant attention, owing to their physical characteristics (size and shape), optical properties and biocompatibility [134]. Gold nanoparticles of varied sizes and diverse morphologies have been employed in the medical sector for various purposes, such as the detection of tumors, and as a carrier for drugs etc. (e.g., Paclitaxel) [134]. Antioxidant AuNPs are mostly derived from extracts of plant parts like leaves and fruits, as shown in Table 3. Markus et al. (2016) reported the identification of Lactobacillus kimchicus DCY51T, a novel probiotic from Korean kimchi, using an intracellular membrane-bound approach, and synthesized antioxidant functionalized AuNPs from these bacteria [107]. It was observed that the AuNPs formed a capping layer which comprised of several amino acids, while the presence of surface-bound proteins rendered them harmless to cancerous cell lines i.e., human colorectal adenocarcinoma (HT29) and murine macrophage (RAW264.7). As compared with gold salts, biological produced gold nanoparticles were found to be better scavengers of free radicals, especially DPPH. A food inhabiting Enterococcus species was identified and used for the production of gold nanoparticles, and was further confirmed through techniques like Fourier transform infrared spectroscopy (FT-IR), UV-vis absorption spectroscopy and transmission electron microscopy (TEM) [109].
Veeraapandian et al. (2012) utilized extracellular proteins of Escherichia coli to synthesize gold nanoparticles of anisotropic nature [110]. The size and shape of AuNPs are greatly influenced by the quantity of protein, as extracellular proteins significantly increases the shelf life of nanoparticles by acting as a capping agent, and thereby conferring them with stability. Manjunath et al. (2017) in their study, used a fungal endophyte, Cladosporium cladosporioides, obtained from Sargassum wightii, seaweed, to synthesize gold nanoparticles [118]. The process of reduction of gold metal salts into gold nanoparticles involved the use of phenolic compounds, and was found to be mediated through the activity of NADPH-dependent reductase enzyme. Lee et al. (2015) used the extracts of a mushroom species, Inonotus obliquus, to develop AuNPs without application of deleterious agents [127]. In another study, AuNPs were synthesized from the extracts of Gracilaria corticate, a red alga inhabiting in the marine environment and using it as a reducing agent [130]. Sharma et al. (2014) used the desiccated biomass of Lemanea fluviatilis to reduce and stabilize the AuNPs [131]. Debnath et al. (2016) developed the gold nanoparticles without the use of extraneous stabilizing and reducing agents by using desiccated biomass of lichens inhabiting in the alpine region of Eastern Himalayas (India) at high altitudes [133]. The gold nanoparticles obtained from lichen Acroscyphus sp. comprised of prismatic and multiply twinned quasi-spherical shapes, while those synthesized from Sticta sp. were simply multiply twinned and, both of these exhibited antioxidant activity.

6.3. Copper Oxide Nanoparticles (Cu2ONPs)

In recent years, copper (Cu) has intrigued researchers for its application in nanoparticle synthesis, due to its readily available nature, and characteristics like catalytic, electrical, optical and mechanical properties [135,136,137]. Copper oxide, being a vital inorganic material, is predominantly used in modern technologies, especially in the field of catalysis, ceramics and superconductor applications. Moreover, it can also be used as electrode active materials for the degradation of nitrous oxide with ammonia, and oxidizing carbon monoxide, hydrocarbon and phenol in the formation of supercritical water [138]. The phytogenic synthesis of Cu2ONPs possessing antioxidant ability is compiled in Table 3. In 2019, Rajeshkumar and his colleagues also reported a similar radical scavenging property of copper nanoparticles synthesized using leaf extract of Cissus arnotiana, when compared with the standard ascorbic acid [102].

6.4. Iron Nanoparticles (INPs)

Iron is another imperative material used for synthesizing nanoparticles, because of unique physiochemical properties like high catalytic activity, low toxicity, high magnetism and microwave absorption ability [139,140,141]. The iron nanoparticles (INPs) are categorized into three groups i.e., (a) iron oxide nanoparticles (IONs), (b) iron oxide hydroxide (FeOOH) nanoparticles, and (c) zero-valent iron (ZVI) nanoparticles [142,143,144,145]. These nanoparticles are used for varied applications like bio-separation, bioprocess intensification, drug delivery, environmental remediation, ferrofluids, food preservation, gene therapy, hyperthermia, magnetic targeting, negative MRI contrast enhancement, pigments, stem cell sorting and manipulation, thermal-ablation and lithium-ion batteries [146]. Table 3 shows the plant mediated INPs with antioxidant potential. In 2015, Muthukumar and Manickam observed the high antioxidant potential of Amaranthus spinosus leaf extract mediated INPs, because of the presence of amaranthine and phenolic compounds in it, which act as the capping agent [104].

6.5. Zinc Oxide (ZnONPs), Selenium (SeNPs) and Nickel Oxide Nanoparticles (NiONPs)

Zinc oxide is a metal oxide generally used in n-type of semiconductors. In recent decades, ZnONPs have drawn significant attention, owing to their widespread application in various fields like electronics, optical and the biomedical sector [147]. The process of production of zinc oxide nanoparticles is cost-effective, safe and easy, and ZnO has been given the status of generally recognized as safe (GRAS) by the US Food and Drug Administration [148,149]. ZnONPs are primarily known for their exceptional semiconducting properties, which can be attributed to the wide band gap of 3.37 eV, high catalytic activity, UV filtering properties and large exciton binding energy of 60 meV, and also have good optical, wound healing and anti-inflammatory properties [147]. ZnONPs have been comprehensively used in the cosmetic industry, in products like sunscreen lotions, for their intrinsic UV filtering properties [150]. They also have widespread applications in the biomedical sector, such as in drug delivery, and also exhibit antidiabetic, antibacterial anticancer and antifungal properties [147]. Chandra et al. (2019) used the leaf extract of Berberis aristata, a plant of medicinal importance, to synthesize zinc oxide nanoparticles [95]. Other studies have employed the yeast species, Pichia kudriavzevii for the extracellular biosynthesis of ZnONPs and the development of ZnONPs by using the fungal strain, Aspergillus niger [116,122].
In addition to these materials, selenium nanoparticles (SeNPs) are also attracting researchers, owing to their enhanced properties like semi-conduction, photoelectrical, photoconduction, catalytic etc., and their potential in optical and electronic instruments. They are known to have lesser toxic effects than selenium (Se) compounds, and are used in the medical field, as they show high therapeutic and anticancer properties [151]. Ramya et al. (2015) used an actinomycetes species, Streptomyces minutiscleroticus M10A62, obtained from magnesite mine, to develop SeNPs [98]. In 2012, Torres and his colleagues used Pantoea agglomerans isolated from the Camarones River to synthesize Se nanoparticles [113]. In 2013, Li and team members fabricated 6-hydroxy-2,5,7,8-tetramethylchroman-2-carboxylic acid (Trolox) coated surface-functionalized nanoparticles of Selenium (Se@Trolox) with antioxidant potential [152]. Moreover, Se@Trolox was found to block the activation of the AKT and MAPK signaling pathway, cisplatin-induced reactive oxygen species (ROS) accumulation, and DNA damage-mediated p53 phosphorylation in HK-2 cells [152].
Recently, nickel oxide has been found to perform different functions in various fields like biomedicine, electronics and magnetism, owing to its properties like anti-bacterial, anti-inflammatory, eco-friendliness, easy usage and high reactiveness [153]. Being highly reactive, it is readily used for catalyzing various organic reactions, like the α-alkylation of methyl ketone, the chemo-selective oxidative coupling of thiols, the hydrogenation of olefins, the synthesis of stilbenes from alcohol through Wittig-type olefination and the reduction of ketones and aldehydes [154,155,156,157,158]. Moreover, it is also found to catalyze inorganic reactions, such as the decomposition of ammonia [159]. Recently, it has been used for developing carbon nanotubes (CNTs) [160]. The plant derived nickel oxide nanoparticles showing antioxidant potential are shown in Table 3.

7. Challenges

Apart from the widespread applications of metallic nanoparticles, owing to their size, chemical composition, shape, stability, functionalization, surface coating and purity, they also possess potential toxic effects [161]. Nanoparticles show characteristically distinct cellular uptake mechanisms, and also exhibit the ability to catalyze the synthesis of ROS, thus, leading to ROS associated toxic effects [162]. The size of the NPs is also known to affect the uptake of NPs and their distribution within the cell significantly, and it has been found that in NPs with the same dosage but having distinct sizes; the small-sized NPs are readily internalized within the cell than the large-sized NPs [163]. Moreover, it has been observed that the small-sized NPs possess relatively high reactivity, as they have a large surface area [164]. The size, surface charge and the material type of NPs determine the aggregating efficiency of MNPs, which further influences the internalization of NPs in the cell, and ultimately affects the NPs associated toxicity [165]. A lot of research is being carried out at a global scale to evaluate the NP’s associated toxic effects. Cho et al. (2009) found that polyethyleneglycol (PEG)-coated AuNPs triggers acute inflammation, as well as apoptosis, in the liver after intravenous injection, and also leads to the aggregation of the AuNPs in the cytoplasmic vesicles, liver and spleen [166]. There are numerous transition metals like copper (Cu), chromium (Cr), iron (Fe), silicon (Si) and vanadium (V), which are also involved in the generation of ROS via the Fenton reaction and the Haber-Weiss mechanism [167]. Apart from this, ZnONPs have also been reported to increase the cytotoxicity, due to the formation of ROS, which causes oxidative damage, and evoke the release of inflammatory mediators, thus, resulting in the senescence of phagocytic RAW264.7 cells and the transformation of BEAS-2B (human bronchial epithelial) cells [168]. These are some of the challenges associated with the use of metallic and metalloid NPs, and continuous efforts are being made to eliminate and overcome these challenges.

