Next Article in Journal
The Effects of Titanium Dioxide Nanoparticles on Osteoblasts Mineralization: A Comparison between 2D and 3D Cell Culture Models
Next Article in Special Issue
Ternary Polyaniline@Bi2O3-BiOCl Nanocomposites as Innovative Highly Active Photocatalysts for the Removal of the Dye under Solar Light Irradiation
Previous Article in Journal
Irreversibility Marangoni Tri-Hybrid Nanoflow Analysis for Thermal Enhancement Applications
Previous Article in Special Issue
Sono-Photocatalytic Activity of Cloisite 30B/ZnO/Ag2O Nanocomposite for the Simultaneous Degradation of Crystal Violet and Methylene Blue Dyes in Aqueous Media
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Biogenic Selenium Nanoparticles in Biomedical Sciences: Properties, Current Trends, Novel Opportunities and Emerging Challenges in Theranostic Nanomedicine

by
Marjorie C. Zambonino
1,
Ernesto Mateo Quizhpe
1,
Lynda Mouheb
2,
Ashiqur Rahman
3,
Spiros N. Agathos
4 and
Si Amar Dahoumane
5,6,*
1
School of Biological Sciences and Engineering, Yachay Tech University, Hacienda San José s/n, San Miguel de Urcuquí 100119, Ecuador
2
Laboratoire de Recherche de Chimie Appliquée et de Génie Chimique, Hasnaoua I, Université Mouloud Mammeri, BP 17 RP, Tizi-Ouzou 15000, Algeria
3
Center for Midstream Management and Science, Lamar University, 211 Redbird Ln., Beaumont, TX 77710, USA
4
Earth and Life Institute, Catholic University of Louvain, B-1348 Louvain-la-Neuve, Belgium
5
Department of Chemical Engineering, Polytechnique Montréal, C.P. 6079, Succ. Centre-Ville, Montréal, QC H3C 3A7, Canada
6
Department of Chemistry and Biochemistry, Université de Moncton, 18, Ave Antonine-Maillet, Moncton, NB E1A 3E9, Canada
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(3), 424; https://doi.org/10.3390/nano13030424
Submission received: 17 December 2022 / Revised: 15 January 2023 / Accepted: 17 January 2023 / Published: 19 January 2023

Abstract

:
Selenium is an important dietary supplement and an essential trace element incorporated into selenoproteins with growth-modulating properties and cytotoxic mechanisms of action. However, different compounds of selenium usually possess a narrow nutritional or therapeutic window with a low degree of absorption and delicate safety margins, depending on the dose and the chemical form in which they are provided to the organism. Hence, selenium nanoparticles (SeNPs) are emerging as a novel therapeutic and diagnostic platform with decreased toxicity and the capacity to enhance the biological properties of Se-based compounds. Consistent with the exciting possibilities offered by nanotechnology in the diagnosis, treatment, and prevention of diseases, SeNPs are useful tools in current biomedical research with exceptional benefits as potential therapeutics, with enhanced bioavailability, improved targeting, and effectiveness against oxidative stress and inflammation-mediated disorders. In view of the need for developing eco-friendly, inexpensive, simple, and high-throughput biomedical agents that can also ally with theranostic purposes and exhibit negligible side effects, biogenic SeNPs are receiving special attention. The present manuscript aims to be a reference in its kind by providing the readership with a thorough and comprehensive review that emphasizes the current, yet expanding, possibilities offered by biogenic SeNPs in the biomedical field and the promise they hold among selenium-derived products to, eventually, elicit future developments. First, the present review recalls the physiological importance of selenium as an oligo-element and introduces the unique biological, physicochemical, optoelectronic, and catalytic properties of Se nanomaterials. Then, it addresses the significance of nanosizing on pharmacological activity (pharmacokinetics and pharmacodynamics) and cellular interactions of SeNPs. Importantly, it discusses in detail the role of biosynthesized SeNPs as innovative theranostic agents for personalized nanomedicine-based therapies. Finally, this review explores the role of biogenic SeNPs in the ongoing context of the SARS-CoV-2 pandemic and presents key prospects in translational nanomedicine.