8. Conclusions

A sedentary lifestyle and the consumption of high carbohydrate, proteins and fat have drastically changed the lives of humans, resulting in the production of ROS, which subsequently leads to oxidative stress. Moreover, these oxidative stresses are known to be linked with various diseases, and many attempts have been made to subside these with different medications. Although conventional therapies using antioxidants have been used in the past, they were mostly found ineffective in treating various diseases because of their incompetence in passing through the blood–brain barrier. To overcome these challenges, several antioxidant functionalized nanoparticles have been derived from biological sources, and evaluated by using scavenging assays under invitro conditions. Antioxidant properties of these nanocarriers make it a suitable candidate for targeted delivery, and can open new opportunities for combating oxidative stress in in vivo conditions. Furthermore, researchers are making continuous effort to slow down the toxic effects associated with the metallic and metalloid nanoparticles.

Author Contributions

Conceptualization, E.N., K.K., and D.K.; Manuscript writing, H.K. and K.B.; Manuscript editing, D.S.D., S.B., and R.V.; Critical revising, E.N., K.K., S.K.B. and D.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the University of Hradec Kralove (Faculty of Science VT2019-2021).

Acknowledgments

We acknowledge university of Hradec Kralove (Faculty of Science, VT2019-2021) for financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dabhade, P.; Kotwal, S. Tackling the aging process with biomolecules: 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]
  2. López-Otín, C.; Blasco, M.A.; Partridge, L.; Serrano, M.; Kroemer, G. The hallmarks of aging. Cell 2013, 153, 1194–1217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Shokolenko, I.N.; Wilson, G.L.; Alexeyev, M.F. Aging: A mitochondrial DNA perspective, critical analysis and an update. World J. Exp. Med. 2014, 4, 46. [Google Scholar] [CrossRef] [PubMed]
  4. Chang, C.H.; Lee, K.Y.; Shim, Y.H. Normal aging: Definition and physiologic changes. J. Korean Med. Assoc. 2017, 60, 358–363. [Google Scholar] [CrossRef] [Green Version]
  5. Harman, D. Aging: A theory based on free radical and radiation chemistry. J. Gerontol. 1956, 11, 298–300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Islam, M.T. Oxidative stress and mitochondrial dysfunction-linked neurodegenerative disorders. Neurol. Res. 2017, 39, 73–82. [Google Scholar] [CrossRef]
  7. Valentão, 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] [PubMed] [Green Version]
  8. Dias, V.; Junn, E.; Mouradian, M.M. The role of oxidative stress in Parkinson’s disease. J. Parkinson’s Dis. 2013, 3, 461–491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Halliwell, B.; Gutteridge, J.M.C. Free Radicals in Biology and Medicine, 4th ed.; Clarendon Press: Oxford, UK, 2007. [Google Scholar]
  10. Bahorun, T.; Soobrattee, M.A.; Luximon-Ramma, V.; Aruoma, O.I. Free radicals and antioxidants in cardiovascular health and disease. Internet J. Med. Update 2006, 1, 25–41. [Google Scholar] [CrossRef] [Green Version]
  11. Kumar, S.; Pandey, A.K. Free radicals: Health implications and their mitigation by herbals. Br. J. Med. Med. Res. 2015, 7, 438–457. [Google Scholar] [CrossRef]
  12. Kumar, S.; Pandey, A.K. Chemistry and biological activities of flavonoids: An overview. Sci. World J. 2013, 2013, 162750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Valko, M.; Izakovic, M.; Mazur, M.; Rhodes, C.J.; Telser, J. Role of oxygen radicals in DNA damage and cancer incidence. Mol. Cell. Biochem. 2004, 266, 37–56. [Google Scholar] [CrossRef] [PubMed]
  14. Valko, M.; Leibfritz, D.; Moncola, J.; Cronin, M.D.; Mazur, M.; Telser, J. Free radicals and antioxidants in normal physiological functions and human disease. Int. J. Biochem. Cell Biol. 2007, 39, 44–84. [Google Scholar] [CrossRef] [PubMed]
  15. Dröge, W. Free radicals in the physiological control of cell function. Physiol. Rev. 2002, 82, 47–95. [Google Scholar] [CrossRef]
  16. Willcox, J.K.; Ash, S.L.; Catignani, G.L. Antioxidants and prevention of chronic disease. Crit. Rev. Food Sci. Nutr. 2004, 44, 275–295. [Google Scholar] [CrossRef]
  17. Pacher, P.; Beckman, J.S.; Liaudet, L. Nitric oxide and peroxynitrite in health and disease. Physiol. Rev. 2007, 87, 315–424. [Google Scholar] [CrossRef] [Green Version]
  18. Genestra, M. Oxyl radicals, redox-sensitive signaling cascades and antioxidants. Cell. Signal. 2007, 19, 1807–1819. [Google Scholar] [CrossRef]
  19. Halliwell, B. Biochemistry of oxidative stress. Biochem. Soc. Trans. 2007, 35, 1147–1150. [Google Scholar] [CrossRef]
  20. 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]
  21. Sharma, P.; Mehta, M.; Dhanjal, D.S.; Kaur, S.; Gupta, G.; Singh, H.; Thangavelu, L.; Rajeshkumar, S.; Tambuwala, M.; Bakshi, H.A.; et al. Emerging trends in the novel drug delivery approaches for the treatment of lung cancer. Chem. Biol. Interact. 2019, 309, 108720. [Google Scholar] [CrossRef]
  22. Aggarwal, V.; Tuli, H.S.; Varol, A.; Thakral, F.; Yerer, M.B.; Sak, K.; Varol, M.; Jain, A.; Khan, M.; Sethi, G. Role of reactive oxygen species in cancer progression: Molecular mechanisms and recent Advancements. Biomolecules 2019, 9, 735. [Google Scholar] [CrossRef] [Green Version]
  23. Liou, G.Y.; Storz, P. Reactive oxygen species in cancer. Free Radic. Res. 2010, 44, 479–496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Qian, Q.; Chen, W.; Cao, Y.; Cao, Q.; Cui, Y.; Li, Y.; Wu, J. Targeting reactive oxygen species in cancer via Chinese herbal medicine. Oxid. Med. Cell. Longev. 2019, 2019, 9240426. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Kumari, S.; Badana, A.K.; Murali, M.G.; Shailender, G.; Malla, R. Reactive oxygen species: A key constituent in cancer survival. Biomark. Insights 2018, 13, 1177271918755391. [Google Scholar] [CrossRef] [Green Version]
  26. Yang, H.; Villani, R.M.; Wang, H.; Simpson, M.J.; Roberts, M.S.; Tang, M.; Liang, X. The role of cellular reactive oxygen species in cancer chemotherapy. J. Exp. Clin. Cancer Res. 2018, 37, 266. [Google Scholar] [CrossRef] [PubMed]
  27. Schieber, M.; Chandel, N.S. ROS function in redox signaling and oxidative stress. Curr. Biol. 2014, 24, R453–R462. [Google Scholar] [CrossRef] [Green Version]
  28. Redza-Dutordoir, M.; Averill-Bates, D.A. Activation of apoptosis signalling pathways by reactive oxygen species. Biochim. Biophys. Acta 2016, 1863, 2977–2992. [Google Scholar] [CrossRef]
  29. Mehta, M.; Dhanjal, D.S.; Paudel, K.R.; Singh, B.; Gupta, G.; Rajeshkumar, S.; Thangavelu, L.; Tambuwala, M.M.; Bakshi, H.A.; Chellappan, D.K.; et al. Cellular signalling pathways mediating the pathogenesis of chronic inflammatory respiratory diseases: An update. Inflammopharmacology 2020, 1–23. [Google Scholar] [CrossRef]
  30. Kessenbrock, K.; Plaks, V.; Werb, Z. Matrix metalloproteinases: Regulators of the tumor microenvironment. Cell 2010, 141, 52–67. [Google Scholar] [CrossRef] [Green Version]
  31. He, F.; Zuo, L. Redox roles of reactive oxygen species in cardiovascular diseases. Int. J. Mol. Sci. 2015, 16, 27770–27780. [Google Scholar] [CrossRef] [Green Version]
  32. Panth, N.; Paudel, K.R.; Parajuli, K. Reactive oxygen species: A key hallmark of cardiovascular disease. Adv. Med. 2016, 2016, 9152732. [Google Scholar] [CrossRef] [Green Version]
  33. Sag, C.M.; Santos, C.X.; Shah, A.M. Redox regulation of cardiac hypertrophy. J. Mol. Cell Cardiol. 2014, 73, 103–111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Zhou, T.; Prather, E.R.; Garrison, D.E.; Zuo, L. Interplay between ROS and antioxidants during ischemia-reperfusion injuries in cardiac and skeletal muscle. Int. J. Mol. Sci. 2018, 19, 417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Van der Pol, A.; Van Gilst, W.H.; Voors, A.A.; Van der Meer, P. Treating oxidative stress in heart failure: Past, present and future. Eur. J. Heart Fail. 2019, 21, 425–435. [Google Scholar] [CrossRef] [PubMed]
  36. Uttara, B.; Singh, A.V.; Zamboni, P.; Mahajan, R.T. Oxidative stress and neurodegenerative diseases: A review of upstream and downstream antioxidant therapeutic options. Curr. Neuropharmacol. 2009, 7, 65–74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Liu, Z.; Zhou, T.; Ziegler, A.C.; Dimitrion, P.; Zuo, L. Oxidative stress in neurodegenerative diseases: From molecular mechanisms to clinical applications. Oxid. Med. Cell. Longev. 2017, 2017, 2525967. [Google Scholar] [CrossRef]
  38. Khaltaev, N.; Axelrod, S. Chronic respiratory diseases global mortality trends, treatment guidelines, life style modifications, and air pollution: Preliminary analysis. J. Thorac. Dis. 2019, 11, 2643–2655. [Google Scholar] [CrossRef]
  39. Boukhenouna, S.; Wilson, M.A.; Bahmed, K.; Kosmider, B. Reactive oxygen species in chronic obstructive pulmonary disease. Oxid. Med. Cell. Longev. 2018, 2018, 5730395. [Google Scholar] [CrossRef]
  40. Thimmulappa, R.K.; Chattopadhyay, I.; Rajasekaran, S. Oxidative stress mechanisms in the pathogenesis of environmental lung diseases. In Oxidative Stress in Lung Diseases; Chakraborti, S., Chakraborti, T., Ghosh, R., Ganguly, N.K., Parinandni, N.L., Eds.; Springer: Singapore, 2020; Volume 2, pp. 103–137. [Google Scholar]
  41. Pizzino, G.; Irrera, N.; Cucinotta, M.; Pallio, G.; Mannino, F.; Arcoraci, V.; Squadrito, F.; Altavilla, D.; Bitto, A. Oxidative stress: Harms and benefits for human health. Oxid. Med. Cell. Longev. 2017, 2017, 8416763. [Google Scholar] [CrossRef]
  42. Young, I.S.; Woodside, J.V. Antioxidants in health and disease. J. Clin. Pathol. 2001, 54, 176–186. [Google Scholar] [CrossRef] [Green Version]
  43. Matill, H.A. Antioxidants. Annu. Rev. Biochem. 1947, 16, 177–192. [Google Scholar] [CrossRef] [PubMed]
  44. German, J. Food processing and lipid oxidation. Adv. Exp. Med. Biol. 1999, 459, 23–50. [Google Scholar] [PubMed]
  45. Jacob, R. Three eras of vitamin C discovery. Subcell. Biochem. 1996, 25, 1–16. [Google Scholar]
  46. Knight, J. Free radicals: Their history and current status in aging and disease. Ann. Clin. Lab. Sci. 1998, 28, 331–346. [Google Scholar]
  47. Halliwell, B. How to characterize an antioxidant- An update. Biochem. Soc. Symp. 1995, 61, 73–101. [Google Scholar]
  48. Shi, H.L.; Noguchi, N.; Niki, N. Comparative study on dynamics of antioxidative action of α- tocopheryl hydroquinone, ubiquinoland α- tocopherol, against lipid peroxidation. Free Radic. Biol. Med. 1999, 27, 334–346. [Google Scholar] [CrossRef]
  49. Levine, M.; Ramsey, S.C.; Daruwara, R. Criteria and recommendation for vitamin C intake. JAMA 1991, 281, 1415–1423. [Google Scholar] [CrossRef]
  50. Yang, X.; Sun, Z.; Wang, W.; Zhou, Q.; Shi, G.; Wei, F.; Jiang, G. Developmental toxicity of synthetic phenolic antioxidants to the early life stage of zebrafish. Sci. Total. Environ. 2018, 643, 559–568. [Google Scholar] [CrossRef]
  51. Aguilera, Y.; Martin-Cabrejas, M.A.; González de Mejia, E. Phenolic compounds in fruits and beverages consumed as part of the mediterranean diet: Their role in prevention of chronic diseases. Phytochem. Rev. 2016, 15, 405–423. [Google Scholar] [CrossRef]
  52. Faustino, M.; Veiga, M.; Sousa, P.; Costa, E.M.; Silva, S.; Pintado, M. Agro-food byproducts as a new source of natural food additives. Molecules 2019, 24, 1056. [Google Scholar] [CrossRef] [Green Version]
  53. Chandra, P.; Sharma, R.K.; Arora, D.S. Antioxidant compounds from microbial sources: A review. Food Res. Int. 2020, 129, 108849. [Google Scholar] [CrossRef]
  54. Dey, T.B.; Chakraborty, S.; Jain, K.K.; Sharma, A.; Kuhad, R.C. Antioxidant phenolics and their microbial production by submerged and solid state fermentation process: A review. Trends Food Sci. Technol. 2016, 53, 60–74. [Google Scholar]
  55. Brewer, M.S. Natural antioxidants: Sources, compounds, mechanisms of action, and potential applications. Compr. Rev. Food Sci. Food Saf. 2011, 10, 221–247. [Google Scholar] [CrossRef]
  56. Kozarski, M.; Klaus, A.; Jakovljevic, D.; Todorovic, N.; Vunduk, J.; Petrović, P.; Niksic, M.; Niksic, M.M.; Griensven, L.V. Antioxidants of edible mushrooms. Molecules 2015, 20, 19489–19525. [Google Scholar] [CrossRef] [Green Version]
  57. Ferreira, I.C.F.R.; Barros, L.; Abreu, R.M.V. Antioxidants in wild mushrooms. Curr. Med. Chem. 2009, 16, 1543–1560. [Google Scholar] [CrossRef] [Green Version]
  58. Munir, N.; Sharif, N.; Naz, S.; Manzoor, F. Algae: A potent antioxidant source. Sky J. Microbiol. Res. 2013, 1, 22–31. [Google Scholar]
  59. Venkatesan, J.; Kim, S.K.; Shim, S.K. Antimicrobial, antioxidant, and anticancer activities of biosynthesized silver nanoparticles using marine algae Ecklonia cava. Nanomaterials 2016, 6, 235. [Google Scholar] [CrossRef]
  60. Fernández-Moriano, C.; Gómez-Serranillos, M.P.; Crespo, A. Antioxidant potential of lichen species and their secondary metabolites. A systematic review. Pharm. Biol. 2015, 54, 1–17. [Google Scholar] [CrossRef]
  61. 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] [PubMed] [Green Version]
  62. Eftekhari, A.; Ahmadian, E.; Panahi-Azar, V.; Hosseini, H.; Tabibiazar, M.; Dizaj, S.M. Hepatoprotective and free radical scavenging actions of quercetin nanoparticles on aflatoxin B1-induced liver damage: In vitro/in vivo studies. Artif. Cells Nanomed. Biotechnol. 2017, 46, 411–420. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. 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] [PubMed]
  64. 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] [PubMed] [Green Version]
  65. 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]
  66. 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]
  67. Deshpande, S.; Patil, S.; Kuchibhatla, S.V.N.T.; Seal, S. Size dependency variation in lattice parameter and valency states in nanocrystalline cerium oxide. Appl. Phys. Lett. 2005, 87, 133113. [Google Scholar] [CrossRef]
  68. Kim, C.K.; Kim, T.; Choi, I.Y.; Soh, M.; Kim, D.; Kim, Y.J.; Jang, H.; Yang, H.S.; Kim, J.Y.; Park, H.K.; et al. Ceria nanoparticles that can protect against ischemic stroke. Angew. Chem. Int. Ed. Engl. 2012, 51, 11039–11043. [Google Scholar] [CrossRef] [PubMed]
  69. 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. Envrion. Toxicol. 2013, 28, 107–118. [Google Scholar] [CrossRef]
  70. Caputo, F.; Nicola, M.D.; 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]
  71. Sonaje, K.; Italia, J.L.; Sharma, G.; Bhardwaj, V.; Tikoo, K.; Kumar, M.N.V.R. Development of biodegradable nanoparticles for oral delivery of ellagic acid and evaluation of their antioxidant efficacy against cyclosporine A-induced nephrotoxicity in rats. Pharm. Res. 2007, 24, 899–908. [Google Scholar] [CrossRef]
  72. Yun, X.; Maximov, V.D.; Yu, J.; Zhu, H.; Vertegel, A.A.; Kindly, M.S. Nanoparticles for targeted delivery of antioxidant enzymes to the brain after cerebral ischemia and reperfusion injury. J. Cereb. Blood Flow. Metab. 2013, 33, 583–592. [Google Scholar] [CrossRef] [Green Version]
  73. Chorny, M.; Hood, E.; Levy, R.J.; Muzykantov, V.R. Endothelial delivery of antioxidant enzymes loaded into non-polymeric magnetic nanoparticles. J. Control. Release 2010, 146, 144–151. [Google Scholar] [CrossRef] [Green Version]
  74. Reddy, M.K.; Labhasetwar, V. Nanoparticle-mediated delivery of superoxide dismutase to the brain: An effective strategy to reduce ischemia-reperfusion injury. FASEB J. 2009, 23, 1384–1395. [Google Scholar] [CrossRef]
  75. Fan, Y.; Yi, J.; Zhang, Y.; Yokoyama, W. Fabrication of curcumin-loaded bovine serum albumin (BSA)-dextran nanoparticles and the cellular antioxidant activity. Food Chem. 2018, 239, 1210–1218. [Google Scholar] [CrossRef] [PubMed]
  76. Elle, R.E.; Rahmani, S.; Lauret, C.; Morena, M.; Bidel, L.P.R.; Boulahtouf, A.; Balaguer, P.; Cristol, J.P.; Durand, J.O.; Charnay, C.; et al. 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] [PubMed]