Graphical Abstract

1. Introduction

Nanotechnology deals with the design of particles, tools and devices in the range between 1 and 100 nm in size with specific functions at the cellular, atomic and molecular levels [1,2]. Nanomedicine is a relatively new but fast-developing field that can potentially elicit a major impact on human health by combining nanotechnology-based techniques and methods with biomedical and pharmaceutical sciences [3,4,5,6,7]. Thus, nanotechnology can remarkably assist the therapy, diagnosis, monitoring and control of biological systems for future development of personalized medicine with tailored and optimized treatments [8,9,10]. Nanomedicine embraces nanopharmaceuticals, nanoimaging, sensing, plus diagnostics and therapy, also known as theranostics [11,12,13,14,15,16,17]. The interdisciplinary fields of nanotechnology and nanomedicine have been propelled to the forefront in investigations from academia, the pharmaceutical industry, clinical organizations, and several national and international funding and regulatory agencies [18,19].
Nanoparticles (NPs) exhibit unique properties over bulk materials, such as small size, high surface area, low polydispersity, colloidal stability, tunable surface charge and chemistry, easy modification, and multi-functionality (Figure 1) [20,21,22,23]. In recent years, NPs have been developed to harness their biological interactions at the molecular and cellular levels with a high degree of specificity towards diagnosing and treating several diseases [24,25,26,27,28,29,30]. The utilization of NPs has opened new therapeutic opportunities for agents which otherwise cannot be used effectively as traditional drug formulations due to their poor bioavailability and instability [31]. In particular, inorganic NPs, with their superior intrinsic chemical, biological, and magnetic properties, have been designed for several diagnostic [32,33,34], therapeutic [35,36], health [37], and agricultural [38,39] applications. Several nanomaterials hold great promise for integrating diagnostic and therapeutic applications, such as monitoring the biodistribution and accumulation at target site, observing and quantifying drug release, and longitudinally evaluating therapeutic efficacy [32,40].
Selenium nanoparticles (SeNPs) have attracted special interest since selenium is an essential micronutrient for the proper functioning of human and animal organisms in trace amounts [41]. Selenium is an important dietary constituent of at least 25 human selenoproteins and enzymes containing selenocysteine [41,42,43]. In the environment, selenium, whose chemical symbol is Se, exists under various oxidation states (2−, 0, 2+, 4+, 6+) and forms, such as selenate (Na2SeO4), selenite (Na2SeO3), selenomethionine (SeMet), selenocysteine (SeCys), and solid-state, zerovalent Se (Se0) [44,45]. Nanoscale selenium has attracted a great deal of interest worldwide due to its high degree of absorption, great bioactivity, low toxicity, and high efficiency in preventing oxidative damage compared to its organic and inorganic counterparts [46,47]. SeNPs possess remarkable anticancer [48,49,50], antioxidant [51], antidiabetic [52,53], antibacterial, and antibiofilm properties [54,55,56]. Equally noteworthy are the diagnostic applications of SeNPs in nanosensors, cellular imaging, epigenetics, and immunochromatography [34,57,58]. SeNPs can be fabricated through different physical, chemical, and biological techniques [59]. Nevertheless, the biological synthesis route, which relies on bacteria, yeast, plants, and microalgae, constitutes a widely explored green alternative that has witnessed tremendous developments and is especially suitable for SeNPs [42,60,61,62,63]. Bionanotechnology, a subset of green nanotechnology, is described as a set of methods that eliminate or reduce the utilization of toxic substances, resulting in cost-effective and eco-friendly alternatives [64,65]. In this sense, biogenic SeNPs are highly biocompatible and stable, due to natural coatings provided by biomolecules that, besides avoiding the use of chemical stabilizers, prevent the aggregation of the particles, improve their pharmacological activity, and protect them against physical and chemical degradation [66,67,68,69].
The translation of Se-based nanotechnology into clinical applications requires not only the development of safe, simple, sustainable, and cost-effective methods for the synthesis of SeNPs, but also a thorough understanding of their relevant physicochemical and biological properties that, in turn, might impact their in vitro and in vivo effects, safety control mechanisms, pharmacokinetics and pharmacodynamics, and potential biomedical applications. Research progress in biogenic SeNPs has implemented a safe-by-design principle, ensuring safety for both human health and the environment. However, this body of research that encompasses research articles and reviews has only discussed individual aspects of biogenic SeNPs, such as the synthesis, properties, and specific biomedical applications. Therefore, this study aims at filling this gap by providing a comprehensive and up-to-date review of the theranostics applications of biosynthesized SeNPs and their potential in translational nanomedicine. Furthermore, the properties, synthesis, and pharmacological activity of SeNPs focused on the molecular mechanisms, cellular interactions, and roles of selenoproteins are presented. Additionally, examples of biocidal and biomedical applications of biogenic SeNPs are detailed. Finally, key aspects that should be addressed to pave the way for clinical applications of biogenic SeNPs are discussed.