  77. Tzankova, V.; Aluani, D.; Kondeva-Burdina, M.; Yordanov, Y.; Odzhakov, F.; Apostolov, A.; Yoncheva, K. Hepatoprotective and antioxidant activity of quercetin loaded chitosan/alginate particles in vitro and in vivo in a model of paracetamol-induced toxicity. Biomed. Pharmacother. 2017, 92, 569–579. [Google Scholar] [CrossRef]
  78. Trombino, S.; Cassano, R.; Ferrarelli, T.; Barone, E.; Picci, N.; Mancuso, C. Trans-ferulic acid-based solid lipid nanoparticles and their antioxidant effect in rat brain microsomes. Colloids Surf. B Biointerfaces 2013, 109, 273–279. [Google Scholar] [CrossRef]
  79. 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] [PubMed]
  80. Hans, M.; Lowman, A. Biodegradable nanoparticles for drug delivery and targeting. Curr. Opin. Solid State Mater. Sci. 2002, 6, 319–327. [Google Scholar] [CrossRef]
  81. Vila, A.; Sanchez, A.; Tobıo, M.; Calvo, P.; Alonso, M. Design of biodegradable particles for protein delivery. J. Control. Release 2002, 78, 15–24. [Google Scholar] [CrossRef]
  82. 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]
  83. 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] [PubMed]
  84. Patil, S.P.; Kumbhar, S.T. Antioxidant, antibacterial and cytotoxic potential of silver nanoparticles synthesized using terpenes rich extract of Lantana camara L. leaves. Biochem. Biophys. Rep. 2017, 10, 76–81. [Google Scholar]
  85. Saratale, R.G.; Benelli, G.; Kumar, G.; Kim, D.S.; Saratale, G.D. Bio-fabrication of silver nanoparticles using the leaf extract of an ancient herbal medicine, dandelion (Taraxacum officinale), evaluation of their antioxidant, anticancer potential, and antimicrobial activity against phytopathogens. Environ. Sci. Pollut. Res. 2018, 25, 10392–10406. [Google Scholar] [CrossRef] [PubMed]
  86. Phull, A.R.; Abbas, Q.; Ali, A.; Raza, H.; Kim, S.J.; Zia, M.; Haq, I.U. Antioxidant, cytotoxic and antimicrobial activities of green synthesized silver nanoparticles from crude extract of Bergenia ciliata. Future J. Pharm. Sci. 2016, 2, 31–36. [Google Scholar] [CrossRef]
  87. 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]
  88. Kalaiyarasan, T.; Bharti, V.K.; Chaurasia, O.P. 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]
  89. Sharma, B.; Deswal, R. Single pot synthesized gold nanoparticles using Hippophae rhamnoides leaf and berry extract showed shape-dependent differential nanobiotechnological applications. Artif. Cells Nanomed. Biotechnol. 2018, 46, 408–418. [Google Scholar] [CrossRef] [Green Version]
  90. 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]
  91. Sathishkumar, G.; Jha, P.K.; Vignesh, V.; Rajkuberan, C.; Jeyaraj, M.; Selvakumar, M.; Jha, R.; Sivaramakrishnan, S. Cannonball fruit (Couroupita guianensis, Aubl.) extract mediated synthesis of gold nanoparticles and evaluation of its antioxidant activity. J. Mol. Liq. 2016, 215, 229–236. [Google Scholar]
  92. Patra, J.K.; Das, G.; Baek, K.H. Phyto-mediated biosynthesis of silver nanoparticles using the rind extract of watermelon (Citrullus lanatus) under photo-catalyzed condition and investigation of its antibacterial, anticandidal and antioxidant efficacy. J. Photochem. Photobiol. B 2016, 161, 200–210. [Google Scholar] [CrossRef]
  93. Mohanta, Y.K.; Panda, S.K.; Jayabalan, R.; Sharma, N.; Bastia, A.K.; Mohanta, T.K. Antimicrobial, antioxidant and cytotoxic activity of silver nanoparticles synthesized by leaf extract of Erythrina suberosa (Roxb.). Front. Mol. Biosci. 2017, 4, 14. [Google Scholar] [CrossRef] [Green Version]
  94. Hamelian, M.; Zangeneh, M.M.; Amisama, A.; Varmira, K.; Veisi, H. Green synthesis of silver nanoparticles using Thymus kotschyanus extract and evaluation of their antioxidant, antibacterial and cytotoxic effects. Appl. Organomet. Chem. 2018, 32, e4458. [Google Scholar] [CrossRef]
  95. Chandra, H.; Patel, D.; Kumari, P.; Jangwan, J.S.; Yadav, S. Phyto-mediated synthesis of zinc oxide nanoparticles of Berberis aristata: Characterization, antioxidant activity and antibacterial activity with special reference to urinary tract pathogens. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 102, 212–220. [Google Scholar] [CrossRef]
  96. Veena, S.; Devasena, T.; Sathak, S.S.M.; Yasasve, M.; Vishal, L.A. Green synthesis of gold nanoparticles from Vitex negundo leaf extract: Characterization and in vitro evaluation of antioxidant-antibacterial activity. J. Clust. Sci. 2019, 30, 1591–1597. [Google Scholar] [CrossRef]
  97. Keshari, A.K.; Srivastava, R.; Singh, P.; Yadav, V.B.; Nath, G. Antioxidant and antibacterial activity of silver nanoparticles synthesized by Cestrum nocturnum. J. Ayurveda Integr. Med. 2020, 11, 37–44. [Google Scholar] [CrossRef] [PubMed]
  98. Subbaiy, R.; Selvam, M.M. Green synthesis of copper nanoparticles from Hibicus rosasinensis and their antimicrobial, antioxidant activities. Res. J. Pharm. Biol. Chem. Sci. 2015, 6, 1183–1190. [Google Scholar]
  99. Ghosh, S.; More, P.; Nitnavare, R.; Jagtap, S.; Chippalkatti, R.; Derle, A.; Kitture, R.; Asok, A.; Kale, S.; Singh, S.; et al. Antidiabetic and antioxidant properties of copper nanoparticles synthesized by medicinal plant Dioscorea bulbifera. J. Nanomed. Nanotechnol. 2015, S6, 007. [Google Scholar]
  100. Sarkar, J.; Chakraborty, N.; Chatterjee, A.; Bhattacharjee, A.; Dasgupta, D.; Acharya, K. Green synthesized copper oxide nanoparticles ameliorate defence and antioxidant enzymes in Lens culinaris. Nanomaterials 2020, 10, 312. [Google Scholar] [CrossRef] [Green Version]
  101. Dobrucka, R. Antioxidant and catalytic activity of biosynthesized CuO nanoparticles using extract of Galeopsidis herba. J. Inorg. Organomet. Polym. Mater. 2018, 28, 812–819. [Google Scholar] [CrossRef] [Green Version]
  102. Rajeshkumar, S.; Menon, S.; Kumar, S.V.; 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. 2019, 197, 111531. [Google Scholar] [CrossRef]
  103. Harshiny, M.; Iswarya, C.N.; Matheswaran, M. Biogenic synthesis of iron nanoparticles using Amaranthus dubius leaves extract as reducing agents. Powder Technol. 2015, 286, 744–749. [Google Scholar] [CrossRef]
  104. Muthukumar, H.; Manickam, 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]
  105. 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]
  106. 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] [PubMed] [Green Version]
  107. Markus, J.; Mathiyalagan, R.; Kim, Y.J.; Abbai, R.; Singh, P.; Ahn, S.; Perez, Z.E.J.; Hurh, J.; Yang, D.C. Intracellularsynthesis of goldnanoparticles with antioxidantactivity by probiotic Lactobacillus kimchicus DCY51T isolated from Koreankimchi. Enzym. Microb. Technol. 2016, 95, 85–93. [Google Scholar] [CrossRef] [PubMed]
  108. Baygar, T.; Ugur, A. Biosynthesis of silver nanoparticles by Streptomyces griseorubens isolated from soil and their antioxidant activity. IET Nanobiotechnol. 2017, 11, 286–291. [Google Scholar] [CrossRef]
  109. Oladipo, I.C.; Lateef, A.; Elegbede, J.A.; Azeez, M.A.; Asafa, T.M.; 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 2017, 173, 250–257. [Google Scholar] [CrossRef]
  110. Veeraapandian, S.; Sawant, S.N.; Doble, M. Antibacterial and antioxidant activity of protein capped silver and gold nanoparticles synthesized with Escherichia coli. J. Bimed. Nanotechnol. 2012, 8, 140–148. [Google Scholar] [CrossRef]
  111. 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]
  112. Ramya, S.; Shanmugasundaram, T.; Balagurunathan, R. Biomedical potential of actinobacterially synthesised 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]
  113. Torres, S.K.; Campos, V.L.; León, C.G.; Rodríguez-Llamazares, S.M.; Rajos, S.M.; González, M.; Smith, C.; Mondaca, M.A. Biosynthesis of selenium nanoparticles by Pantoea agglomerans and their antioxidant activity. J. Nanopart Res. 2012, 14, 1236. [Google Scholar] [CrossRef]
  114. 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]
  115. Sivasankar, P.; Seedevi, P.; Poongodi, S.; Sivakumar, M.; Murugan, T.; Sivakumar, L.; Sivakumar, K. 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]
  116. 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] [PubMed] [Green Version]
  117. 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, 5683–5696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Manjunath, H.M.; 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]
  119. Manjunath, H.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]
  120. 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. 3Biotech 2016, 6, 132. [Google Scholar] [CrossRef] [Green Version]
  121. Saravanakumar, K.; Wang, M.H. Trichoderma based synthesis of anti-pathogenic silver nanoparticles and their characterization, antioxidant and cytotoxicity properties. Microb. Pathog. 2018, 114, 269–273. [Google Scholar] [CrossRef]
  122. Gao, Y.; Anand, M.A.V.; Ramachandran, V.; Karthikkumar, V.; Shalini, V.; Vijayalakshmi, S.; Ernest, D. Biofabrication of zinc oxide nanoparticles from Aspergillus niger, their antioxidant, antimicrobial and anticancer activity. J. Clust. Sci. 2019, 30, 937–946. [Google Scholar] [CrossRef]
  123. Govindappa, M.; Farheen, H.; Chandrappa, C.P.; Rai, R.V.; Raghavendra, V.B. Mycosynthesis of silver nanoparticles using extract of endophytic fungi, Penicillium species of Glycosmis mauritiana, and its antioxidant, antimicrobial, anti-inflammatory and tyrokinase inhibitory activity. Adv. Nat. Sci. Nanosci. Nanotechnol. 2016, 7, 035014. [Google Scholar] [CrossRef]
  124. Nagajyothi, P.C.; Sreekanth, T.V.M.; Lee, J.I.; Lee, K.D. Mycosynthesis: Antibacterial, antioxidant and antiproliferative activities of silver nanoparticles synthesized from Inonotus obliquus (Chaga mushroom) extract. J. Photochem. Photobiol. B 2014, 130, 299–304. [Google Scholar] [CrossRef]
  125. Popli, D.; Anil, V.; Subramanyam, A.B.; Namratha, M.N.; Ranjitha, V.R.; Rao, S.N.; Rai, R.V.; Govindappa, M. Endophyte fungi, Cladosporium species-mediated synthesis of silver nanoparticles possessing in vitro antioxidant, anti-diabetic and anti-Alzheimer activity. Artif. Cells Nanomed. Biotechnol. 2018, 46, 676–683. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. 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]
  127. 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]
  128. 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. Bimed. Anal. 2020, 178, 112970. [Google Scholar] [CrossRef]
  129. 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]
  130. 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]
  131. 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]
  132. Dasari, S.; Suresh, K.A.; Rajesh, M.; Reddy, C.S.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]
  133. Debnath, R.; Purkayastha, D.D.; Hazra, S.; Ghosh, N.N.; Bhattacharjee, C.R.; Rout, J. Biogenic synthesis of antioxidant, shape selective gold nanomaterials mediated by high altitude lichens. Mater. Lett. 2016, 169, 58–61. [Google Scholar] [CrossRef]
  134. Kumar, H.; Bhardwaj, K.; Kuča, K.; Kalia, A.; Nepovimova, E.; Verma, R.; Kumar, D. Flower-basedgreensynthesis of metallicnanoparticles: Applicationsbeyondfragrance. Nanomaterials 2020, 10, 766. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Guajardo-Pachecoa, M.J.; Morales-Sanchz, J.E.; González-Hernándezc, J.; Ruiz, F. Synthesis of copper nanoparticles using soybeans as a chelant agent. Mater. Lett. 2010, 64, 1361–1364. [Google Scholar] [CrossRef]
  136. 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]
  137. He, Y. A novel solid-stabilized emulsion approach to CuO nanostructures microspheres. Mater. Res. Bull. 2007, 42, 190–195. [Google Scholar] [CrossRef]
  138. Motogoshi, R.; Oku, T.; Suzuki, A.; Kikuchi, K.; Kikuchi, S.; Jeyadevan, B.; Cuya, J. Fabrication and characterization of cupprious oxide: Fullerene solar cells. Synth. Met. 2010, 160, 1219–1222. [Google Scholar] [CrossRef]
  139. Herlekar, M.; Barve, S.; Kumar, R. Plant-mediated green synthesis of iron nanoparticles. J. Nanopart. Res. 2014, 2014, 140614. [Google Scholar] [CrossRef] [Green Version]
  140. Huber, D.L. Synthesis, properties, and applications of iron nanoparticles. Small 2005, 1, 482–501. [Google Scholar] [CrossRef]
  141. 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, 63–73. [Google Scholar] [CrossRef]
  142. 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]
  143. 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] [PubMed]
  144. 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]
  145. 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] [PubMed] [Green Version]
  146. 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]
  147. Agarwal, H.; Kumar, S.V.; Rajeshkumar, S. A review on green synthesis of zinc oxide nanoparticles -An eco-friendly approach. Res. Effic. Technol. 2017, 3, 406–413. [Google Scholar] [CrossRef]
  148. Jayaseelan, C.; Rahuman, A.A.; Kirthi, A.V.; Marimuthu, S.; Santhoshkumar, T.; Bagavan, A.; Guarav, K.; Karthik, L.; Rao, K.V. Novel microbial route to synthesize ZnO nanoparticles using Aeromonas hydrophila and their activity against pathogenic bacteria and fungi. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2012, 90, 78–84. [Google Scholar] [CrossRef]
  149. Pulit-prociak, J.; Chwastowski, J.; Kucharski, A.; Banach, M. Applied surface science functionalization of textiles with silver and zinc oxide nanoparticles. Appl. Surf. Sci. 2016, 385, 543–553. [Google Scholar] [CrossRef]
  150. Wodka, D.; Bielaníska, E.; Socha, R.P.; Elzbieciak-Wodka, M.; Gurgul, J.; Nowak, P.; Warszyński, P.; Kumakiri, I. Photocatalytic activity of titanium dioxide modified by silver nanoparticles. ACS Appl. Mater. Interfaces 2010, 2, 1945–1953. [Google Scholar] [CrossRef]
  151. Wadhwani, S.A.; Shedbalkar, U.U.; Singh, R.; Chopade, B.A. Biogenic selenium nanoparticles: Current status and future prospects. Appl. Microbiol. Biotechnol. 2016, 100, 2555–2566. [Google Scholar] [CrossRef]
  152. 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. 2013, 1, 6365–6372. [Google Scholar] [CrossRef]
  153. Din, M.I.; Rani, A. Recent advances in the synthesis and stabilization of nickel and nickel oxide nanoparticles: A green adeptness. Int. J. Anal. Chem. 2016, 2016, 3512145. [Google Scholar]
  154. Saxena, A.; Kumar, K.; Mozumdar, S. Ni-nanoparticles: An efficient green catalyst for chemo-selective oxidative couplingof thiols. J. Mol. Catal. A Chem. 2007, 269, 35–40. [Google Scholar] [CrossRef]
  155. Alonso, F.; Riente, P.; Yus, M. Hydrogen-transfer reduction of carbonyl compounds promoted by nickel nanoparticles. Tetrahedron 2008, 64, 1847–1852. [Google Scholar] [CrossRef]
  156. Dhakshinamoorthy, A.; Pitchumani, K. Clay entrapped nickel nanoparticles as efficient and recyclable catalysts forhydrogenation of olefins. Tetrahedron Lett. 2008, 49, 1818–1823. [Google Scholar] [CrossRef]
  157. Alonso, F.; Riente, P.; Yus, M. Wittig-type olefination of alcohols promoted by nickel nanoparticles: Synthesis ofpolymethoxylated and polyhydroxylated stilbenes. Eur. J. Org. Chem. 2009, 2009, 6034–6042. [Google Scholar] [CrossRef] [Green Version]
  158. Alonso, F.; Riente, P.; Yus, M. Alcohols for the α-alkylationof methyl ketones and indirect aza-wittig reaction promoted bynickel nanoparticles. Eur. J. Org. Chem. 2008, 2008, 4908–4914. [Google Scholar] [CrossRef]
  159. Li, X.K.; Ji, W.J.; Zhao, J.; Wang, S.J.; Au, C.T. Ammonia decomposition over Ru and Ni catalysts supported on fumed SiO2, MCM-41, and SBA-15. J. Catal. 2005, 236, 181–189. [Google Scholar] [CrossRef]
  160. Li, Y.; Zhang, B.; Xie, X.; Liu, J.; Xu, Y.; Shen, W. Novel Nicatalysts for methane decomposition to hydrogen and carbonnanofibers. J. Catal. 2006, 238, 412–424. [Google Scholar] [CrossRef]
  161. Al-Rawi, M.; Diabaté, S.; Weiss, C. Uptake and intracellular localization of submicron and nano-sized SiO₂ particles in HeLa cells. Arch Toxicol. 2011, 85, 813–826. [Google Scholar] [CrossRef]
  162. Hussain, S.M.; Hess, K.L.; Gearhart, J.M.; Geiss, K.T.; Schlager, J.J. In vitro toxicity of nanoparticles in BRL 3A rat liver cells. Toxicol. Vitr. 2005, 19, 975–983. [Google Scholar] [CrossRef]
  163. Clift, M.J.D.; Rothen-Rutishauser, B.; Brown, D.M.; Duffin, R.; Ronaldson, K.; Proudfoot, L.; Guy, K.; Stone, V. The impact of different nanoparticle surface chemistry and size on uptake and toxicity in a murine macrophage cell line. Toxicol. Appl. Pharmacol. 2008, 232, 418–427. [Google Scholar] [CrossRef] [PubMed]