2. Selenium and Nanoselenium: General Information

Nanotechnology is emerging as an extraordinary platform of technical solutions for complex medical challenges. Nanomedicine involves nanotherapeutics [2], nanopharmaceuticals [70], nanoimaging [71], and theranostics [36]. Compared to conventional medicine, precision nanomedicine offers great physical and biological benefits, such as enhanced efficacy, improved pharmacokinetics and safety, reduced toxicity, and increased tissue selectivity of drug formulations [4,72]. Particularly, the major issue, in the case of Se, in the translation from bench to bedside, is the narrow window from therapeutic effect to toxicity due to the small margin of dosage error [73]. As mentioned above, selenium is an essential micronutrient, playing an important role in endocrine, reproductive, cardiovascular, and immune processes. It also acts as a pleiotropic agent associated with biotherapy and drug delivery for a better immune response and cancer prevention [74].
In addition to its above-mentioned inorganic and organic forms, Se is the principal constituent of some selenoenzymes, such as glutathione peroxidases (GPXs), thioredoxin reductases (TrxR), and deiodinases (DIO), which are essential in biochemical reactions of biological defense systems, including antioxidant activity [75]. Se-biogenic compounds can be found in living organisms in the form of methylated species, selenoamino acids, selenoproteins, selenium peptides, selenoenzymes, selenoamino carboxylic acids, and selenium derivatives of pyrimidines, purines, coenzyme A, cholines, and steroids, among others [76]. Most of these molecules play an important role in the organism’s defense against oxidative stress [77]. They also exhibit remarkable antioxidant and pro-oxidant effects limited by the dose, life span, chemical form of Se compound, route of administration, and oxidation state [78]. Table 1 classifies selenium compounds.
Usually, conventional selenium dietary supplements exhibit a low degree of absorption and enhanced toxicity [58]. Therefore, selenium is viewed as a controversial nutrient since high doses are toxic, provoking death, whereas selenium deficiencies can yield chronic and fatal health issues, such as diabetes, thyroid dysfunction, arthrophyma, Keshan disease, and cognitive problems [74,80]. Selenium toxicity has been tracked for decades, showing that low levels elicit efficacious anticarcinogenic activity while high levels can generate carcinogenesis, cytotoxicity, and genotoxicity (Figure 2) [81,82]. Indeed, there is some scientific consensus that the high pro-oxidant property of different redox-active forms of selenium compounds is the key factor to efficiently and selectively combat cancer [82,83,84,85]. For example, methyl selenocysteine (MeSeCys) and methyl selenic acid (MeSeA) were found to be potential anticarcinogenic selenocompounds with little toxicity and high bioavailability as indicated by the increased glycoprotein selenoprotein P (SEPP) biosynthesis [86]. Cao et al. reported the remarkable antitumor activity of MeSeCys in preclinical trials when combined with four different cytostatic drugs (cyclophosphamide, cisplatin, oxaliplatin, and irinotecan), offering protection against organ-specific toxicity [87]. Moreover, selenocysteine, a naturally occurring selenoamino acid, may be a promising anticancer candidate, as it enhances the apoptosis of the A375 human melanoma cell line when combined with 5-fluorouracil [88].
To date, the anticancer activity of selenium is not yet fully demonstrated, especially its in vivo therapeutic efficacy. The main challenge lies in delivering specific concentrations of redox-active selenium directly to the target site (tumor or metastatic cells) to bring about the cytotoxic effect [89]. In the quest of innovative systems to upgrade the bioavailability and the controlled release of selenium, much attention has been focused on SeNPs, which have appeared as the answer to the toxicological concerns due to their novel properties, including high specific surface area, high degree of absorption, and low toxicity when compared to their inorganic and organic analogs [46,58]. As the alternative to elemental zero-valent selenium (Se0), nanoscale selenium offers the advantage of a significantly low toxicity without affecting its ability to upregulate selenoenzymes at nutritional levels and induce phase II enzymes at supra-nutritional levels [90]. Nanoscale selenium is bright red, highly stable, and soluble; it has been processed for pharmaceutical and medical applications due to its anticancer, antimicrobial, antioxidant, and antidiabetic activity [62,91]. When released from the NP surface, the unstable Se0 atoms are readily oxidized to inactive forms. To prevent this instability, proteins and polysaccharides are mainly used as nanocarriers; these include chitosan [51,92], egg white lysozyme [93], β-lactoglobulin [94], acacia gum, and carboxymethyl cellulose [95].
SeNPs have been integrated in simultaneous treatments involving immunotherapy, chemotherapy, and radiotherapy because they not only possess a sensitive response to radiation stimuli but also exhibit excellent anticancer activity and immune checkpoint inhibitor effect [96]. Indeed, SeNPs deliver the chemotherapeutic drug doxorubicin (DOX) to tumor sites by systemic administration, thereby exerting immunomodulatory activity by enhancing natural killer (NK) cell function [96]. Moreover, SeNPs can effectively increment the persistence of cytokine-induced killer (CIK) cells in peripheral blood in the body. For instance, the combination of SeNPs and CIK cells induces natural killer cells to infiltrate into tumors, and shapes tumor-associated macrophages to trigger powerful immune responses for effective cancer immunotherapy [35]. In addition, SeNPs enable signal transduction from the lysosomes to the nucleus and further potentiate γδ T cell anti-tumor cytotoxicity, promote the production of surface receptors present at the immune cells, typically NKG2D, CD16, CD44, and IFN-γ, and inhibit the expression of PD-1 receptors [97].
The effects of SeNPs at the cellular and tissue levels have been thoroughly investigated in, for example, type 2 diabetes mellitus (T2DM) treatment [73], immune and antioxidative responses [98,99], atherosclerosis treatment [100], and semen quality and testis ultrastructure studies [101]. In addition, several studies indicate that nanostructured selenium is gaining attention among dietary supplements and therapeutic agents [76]. In addition, its immunostimulatory [63] and protective effects against heavy metal intoxication [102,103,104] are well-documented. For instance, Sheiha et al. reported the effects of nanoscale selenium supplementation on the growth performance, kidney and liver functions, carcass traits, antioxidant indices, and inflammatory cytokines of growing rabbits subjected to thermal stress [105]. Moreover, Tran et al. combined the carcinostatic activity of Se nanoclusters with the mechanical properties of titanium to build a new anticancer bone implant [106]. Furthermore, Bartůněk et al. evaluated the use of PEGylated SeNPs for specific antimicrobial coatings [107]. Lastly, SeNPs were used as coatings to inhibit biofilm formation [56,108,109]. Figure 3 summarizes key properties and applications of SeNPs [58].

3. Selenium Nanoparticles: Methods of Synthesis

Owing to their unique surface activity and particle dispersion, SeNPs offer several advantages over bulk selenium-based materials, such as larger biological activity, higher catalytic efficacy, greater bioavailability, and lower toxicity [42]. The ultimate outlook of precision nanomedicine has consisted in designing and constructing smart NPs for clinical translation. However, the ability to fabricate NPs free of any toxic or hazardous substances is very challenging, especially for applications in nanomedicine. This quest has driven the evolution of different approaches to synthesize SeNPs.