  164. Rabolli, V.; Thomassen, L.C.; Uwambayinema, F.; Martens, J.A.; Lison, D. The cytotoxic activity of amorphous silica nanoparticles is mainly influenced by surface area and not by aggregation. Toxicol. Lett. 2011, 206, 197–203. [Google Scholar] [CrossRef] [PubMed]
  165. Morais, T.; Soares, M.E.; Duarte, J.A.; Soares, L.; Maia, S.; Gomes, P.; Pereira, E.; Fraga, S.; Carmo, H.; De Lourdes Bastos, M. Effect of surface coating on the biodistribution profile of gold nanoparticles in the rat. Eur. J. Pharm. Biopharm. 2012, 80, 185–193. [Google Scholar] [CrossRef] [Green Version]
  166. Cho, W.S.; Cho, M.; Jeong, J.; Choi, M.; Cho, H.Y.; Han, B.S.; Kim, S.H.; Kim, H.O.; Lim, Y.T.; Chung, B.H.; et al. Acute toxicity and pharmacokinetics of 13 nm-sized PEG-coated gold nanoparticles. Toxicol. Appl. Pharmacol. 2009, 236, 16–24. [Google Scholar] [CrossRef]
  167. Knaapen, A.M.; Borm, P.J.; Albrecht, C.; Schins, R.P. Inhaled particles and lung cancer. Part A: Mechanisms. Int. J. Cancer 2004, 109, 799–809. [Google Scholar] [CrossRef] [PubMed]
  168. Xia, T.; Kovochich, M.; Liong, M.; Mädler, L.; Gilbert, B.; Shi, H.; Yeh, J.I.; Zink, J.I.; Nel, A.E. Comparison of the mechanism of toxicity of zinc oxide and cerium oxide nanoparticles based on dissolution and oxidative stress properties. ACS Nano 2008, 2, 2121–2134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Side effects of oxidative stress on human body. COPD—Chronic obstructive pulmonary disease; ARDS—Acute respiratory distress syndrome; OCD—Obsessive-compulsive disorder; ADHD—Attention-deficit/hyperactivity disorder.
Figure 1. Side effects of oxidative stress on human body. COPD—Chronic obstructive pulmonary disease; ARDS—Acute respiratory distress syndrome; OCD—Obsessive-compulsive disorder; ADHD—Attention-deficit/hyperactivity disorder.
Nanomaterials 10 01334 g001
Table 1. Antioxidants from different biological sources.
Table 1. Antioxidants from different biological sources.
SourceAntioxidantsRef.
BacteriaThiazostatins A, 5-(2,4-Dimethylbenzyl) Pyrrolidin-2-One, Phenazoviridin, Benthophoenin, Benthocyanins A, B and C, Benzastatins C, Benzastatins A, (Z)-1-((1-Hydroxypenta-2,4-Dien-1-Yl)Oxy)Anthracene-9,10-Dione[53]
PlantsGallic acid, Protocatechuic acid, p-Coumaric acid, Caffeic acid, Rosmarinic acid, Carnosol, Carnosic acid, Rosmanol, Rutin, Epicatechin gallate, Epigallocatechin gallate, Epicatechin, Quercetin (flavanol), Eugenol, Carvacrol, Safrole, Thymol, Myristicin, Menthol, 1,8-Cineol, α-Terpineol, p-Cymene, Cinnamaldehyde, Piperine, Flavone, Flavonol, Chalcone, Flavanone, Anthocyanin, Anthocyanidin-3,5-glucoside, Alpha tocopherol, Gamma tocopherol, Ascorbic acid, Ascorbyl palmitate, Propyl gallate, Resveratrol[55]
FungiIsopestacin, Pestacin, Atrovenetin, 2-Acetonyl-2,4,9,-Trihydroxy-6-Methoxy-7-Methyl-1HPhenalene-1,3(2H)-Dione, Graphislactone, 4,6-dihydroxy-5-methoxy-7-methyl-1,3-dihydroisobenzofuran, 4,5,6-trihydroxy-7-methyl-1,3-dihydroisobenzofuran, 4,6-dihydroxy-5-methoxy-7-methylphthalide, Kojic acid, Phomapyrone C, Versicolone A, Terremutin, Terreic Acid, Neoechinulin A, Candidusin B, 3″-Dihydroxyterphenyllin and 3-Hydroxyterphenyllin, p-Hydroxybenzoic acid, Protocatechuic acid, Gallic acid, Gentisic acid, Vanillic acid, 5-Sulfosalicylic acid, Syringic acid¸ Veratric acid, Vanillin, Cinnamic acid, p-Coumaric acid, o-Coumaric acid, Caffeic acid, Ferulic acid, 3-O-Caffeoylquinic acid, 4-O-Caffeoylquinic acid, 5-O-Caffeoylquinic acid, Quercetin, Rutin, Kaempferol, Myricetin, Chrysin, Catechin, Hesperetin, Naringenin, Naringin, Formometin, Biochanin, Pyrogallol, Resveratrol, Ellagic acid, Tannic acid, Sinapic acid, Flavonols, Flavones, Isoflavones, Flavanones, Anthocyanidins, Flavanols, Vitamin C, Vitamin E, Homogentisic acid[53,56,57]
Algaeβ-carotene, Lutein, Bromophenol, Carrageenan, Fucophlorethols, Fucoxanthin, Galactan sulphate, Phlorotannins, Phycoerythrin, Porphyran, Shinorine, Catechin, Epicatechin, Gallate, Alginic acid, Laminaran, Vitamin A, Phloroglucinol, Eckol, Fucodiphlorethol G, Phlorofucofuroeckol A, 7-phloroeckol, Dieckol, 6,6′-bieckol, Triphlorethol-A, 2,7′-phloroglucinol-6,6′-bieckol[58,59]
Lichens 1-Chloropannarin, 2-O-Methylsekikaic acid, Atranol, Atranorin, Barbatic acid, Boninic acid, Chloroatranol, Chloroatranorin, Chlorohematommic acid, Cryptostictinolide, Divaricatic acid, Ergosterol peroxide, Ethyl chlorohematommate, Evernic acid, Fumarprotocetraric acid, Gyrophoric acid, Hematommic acid, Lecanoric acid, Lecanoric acid, Methyl orsellinate, Orcinol, Physodic acid, Protrocetraric acid, Sekikaic acid, Umbilicaric acid [60]
Table 2. In vivo and in vitro delivery of antioxidants by different nanoparticles.
Table 2. In vivo and in vitro delivery of antioxidants by different nanoparticles.
NanoparticlesDelivered Antioxidant/EnzymesMethod of PreparationCharacterizationSizeTest SystemBiological EffectsRef.
Poly(lactide-co glycolide) (PLGA) and Polycaprolactone (PCL) Ellagic acidEmulsion -diffusion-evaporationDLS, Zeta potentialNDOvernight fasted male Sprague Dawley (SD) ratsPrevent cyclosporine A (CyA)-induced nephrotoxicity[71]
Polybutylcyanoacrylate, Liposomes, Poly(Lactide-co-Glycolide)Superoxide dismutase (SOD)Emulsion solvent evaporationDLS, Zeta potentialNDC57BL/6 miceNanoparticles displayed protection against reperfusion injury and ischemia when applied after injury reduced in infarct volume with a 50% to 60%, lowered inflammatory markers, and improved in mice behavior[72]
Iron oxide (magnetite)Catalase and SODNanoprecipitationZeta potential, TEM303 ± 38 nm (Catalase loaded), 350 ± 10 nm (SOD loaded)Bovine aortic endothelial cells (BAEC), Primary human umbilical vein endothelial cells (HUVEC)Cultured endothelial cells rapidly take magnetically responsive nanoparticles (MNP) under magnetic guidance catalase-loaded providing increased resistance to oxidative stress (62 ± 12% cells rescued from hydrogen peroxide induced cell death vs. 10 ± 4% under non-magnetic conditions)[73]
Poly(lactide-co glycolide) (PLGA)SODEmulsion solvent evaporationTEM, DLS, Zeta potential81 ± 4 nmMale Sprague-Dawley ratsNPs encapsulated by superoxide dismutase helps in reduction of cerebral injury and promote neurological recovery in a rat cerebral ischemia-reperfusion model[74]
Bovine serum albumin (BSA)-dextranCurcuminSelf-assemblyTEM, DLS, Zeta potential115 nmCaco-2 cellsAt 5 μg/mL curcumin in BSA-dextran nanoparticle the CAA (cellular antioxidant activity) value was 65.35, significantly more that of free curcumin (48.61) at the same concentration, showing that nanoencapsulation increased the uptake of curucmin (p < 0.05). Curucmin-loaded BSA-dextran nanoparticle EC50 values was 3.27 μg/mL, indicating the CAA of curcumin was enhanced by nanoparticle-based delivery systems[75]
Mesoporous silicaCaffeic acid, RutinNDTEM, Zeta potential200 nmCaco-2 and the epidermal HaCaT cell linesAfter 24 h incubation of cells with grafted nanoparticles the best results were given by Rutin in terms of antioxidant capacities preservation during coupling procedures, decrease of ROS level and cellular toxicity alleviation. Rutin protective effects were found more apparated in HaCaT than in Caco-2 cells, revealing much cellular specificity towards defense against oxidative stress; MSN-RUT has ability to stimulate a strong Nrf2 protective response in HaCaT cells, accompanied by a comparable induction of HO-1 mRNA. These responses level in Caco-2 cells was again less important.[76]
Chitosan/alginateQuercetinGelationZeta potentialNDHuman hepatocellular carcinoma HepG2 cells and Male Wistar rats (paracetamol-induced liver injury)Pretreatment of HepG2 cells with (10 µg/mL) encapsulated quercetin significantly reduction in cell viability in H2O2-induced oxidative stress (0.1 mM H2O2), thus showing an efficacious in vitro protection; oral pretreatment with encapsulated quercetin (0.18 mg/kg b.w., 7 days) significantly reduced the increased serum transaminases ALT and AST levels, reduced the lipid peroxidation and restored the gluthation (a marker of cell antioxidant defence system) levels[77]
Stearic acid- and stearyl ferulate-based solid lipidTrans-ferulic acidMicroemulsionDLS, TEM505 ± 8.2 (SLN-FA), 600 ± 3.4 nm (SLN-SF-FA)Male ratsBoth SLN-SF-FA and SLN-FA dose-dependently reduced lipid peroxidation induced by the three oxidants (NADPH/ADP-Fe3+, AAPH and SIN-1). SLN-SF-FA showed high efficiency (EC50) and potency (maximal activity) against NADPH/ADPFe3+ and AAPH-induced lipid peroxidation[78]
TEM—transmission electron microscopy; DLS—dynamic light scattering; ND—not defined.
Table 3. Antioxidant potential of functionalized nanoparticles.
Table 3. Antioxidant potential of functionalized nanoparticles.
Antioxidant SourceTypes of NanoparticlesBiological ExtractTemperatureReaction TimeCharacterizationsMorphologySizeStabilityAntioxidant ActivityRef.
PlantSilverLantana camara L. leaves extract (terpenes rich)RT24 hUV-Vis, XRD, Zeta potential; SEMSphere425 nmNdA10 μL of AgNP (2 mg/mL), spot intensity was found good and comparable with ascorbic acid[84]
SilverTaraxacum officinale leaf extractRT15 minUV-Vis, XRD, FTIR, HRTEMSphere15 nm4 monthsThe efficiency of AgNPs were found against ABTS radicals, displayed an IC50 value of 45.6 μg/mL; scavenging potential of Nitric Oxide is 72.1% at 100 μg/mL concentration with IC50 value of 55.2 μg/mL[85]
SilverBergenia ciliate crude extractRT3 hUV-Vis, SEM, FTIR,Sphere35 nmNdResults of DPPH activity showed the effective free radical % scavenging potential of Bergenia ciliate AgNPs is 59.31%[86]
SilverClerodendrum phlomidis L. leaves extractRT10 minUV-Vis, SEM, TEM, EDAX, FT-IRSphere23–42 nmNdAgNPs exhibited remarkable antioxidant activity than the crude extract using phosphomolybdate assay, ferric reducing power, superoxide radical scavenging activity and DPPH assay[87]
SilverHippophae rhamnoides L. leaves extractRT24 hUV-Vis, TEM, HRTEM, FTIRSphere10–40 nm1 yearThe SBT@AgNPs showed excellent DPPH radical scavenging capacity. The results also revealed that the antioxidant properties of the samples depends on dose as their concentrations (5–25 µg mL−1) increase their percentage DPPH radical scavenging abilities also increased [88]
GoldHippophae rhamnoides ssp. Turkestanica leaves and berries extractRT2 min (LE), 15 min (BE)UV-Vis, HRTEM, FTIR, XRDTriangles, hexagon and sphere (BE AuNPs), Sphere (LE AuNPs)55 nm (BE AuNPs), 27 nm (LE AuNPs)5 monthsColorimetric DPPH assay at (80 µg/mL) concentration showed, free radical scavenging activity was found maximum in LE AuNPs (81%) and BE (70%) AuNPs. LE AuNPs nanospheres (IC50 49 µg) revealed a little better (14%) antioxidant capacity as compared to BE nanotriangles (IC50 57 µg)[89]
SilverMorus alba leaf extractRT10 minUV-Vis, FTIR, SEM, FESEM, EDX, HRTEM, XRD, DLSSphere12–39 nmNdDose dependent antioxidant activity against free radicals like DPPH, ABTS+, superoxide and nitric oxide[90]
GoldCouroupita guianensis Aubl. fruit extract70 °C60 minUV-Vis, FTIR, TEM, XRD, DLS, Zeta potentialCubic26 nm45 daysFor DPPH assay, CGAuNPs IC50 was 37 μg/mL; CGAuNPs were potent in scavenging the hydroxyl radicals with IC50 values of 30 and 36 μg/mL respectively; CGAuNPs superoxide scavenging activity increased with increasing concentrations and was observed as 89.8% inhibition rate[91]
SilverCitrullus lanatus rind extractRT24 hUV-Vis, SEM, EDX, FTIR, XRDSphere109.97 nmNdAgNPs DPPH free radical scavenging activity at 20–100 μg/mL ranged from 21.65% to 60.97%; AgNPs ABTS radical scavenging activity was 11.25% to 55.26% at a concentration of 20–100 μg/mL; AgNPs Nitric oxide scavenging activity was 9.05% to 54.15% at a concentration of 20–100 μg/mL[92]
SilverErythrina suberosa (Roxb.) leaf extractRTOver nightUV-Vis, ATR-FTIR, DLS, TEM,Sphere12–115 nmNdAgNPs antioxidant potential was estimated by DPPH radical scavenging assay having IC50 30.04 μg/mL[93]
SilverThymus kotschyanus extractRT30 minUV-Vis, FTIR, XRD, EDS, SEM, AFM, HRTEMSphere50–60 nmNdAgNPs DPPH free radical scavenging activities demonstrate effective inhibition as compared to BHT as the standard antioxidant[94]
Zinc OxideBerberis aristata leaf extract70 °CNDUV-Vis, XRD, FTIR, SEM, EDX, DLSNeedle90–110 nmNdB. aristata leaves extract ZnO nanoparticles showed percent inhibition of 32.06% at the concentration of 1 μg/mL and for 5 μg/mL it was to be 61.63%[95]
GoldVitex negundo leaf extractRTNDUV-Vis, XRD, FTIR, TEMSphere20–70 nmNdRadical scavenging activity of DPPH shown that at a 120 µg/mL concentration, the scavenging activity NPs reached 84.64% and the IC50 of the NPs was found to be 62.18 µg; The nitric oxide assay results revealed that the antioxidant property of NPs at a concentration of 120 µg/mL, the NPs scavenging activity reached 69.79% with IC50 estimated at 70.45 µg[96]
SilverCestrum nocturnum leaf extractRT1 weekXRD, TEM, EDS, SEM, FTIR,Sphere20 nmNdAgNPs antioxidant activity for DPPH method was 29.55%[97]
Copper OxideHibicus rosasinensis leaf extractRT48 hUV-Vis, FTIR, TEMNDNDNdGood antioxidant activity from FRAP assay[98]
Copper OxideDioscorea bulbifera tuber extract40 °C5 hUV-Vis, TEM, EDS, XRD, DLSSphere86–126 nmNdShowed 40.81 ± 1.44%, 79.06 ± 1.02% and 48.39 ± 1.46% scavenging activity against DPPH, nitric oxide and superoxide radicals respectively[99]
Copper OxideAdiantum lunulatum whole plant extractRT1 hUV-Vis, DLS, TEM, EDX, XRD, FTIRSphere6.5 nmNdCAT, APX, and SOD activities have steadily increased according to the increasing concentration of copper nanoparticles treatment to Lens culinaris[100]
Copper OxideGaleopsidis herba. G. herba extract25 °C24 hUV-Vis, SEM, FTIR, TEMSphere10 nmNdShowed good scavenging activity against DPPH[101]
Copper OxideCissus arnotiana leaf extractRT4 hUV-Vis, XRD, SEM, TEMSphere80–90 nmNdThe antioxidant property observed was comparatively equal with the standard antioxidant agent ascorbic acid at a maximum concentration of 40 μg/mL DPPH assay[102]
IronAmaranthus dubius leaf extract60 °C90 minUV-Vis, FTIR, XRD, SEMSphere43–220 nmNdShowed high antioxidant activity against DPPH[103]
IronAmaranthus spinosus leaf extractRT90 minUV-Vis, FTIR, TEM, EDX, XRDSphereNDNdAntioxidant efficiency was observed to be 93% against DPPH[104]
IronAsphodelus aestivus Brot. extract50–60 °C20 minUV-Vis, FTIR, TEM, EDX, SEM, XRDNS20–25 nmNdAntioxidant activity against DPPH (IC50: 3.48 µg/mL) and ABTS (60.52%)[105]
Nickel OxideStevia rebaudiana Bertoni leaf extract100 °C2 hUV-Vis, XRD, SEM, TEM, FTIRSphere20–50 nmNdAntioxidant efficiency was observed to be 70% against DPPH[106]
BacteriaGoldLactobacillus kimchicus DCY51T biomassRT12 hUV-Vis, FE-TEM, XRD, DLS, FTIRSphere13 nmNSLowest concentration of the biosynthesized AuNps DPPH percentage scavenging ability was 15.85 ± 0.49 and when concentration was increased to 500 μg/mL this scavenging ability was increased to 60 ± 1.82[107]
SilverStreptomyces griseorubens AU2 cell free supernatantRT48 hUV-Vis, FTIR, TEM, SEM, XRDSphere5–20 nmNdDPPH free radical scavenging activity of AgNPs showed at various concentrations viz. 9.66, 14.27, 15.59, 23.46 and 54.99%[108]
GoldEnterococcus species cell free extractRT30 minUV-Vis, TEM, EDX, FTIRSphere8–50 nmNdAuNPs have ability to scavenge DPPH at all the investigated concentrations (1–40 μg/mL), yielding activities of 33.24–51.47%[109]
Gold and SilverEscherichia coli cell proteinRTNDUV-Vis, XRD, FTIR, TEMTriangular, circular, hexagonal (AuNPs), Sphere (AgNPs)10–100 nm (AuNPs), 10–50 nm (AgNPs)3 monthsEC75 (for scavenging 75% effective concentration) of protein capped gold nanoparticles is 916 µg/mL[110]
SilverStreptomyces naganishii MA7 biomassRT72 hUV-Vis, FTIR, XRD, EDX, AFM, SEM, TEM, HRTEMSphere5–50 nmNdAt a concentration of 1000 μg/mL, AgNPs showed good reducing power comparatively than ascorbic acid (vitamin C)[111]