3.1. Physical Methods

Physical techniques include, to name a few, microwave irradiation, ultraviolet irradiation, laser ablation, and ultrasonic field treatment [91]. Bright red selenium nanoballs, nanotubes and multi-armed nanorods, with diameters ranging from 20 to 130 nm, were obtained via a microwave approach in which the L-asparagine/H2SeO3 concentration ratio and the irradiation time controlled the NP diameter and morphology [110]. Using a similar method, pure hexagonal phase SeNPs were obtained with selenium tetrachloride as the precursor under microwave irradiation [111]. Relying on γ-rays, water-stable SeNPs are produced in the presence of various natural macromolecules, such as citrus pectin, sodium alginate, chitosan, and aqueous extract of fermented fenugreek (Trigonella foenum-graecum) powder [112]. Moreover, SeNPs were synthesized via nanosecond pulsed laser ablation by irradiating selenium pellets while chitosan acted as the capping agent [113]. Other studies reported the fabrication of selenium nanostructures via laser ablation and discussed their antibacterial activity [114,115,116]. Moreover, cubic-like SeNPs were fabricated by employing a self-assembly process [117]. In addition, SeNPs were obtained via sonochemistry [118]. Physical procedures offer advantages over chemical ones since the latter may require a final calcination step that makes them unsuitable for targeted applications [58].

3.2. Chemical Methods

The chemical synthesis of SeNPs is the most conventional and widespread method, which comprises the reduction of metal salts using chemical reducing agents in aqueous or organic media [59]. In addition, the inclusion of stabilizers (e.g., polysaccharides) in the reaction mixture enables the size and shape control of the NPs, ensures their colloidal stability, and avoids their aggregation; hence, the stabilizing molecules are of paramount importance [119]. Basically, chemical methods constitute a time-saving strategy but can be highly expensive and environmentally harmful. Several studies have detailed the chemical synthesis of SeNPs [48,54,120]. For instance, a solution-phase approach yielded monodisperse spherical SeNPs of 20 nm by reducing selenous acid solution using ascorbic acid and water-soluble polysaccharides, such as chitosan, konjac glucomannan, acacia gum, and carboxymethyl cellulose [95]. Size-controlled, stable, and positively charged SeNPs, with sizes ranging from 30 to 150 nm, were fabricated using selenous acid and chitosan to build special chain-shaped intermediates to enhance the cellular uptake and anticancer activity [121]. Another study yielded uniform spherical SeNPs of 103 nm through the reduction of selenic acid with ascorbic acid in the presence of chitosan [92]. These nanostructures continuously grew via a “bottom–up” approach and “top–down” shrinkage processes, generating a stable nanosystem towards pH and enzyme treatment. On the other hand, Chung et al. reported on the synthesis of bovine serum albumin (BSA)-coated selenium nanoparticles (BSA-SeNPs) of 20–800 nm through a one-step reaction relying on sodium selenite, ascorbic acid, and BSA [122]. During the fabrication process, the agitation speed (100 rpm, 400 rpm, and 600 rpm) and sodium selenite to ascorbic acid molar ratios (1:2, 1:4, 1:5, 1:6) impacted the NP size. Another study described the concomitant reduction of sodium selenite by glutathione, and the formation of SeNPs with an effective antibacterial activity of over 99% when coating polyvinyl chloride endotracheal tube substrates [123]. Monodisperse, spherical SeNPs of 46 nm in size were obtained by reducing SeCl4 in the presence of ascorbic acid [124]. A facile and versatile electrochemical approach produced spherical SeNPs of 43–85 nm in size using selenium powder doped-carbon paste electrode [125]. A simple wet chemical method relying on sodium selenosulfate in ionic liquid resulted in spherical SeNPs of 76–150 nm in diameter [126].
The use of chitosan (CTS) as an effective material to promote NP production has been extensively investigated owing to its exceptional properties in terms of biocompatibility, biodegradation, and enzymatic resistance [127]. This enables its various applications in the biomedical field, such as in tissue engineering, drug delivery, wound healing, and gene therapy [128]. For example, spherical SeNPs were embedded into CTS microspheres through a spray-drying method for selenium oral delivery with a high efficiency and good biosafety [129]. Selenium nanoencapsulation within CTS networks exhibited decreased toxicity, enhanced antioxidant activity, and controlled in vitro release [130]. The effects of selenium-loaded CTS nanoparticles were assessed in terms of cellular selenium retention, cell survival, and DNA damage in response to selenium exposure, giving rise to novel selenium delivery systems with high specificity and low toxicity for dietary and therapeutic applications [127]. To highlight the promise of SeNPs in cancer treatment, 105-nm SeNPs modified with ferulic acid were synthesized via a simple, low-cost approach to investigate their antitumor activity and DNA-binding affinity [131]. Lastly, a novel, potentially scalable room-temperature procedure allowed the fabrication of SeNPs using selenium oxide as the precursor and lignosulfonate as the stabilizer [132].

3.3. Biological Methods

The synthesis of SeNPs via biological methods is receiving increased attention. This green and sustainable synthesis overcomes several drawbacks, including the cost, complexity, and toxicity concerns, and improves the effectiveness of the process [60]. In this regard, several organisms, such as plants, algae, fungi, and bacteria, have been examined in the biogenesis of SeNPs. For example, Medina et al. fabricated spherical SeNPs of 90–150 nm using Staphylococcus aureus, methicillin-resistant S. aureus (MRSA), Escherichia coli and Pseudomonas aeruginosa, and found that the antibacterial potential of SeNPs made using a specific bacterium was more efficient against the same bacterial species [133]. Lampis et al. used Stenotrophomonas maltophilia to synthesize SeNPs of 160–250 nm, depending on the age of the cultures [134]. Kora et al. isolated the selenite-reducing bacterium Bacillus cereus from a lake contaminated by industrial waste to fabricate amorphous and spherical SeNPs of 93 nm [135]. Khoei et al. produced biogenic SeNPs via intra- and extra-cellular pathways using two strains of Burkholderia fungorum; this formation was attributed to cytoplasmic enzymatic activation mediated by electron donors [136]. In addition, Kamnev et al. detailed the extracellular production of SeNPs through selenite reduction by living biomass of the rhizobacterium Azospirillum brasilense [137].
In the same vein, fungi are also efficient Se-reducing organisms, able to synthesize Se0 as well as Se-methylated compounds [60,138,139]. Various fungi have been explored for the green fabrication of SeNPs [140,141,142,143,144,145]. For instance, Pleurotus ostreatus-treated aqueous extract of fermented powdered fenugreek seeds was an effective capping and reducing agent to produce SeNPs due to the high amounts of amino acids, proteins, and other reducing agents [112]. Another study reported that the cell wall, cytoplasm, and proteins of Mariannaea sp. provided templates for the reduction of Se(IV) to Se(0) through various detoxification mechanisms [146]. Mycogenic SeNPs were obtained from three fractions of Trichoderma atroviride: namely, culture filtrate, cell lysate, and cell wall debris; these NPs exhibited antifungal activity against several phytopathogens [147]. A simple and efficient method relying on Cordyceps sinensis exopolysaccharides yielded well-dispersed and stable SeNPs [148]. Monodispersed SeNPs of ~22 nm in size, synthesized by P. chrysogenum filtrate, were incorporated within carbon nanotubes under γ-irradiation [108]. Similarly, the aqueous extract of Aspergillus oryzae-fermented lupin was found to reduce Se ions into spherical, isotropic, and poly-dispersed SeNPs under γ-irradiation [149].
On the other hand, some researchers favor yeast-mediated biosynthesis of metal NPs since yeast biomass production is simple, easy to obtain and scalable [60]. For example, several strains of Saccharomyces reduced selenium anions to subsequently form SeNPs [150,151,152]. A suitable green analytical procedure using the yeast S. boulardii enabled the direct monitoring of SeNPs synthesis yield [151]. In the same vein, the genetically engineered metal-resistant strain of Pichia pastoris is an efficient nanofactory for intracellular SeNP biosynthesis [153]. Finally, the yeast Nematospora coryli gives rise to intracellular spherical SeNPs of 50–250 nm that display potential anti-Candida and antioxidant activities [154].
Plants have been extensively used for SeNP biosynthesis [155,156,157,158,159]. For instance, Withania somnifera possesses constituents, such as alkaloids, flavonoids, phenolics, tannins, and terpenoids, which act as good bio-reductants and capping agents for the synthesis of SeNPs [160]. Phenolic and alcoholic compounds present in guava (Psidium guajava) leaf extract are responsible for the synthesis of SeNPs and their stabilization [161]. Cassia auriculata leaf extract reduces selenite ions into SeNPs [162]. Lemon [163], prickly pear [164], Abelmoschus esculents [165], orange peel [166] Macleaya cordata [167], and Hibiscus sabdariffa [168], are examples of plants involved in the biosynthesis of SeNPs. In addition, a facile single-step and green in situ method relying on the novel RTFP-3 polysaccharide, extracted from Rosa roxburghii fruit, enabled the production of size-controlled and stable SeNPs [169]. In addition to some of the previously mentioned biological properties, biogenic SeNPs synthesized using Clausena dentata leaf extract exhibit a strong, dose-dependent mosquito larvicidal activity [170].

4. Properties of Selenium Nanoparticles

The morphology, size, and properties (i.e., physical, chemical, biological) of nanomaterials are determined by different factors and reaction parameters, such as synthesis techniques, starting materials, specific surfactants or additives, pH, reaction time, reaction temperature, and the nature of the solvent [171,172]. This section reviews several physicochemical, optoelectronic, catalytic, and biological properties of SeNPs that will help understand their biomedical applications.