SeleniumStreptomyces minutiscleroticus M10A62 biomassRT72 hUV-Vis, XRD, HRTEM, FTIR, EDXSphere10–250 nmNdSeNPs actinobacterially synthesized were found strong free radical scavenging activity compared with standard ascorbic acid was proved by positive DPPH activity. Free radical scavenging activity depends on concentration as increases with increased concentration of SeNPs[112]
SeleniumPantoea agglomerans UC-32RT24 hTEM, EDS, SEMSphere100 nmNdHigh antioxidant activity in human umbilical vein
endothelial cells
[113]
Gold and SilverGordonia amicalis HS-11 cell free supernatant100 °C10 minUV-Vis, XRD, TEM, FTIRGrain5–25 nmNdCFS synthesized AgNPs and AuNPs respectively, showed 88.5 and 87.75% inhibition towards hydroxyl radicals; CFS mediated AuNPs inhibited nitric oxide radicals by 67.5%and with AgNPs the inhibited by 61.5%[114]
SilverStreptomyces violaceus MM72 exopolysaccharidesRT1 hTEM, EDX, XRDND30 nmNdThe DPPH radical scavenging activity shown by SNPs of 89.5% at 50 μg/mL concentration; SNPs exhibited the more total antioxidant activity of 0.730 at 50 μg/mL concentration; H2O2 scavenging activity of SNPs was evaluated at different concentrations, 50 μg/mL exhibited a higher activity of 72.5% which was significantly higher than that of the standard L-ascorbic acid; SNPs (50 μg/mL) had a nitric oxide scavenging activity of 60.1%; Ferric reducing power of the SNPs was estimated by the reduction of Fe/ferricyanide and the inhibition was observed at 0.390 AU at 50 μg/mL[115]
FungiZinc OxidePichia kudriavzevii cell free extractRT36 hUV-Vis, XRD, TEM, Zeta potentialHexagonal10–61 nmNdDPPH radical scavenging activities IC50 values were 10 ± 0.52, 5.26 ± 0.42 and 25.46 ± 0.35 µg/mL for ZnO/T1, ZnO/T2 and ZnO/T3 respectively[116]
SilverPestalotiopsis microspora filtrateRT24 hUV-Vis, FTIR, XRD, TEM, DLSSphere2–10 nmNdIC50concentrations (the concentration of the sample required to scavenge 50% radicals) of the biosynthesized AgNPs were found to be 76.95 ± 2.96 μg/mL; Biosynthesized AgNPs also exhibited effective scavenging activity against H2O2 radicals and the maximum scavenging activity of 51.14% ± 1.78% was observed at the highest concentration of 100 μg/mL[117]
GoldCladosporium cladosporioides filtrateRTNDUV-Vis, FESEM, EDX, XRD, FTIR, DLS, AFMCubic100 nm6 monthsDPPH radical scavenging capacity of the AuNPs was found to be dose dependent; AuNPs was subjected to reducing power assay where it showed moderate activity of 1.51 ± 0.03 mg of AAE/g sample[118]
SilverCladosporium cladosporioides filtrateRTNDUV-Vis, FESEM, XRD, FTIR, DLS, AFMSphere100 nmNdAgNPs showed potent antioxidant potential and their radical scavenging ability was increasing with increment in their concentration[119]
SilverAspergillus versicolor ENT7 filtrateRT24 hUV-Vis, TEM, XRD, FTIRSphere3–40 nmNdThe radical scavenging activity for the AgNPs at 100 µg/mL was determined as 60.04% which is close to 68.52% obtained for the standard ascorbic acid at the same concentration. IC50 value for the AgNPs is found to be 60.64 µg/mL[120]
SilverTrichoderma atroviride cell free filtrate40 °CNDUV-Vis, TEM, EDS, FTIRVariables15–25 nmNdAgNPs exhibited quite higher DPPH scavenging activity at concentration dependent manner with IC50 of 45.6 μg/mL[121]
Zinc OxideAspergillus niger cell free filtrateRT24 hUV-Vis, FTIR, XRD, DLS, SEMRod and cluster80–130 nmNdMaximum DPPH scavenging of 57.74% was obtained at 100 µg/mL concentration of ZnONPs; ABTS assay scavenging of 73.58% was obtained at 100 µg/mL concentration of ZnONPs[122]
SilverPenicillium species extractRT10 minUV-Vis, SEM, XRDSphere18 nmNdFRAP reducing ability potentially reduced by PsAgNPs was 1109.41 where the standard ascorbic acid shows 1648.52 μm[123]
SilverInonotus obliquus extractRT80 minUV-Vis, SEM, EDX, TEM, XRD, FTIRSphere14.2–35.2 nm8 monthsFree radical scavenging activity of the AgNPs on ABTS radicals was found to increase with increase in the concentration, showing maximum inhibition (76.57%) at 1 mM and minimum inhibition (60.98%) at 0.125 mM solution[124]
SilverCladosporium species extractRT1 hUV-Vis, SEM, XRD, FTIRSphere24 nmNdCsAgNPs exhibited potent antioxidant potential and as the concentration increases the radical scavenging ability also increases[125]
SilverGanoderma lucidium and Agaricus Bisporus extractRT and 60 °C12 h (RT), 5 h (60 °C)UV-Vis, SEM, XRD, FTIRRod10–80 nmNdEHT synthesized AgNPs shows higher antioxidant activity (75% ± 0.24%) in comparison of standard (70.34% ± 0.03%)[126]
GoldInonotus obliquus extractRT30 minUV-Vis, SEM, EDX, TEM, AFM, XRD, FTIRSphere, triangle, hexagonal and rod23 nmNdABTS scavenging effect increased with increasing concentration of AuNPs. The ABTS radical scavenging activity showed a maximum and minimum at 1 mM and 0.125 mM, respectively[127]
SilverGanoderma lucidium extractRTNDUV-Vis, TEM, XRD, FTIR, XPSSphere15–22 nmNdDPPH free radicals scavenging activity of AgNPs raised from 32.57% to 54.16% when concentration raised from 50 mg/L to 100 mg/L[128]
SilverGanoderma lucidium extract60 °C24 hUV-Vis, XRD, SEM, FTIRSphere10–30 nmNdPercentage of inhibition by silver nanoparticles showing
maximum with 73.49% and minimum with 55.34% at 100 and 10 μg/mL respectively
[129]
AlgaeGoldGracilaria corticata extract40 °C4 hUV-Vis, SEMND45–57 nmNdSynthesized AuNPs revealed a good capacity of DPPH free radical scavenging[130]
GoldLemanea fluviatilis (L.) C.Ag. extractRT12 hUV-Vis, XRD, TEM, HRTEM, DLS, FTIRSphere5–15 nmUp to 3 monthsDPPH scavenging (%) vs. different weights of the sample was found to be 18.10 mg[131]
SilverEcklonia cava extractRT72 hUV-Vis, FTIR, XRD, TEM, DLSSphere43 nmNd250 µg/mL of Ecklonia cava extract or biosynthesized AgNPs was mixed with DPPH solution, ca. 50% of scavenging activity was achieved. DPPH radical scavenging activities of Ecklonia cava extract and biosynthesized AgNPs were similar at the same concentrations (e.g., 100, 250, and 500 µg/mL)[59]
LichensSilverParmeliopsis ambigua, Punctelia subrudecta, Evernia mesomorpha, Xanthoparmelia plitti mycelia matsRT24 hUV-Vis, SEM, FTIRVariables150–250 nmNdSNPs samples total antioxidant capacity are 2.55 ± 0.05, 3.35 ± 0.04, 2.90 ± 0.01, 2.22 ± 0.01 µg AA/g respectively. Punctelia subrudecta displayed a higher activity (3.35 ± 0.04 µg AA/g) than the remaining samples; The hydroxyl radical scavenging activity showed that values were 7.75 ± 0.10, 34.48 ± 1.19, 31.97 ± 1.87 and 27.21 ± 1.39 µg/mL for all four samples respectively; Among the tested lichen SNPs, Punctelia subrudecta (Punctelia subrudecta) gave highest reducing power[132]
GoldAcroscyphus sphaerophoroides Lev, Sticta nylanderiana extractRT12 hUV-Vis, FTIR, XRD, TEMQuasi-spherical and prismatic (Acroscyphus sp.), Twinned (Sticta sp.)5–35 nm (Acroscyphus sp.), 20–50 nm (Sticta sp.)Up to 3 months1.66 and 4.48 mg sample concentration (SC50) were found[133]
RT–room temperature; Uiv-vis—ultraviolet–visible spectroscopy; TEM—transmission electron microscopy; SEM—scanning electron microscopy; FT-IR—Fourier-transform infrared spectroscopy; XRD—X-ray powder diffraction; EDX—energy dispersive X-ray spectroscopy; DLS—dynamic light scattering; HRTEM—high-resolution transmission electron microcopy; AFM—atomic force microscopy; ND—not defined; NS—not specified; Nd—not determ.

Share and Cite

MDPI and ACS Style

Kumar, H.; Bhardwaj, K.; Nepovimova, E.; Kuča, K.; Singh Dhanjal, D.; Bhardwaj, S.; Bhatia, S.K.; Verma, R.; Kumar, D. Antioxidant Functionalized Nanoparticles: A Combat against Oxidative Stress. Nanomaterials 2020, 10, 1334. https://doi.org/10.3390/nano10071334

AMA Style

Kumar H, Bhardwaj K, Nepovimova E, Kuča K, Singh Dhanjal D, Bhardwaj S, Bhatia SK, Verma R, Kumar D. Antioxidant Functionalized Nanoparticles: A Combat against Oxidative Stress. Nanomaterials. 2020; 10(7):1334. https://doi.org/10.3390/nano10071334

Chicago/Turabian Style

Kumar, Harsh, Kanchan Bhardwaj, Eugenie Nepovimova, Kamil Kuča, Daljeet Singh Dhanjal, Sonali Bhardwaj, Shashi Kant Bhatia, Rachna Verma, and Dinesh Kumar. 2020. "Antioxidant Functionalized Nanoparticles: A Combat against Oxidative Stress" Nanomaterials 10, no. 7: 1334. https://doi.org/10.3390/nano10071334

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