4.1. Physicochemical Properties

Selenium (Se) is a metalloid that possesses intermediate properties between a metal and a non-metal. It is stable and is not oxidized at room temperature [43]. Selenium shares several chemical and physical properties with its other non-metal counterparts found in the oxygen family (group 16 of the Periodic Table): sulfur and tellurium. Its outer electronic configuration is 4s24p4. The atomic number and weight of Se are 34 and 78.96, respectively [173]. Selenium possesses over 20 different isotopes, among which only 6 are stable: 74Se, 76Se, 77Se, 78Se, 80Se, and 82Se [174]. Its melting point is relatively low (~217 °C) and its photoconductivity is high (~8 × 104 S·cm−1) [175]. Moreover, selenium shows a catalytic activity toward organic hydration and oxidation reactions, intrinsic chirality, high refractive index, large birefringence, and relatively large piezoelectric, thermoelectric, and nonlinear optical responses [176,177]. In terms of allotropy, selenium can exist in amorphous (a-Se) and crystalline varieties (c-Se): gray (trigonal) selenium (containing Sen helical chain polymers) known as “metallic” selenium; rhombohedral selenium (containing Se6 molecules); three deep-red monoclinic forms: α-, β- and γ-selenium (containing Se8 molecules); amorphous red selenium, and black vitreous selenium [178]. Crystalline selenium is thermodynamically the most stable structure, exhibiting an atomic radius of 1.17 Å [179]. Several studies indicate that the phase transition between c-Se and a-Se occurs in the charge/discharge process [44,180,181]. a-Se is an efficient photoconversion material, frequently used in several imaging applications, including ultrahigh-sensitivity pickup tubes [182] and solid-state image sensors [183]. Nevertheless, a-Se has a poor spectral response at long wavelengths and requires a high operation voltage [184].
c-Se structure appears as an alternative to a-Se in the photoconversion layer of solid-state image sensors. Crystalline selenium, its most stable form, consists of a long helical chain arranged in a hexagonal array [184]. Several reports detailed the synthesis and properties of c-Se. For instance, Takiguchi et al. fabricated single crystals of trigonal Se with a cylindrical shape and a diameter of about 8 mm [185]. Liu et al. studied the photoconductance of single-crystalline selenium nanotubes (SC-SeNTs) under a 633 nm illumination of various intensities; their results suggest that SC-SeNTs are potentially good photo-sensor materials as well as very effective solar cell materials [186]. Moreover, the synthesis of uniform nanowires of c-Se with uniform lateral dimensions in the range of 10–30 nm was documented [175]; these nanowires can potentially be converted, under adequate conditions, into other functional materials, such as ZnSe and CdSe. Another study assembled c-Se films via doping with different amounts of various halogens, such as chlorine (Cl): 0.50 and 500 ppm; bromine (Br): 50 ppm; and iodine (I): 50 ppm, to investigate the concentration effect on surface enhancement [184]. This technology is suggested to help in the design of super-high-definition imaging systems.
The physicochemical properties of SeNPs have been extensively explored. Chen et al. indicated that in the case of chitosan-stabilized SeNPs the molecular weight of chitosan regulates the Se biological and physicochemical properties, including crystallinity, surface charge density, and hydrophobicity [187]. Zhang et al. proved that chitosan-selenium nanoparticles (CTS-SeNPs), of 80–120 nm in size and different weights, exhibited excellent physicochemical stability after 30 days of storage [92]. Yu et al. showed that CTS-SeNPs, with a particle size smaller than 180 nm, remained stable for 60 days [121]. In addition, Hageman et al. studied the effects of pH (6–9) and temperature (20–50 °C) on the structure, morphology, and stability of biogenic SeNPs using scanning electron microscopy, X-ray diffraction, and light microscopy [188]. As a result, selenium particle crystallinity, shape, and color can be controlled by temperature and pH; for instance, gray crystalline hexagonal acicular SeNPs form at mild temperatures or high pH, whereas red amorphous nanospheres prevail at low temperatures and low pH.

4.2. Optoelectronic Properties

Owing to their inherent quantum confinement, SeNPs have distinct, striking shape- and size-dependent physical properties. Selenium is a typical semiconductor with a band gap of 1.6 eV (775 nm) [189]. Due to its ability to absorb X-rays and its high resistivity, ranging from 1012 to 1014 Ω, Se has been considered as an outstanding option for photodetectors and xerographic photoreceptors with ultra-low dark current and high sensitivity [190,191,192,193]. Since selenium is one of the primary substances that possess photoelectric conductivity [57], potential detection applications of a-Se have been investigated, mainly due to its ultra-high photosensitivity, by using avalanche multiplication inside the solid, for example, in imaging photodetectors using low-dose X-rays [194,195], X-ray photoelectron spectroscopy (XPS) and Raman spectroscopy [196], nitrogen (N)-doped diamond cold cathode [197], or driven by a diamond cold cathode [198]. On the other hand, c-Se possesses lower concentrations of selenium than a-Se, thus exhibiting less non-radiative recombination loss. Besides being low cost and highly scalable [199], c-Se has been employed to fabricate solar cells [179,200]. Moreover, Sharma et al. used a genetic algorithm-based code, which consists of universal structure prediction evolutionary xtallography (USPEX) and molecular dynamics, to obtain at least 70 distinct equilibrium geometries for each selenium cluster [201]. The authors analyzed the structural features of Se clusters, including the bond length, bond angle, point symmetry, and shape of the geometries, demonstrating that the lowest energy geometries are one-dimensional rings (buckled or distorted) with each atom possessing only two nearest neighbors.
The optical properties of nanomaterials are highly influenced by multiple factors, such as the size, shape, surface modification, doping, and interactions with other materials [202,203,204]. Unique features, such as the nanoscale dimension, increased energy level spacing (quantum effect), and surface plasmon resonance, determine the size-dependent optical properties, enabling several applications in the biomedical field, energy, catalysis, and environment [174,205]. The NP size distribution can be estimated by optical absorption and luminescence spectra generated by quantum confinement effect [206]. For instance, Rajalakshmi and Arora found that a 0.235 eV blueshift appears in the optical absorption and photoluminescence (PL) energy of SeNPs, which is useful to estimate the particle size [207]. Lesnichaya et al. showed that SeNP polydispersity broadens the absorption and excitation-dependent luminescence spectra [206]. In addition, laser irradiation reduces the size of spherical β-SeNPs (69 nm) below 3 nm and converts them into more closely packed α-Se quantum dots (QDs); then, surface defect density and electron trap level of QDs increase with irradiation time, which decreases energy levels [208]. Another study used optical spectroscopy to show a usual blue shift in the optical spectra of α-monoclinic SeNPs of the order of 40 Å, demonstrating a band gap widening [176]. This blueshift in the band gap energy of Se in comparison with its bulk counterpart appears when the particle size is smaller than its Bohr excitation radius; thus the bandgap is enlarged due to the quantum confinement effect [189]. In this connection, biomolecules, such as proteins and amino acids, absorb light and provide thermodynamic stability [209,210]. Fourier transform infrared (FTIR) analysis suggests that a strong interaction between Se atoms and proteins present in P. alcaliphila may be responsible for a drastically decreased intensity of spectral peaks of SeNPs [211]. The optical properties can promote light-induced release of drugs either covalently bonded to or encapsulated with SeNPs [57].

4.3. Catalytic Properties

Selenium has been used as a platinum-free, methanol-tolerant cathode material with great stability and electrocatalytic activity, generating chemical resistance to oxidation and hydrolysis [212,213]. In recent years, SeNPs have attracted special attention, particularly due to their unique redox properties, large surface areas, efficient catalytic activity, and low toxicity [214]. Relying on UV–Visible spectroscopy, Dumore and Mukhopadhyay employed the 1-diphenyl-2-picrylhydrazyl free radical scavenging (DPPH-FRS) reaction as a model to monitor the catalytic activity of aqueous selenium nanoparticles (Aq-SeNPs) at different pH (6, 6.5, 7 and buffered 7) [215]. Following pseudo-first-order kinetics, the FRS reaction depended on DPPH concentration since the rate of DPPH-FRS reaction increased proportionally with the amount of Aq-SeNPs, proving that the catalytic reaction occurs at the NP surface. Likewise, other studies have reported on the excellent electrocatalytic performance of selenium-containing compounds, either as a counter-electrode material for dye-sensitized solar cells [216] or as a cathode catalysts for methanol fuel cells [217].
Semiconductor chalcogens, such as those made of selenium, have a direct bandgap, and can be potentially used for the degradation of dyes due to their thermoconductivity, anisotropy, and high photoconductivity [218,219,220,221,222,223,224,225,226,227]. For instance, Ameri et al. described the photocatalytic discoloration of the anionic triphenylmethane dye, bromothymol blue (BTB), using biogenic SeNPs under ultraviolet (UV) illumination (15 W) for 60 min [228]. Another study showed that single-crystalline Se nanorods (SeNRs) were able to degrade methylene blue (MB) in the dark after a short period of irradiation [229]; the superb catalytic performance of SeNRs over commercial nanoparticles was due to the efficient interior charge carrier transfer, and thus the enhanced carrier utilization efficiency. Likewise, Tripathi et al. evaluated the photocatalytic activity of biogenic fluorescent SeNPs in MB decomposition under UV irradiation [230]. Zhang et al. demonstrated the visible light-driven photocatalytic capacity of super-long single-crystalline t-SeNRs for methyl orange (MO) degradation [231]. A similar study reported that monoclinic, spherical SeNPs degraded rhodamine B (RhB) in the presence of H2O2 more efficiently than t-SeNRs by comparing to results published by another group [232]. This investigation highlights the mechanism of the remaining photomemory effect mechanism of pre-irradiated spherical SeNPs in the dark.
Furthermore, when doped with selenium, bismuth sulfides increased the degradation rate of MB under visible-light irradiation [233]. These improvements were presumably caused by photoelectrons and holes generated by Se dopants in Bi2S3 photocatalysts. Moreover, Se-doped copper oxide NPs (Se-doped CuO NPs) were used to build a photo-Fenton based degradation system for 4-bromophenol under UV irradiation for 90 min in the presence of H2O2, achieving a rate of 0.057 min−1 [234]. Therefore, doping Se confers extraordinary photo-absorption properties, increases NP surface area, and enhances the in situ generation of hydroxyl radicals. In addition, biosynthesized SeNPs using W. somnifera leaf extract exhibit excellent photocatalytic activity in the gradual degradation of MB from deep blue to colorless dye solution under sunlight; besides holes, superoxide and hydroxyl radicals were identified to be involved in this process [160].
SeNPs, obtained using the proper reducing agents, such as ascorbic acid, have proven to be a suitable adsorbent for the removal of copper cations from aqueous solution [235,236]. Further, biogenic SeNPs have been employed as effective and fast adsorbents for zinc ions, mainly through inner-sphere complexation [237]. Another study showed that negatively charged biogenic SeNPs, produced by aerobic granular sludge in a sequencing batch reactor (SBR), efficiently removed Cd(II) [238]. The resulting monolayer maximum adsorption capacity was 59.7 mg g−1, enhanced by increasing pH but decreased by increasing adsorbent dosage. In addition to this, selenium combined with ruthenium NPs increased the electrocatalytic oxygen reduction reaction (ORR) by enhancing the oxygen adsorption site and via the electron bridge features of selenium [239].

4.4. Biological Properties

As an essential micronutrient, selenium is integrated into 25 selenoproteins in the form of the amino acid selenocysteine (SeCys). In addition, selenium modulates a myriad of key biological processes, such as the cellular response to oxidative stress, cellular differentiation, immune response, redox signaling, and protein folding [75,240]. Moreover, selenium plays important biological roles in maintaining thyroid activity, immunity, and homeostasis through the production of oxidoreductases, such as glutathione peroxidases (GPX), iodothyronine deiodinase (DIO) and thioredoxin reductase (TrxR), and the plasma selenium transport protein (SePP1) [75,80], and preventing several pathologies, such as cancer, diabetes, and aging-related diseases, to name a few [241,242,243]. The main selenoprotein families are GPXs that include (i) five Se-dependent members and other non-Se-dependent GPX isoenzymes, which have oxidoreductase functions and also regulate the immune response; (ii) DIOs that catalyze the conversion of T4 (thyroxine) to T3 (triiodothyronine) and rT3 (reverse T3); and (iii) TrxR, which modulates the transcription and signal transduction functions [244,245] (Table 2). To investigate its biomedical applications, the biological properties of selenium need to be understood; however, they are not yet fully unraveled. Hence, the present critical review outlines some key biological features of SeNPs.

4.4.1. Antioxidant Properties

An antioxidant is a substance that greatly inhibits or delays the oxidation mechanism while the antioxidant activity measures the inhibition rate of the oxidation process [249]. The antioxidant activity of SeNPs is principally associated with the mammalian selenoenzymes GPX, TrxR, and IDO [250]. Selenium, as part of the antioxidant defense system in the liver, plays an essential role against oxidative stress. It has been demonstrated that Se supplementation can enhance enzyme levels, such as GPX, that prevent reactive oxygen species (ROS) accumulation and decrease cellular damage [251,252]. GPXs are able to actively detoxify a variety of peroxides, such as H2O2, phospholipid and fatty acid hydroperoxides, and hydroperoxyl groups of thymine [174]. TrxRs exhibit a detoxifying action by forming a redox system with its substrate, thioredoxin. These metabolic processes produce the most common free radicals in nature: reactive nitrogen species (RNS) and ROS. ROS derive from oxygen and include peroxyl radical, superoxide radical, perhydroxyl radical, hydroxyl radical, and non-free radical species, such as hydrogen peroxide and singlet oxygen. RNS and ROS are highly unstable, as their outermost electron shell is occupied by an unpaired electron; this leads to the removal of electrons from other compounds to attain stability, which yields a chain reaction cascade that may produce more reactive species.
Research has demonstrated that excessive levels of ROS may cause oxidative stress and redox imbalance in the cell [253]. This can disrupt or damage proteins, DNA, and lipids, resulting in cardiovascular and neurodegenerative diseases, e.g., Parkinson’s and Alzheimer’s [254]. Selenium has attracted attention because its antioxidant properties are predominantly exerted owing to its incorporation into selenoproteins that can catalyze the reduction of disulfide bonds in proteins and peptides [255,256]. Indeed, Se is the main component of the antioxidant enzymes glutathione peroxidases (GPXs), thioredoxin reductases (TrxRs), and iodothyronine deiodinases (DIOs) that protect cells from oxidative stress. For example, selenite, which is an essential dietary supplement for mammals, is present in the active center of the antioxidant enzyme GPX and protects membrane lipids and macromolecules from oxidative stress [257]. Notably, in vitro and in vivo investigations have demonstrated that all selenium compounds under different oxidation states (2−, 0, 4+ and 6+) enhance selenoprotein expression; thus, selenium compounds under different oxidation states have shown great bioavailability as precursors for selenoprotein biosynthesis [89,258]. For instance, Tobe and Mihara found that selenide is involved in synthesis of selenophosphate synthetase (SPS), which consequently produces selenophosphate (SeP), the key selenium donor for the synthesis of selenoproteins and selenium-modified tRNA [259]. Hence, it is of great importance to understand the biotransformation of selenium and reaction mechanisms of the enzymes implicated in selenium metabolism.
SeNPs have proven efficient in enhancing the activity of selenoenzymes to combat oxidative stress with equal effectiveness and less toxicity when compared to MeSeCys, SeMet, and selenite [73]. SeNPs possess radical scavenging properties and reduce oxidants, including 1,1-diphenyl-2-picrylhydrazyl (DPPH), superoxide anion (O2•−), singlet oxygen (1O2), and carbon-centered free radicals [42,260,261,262]. This function is size-dependent since smaller SeNPs possess a higher free radical scavenging potential. Moreover, SeNPs are shown to restore T3, T4, glutathione (GSH), superoxide dismutase (SOD), and catalase levels in animal models, and decrease K2Cr2O7-induced oxidative stress in thyroid glands [73,103].

4.4.2. Scavenging Mechanism of Reactive Oxygen Species

ROS are chemically reactive molecules, produced through a myriad of extra- and intra-cellular pathways, that include at least one oxygen atom in each molecule [263]. They include free radicals, a species containing one or more unpaired electrons of oxygen, such as superoxide (O2•−), hydroxyl radical (OH), and singlet oxygen (1O2), as well as nonradical oxidizing agents, such as hypochlorous acid (HOCl) and hydrogen peroxide (H2O2), formed by the partial reduction of oxygen and singlet oxygen [263]. Mitochondria are the main intracellular source of O2•−, produced by a side reaction of the respiratory chain. The superoxide anions are formed through the conversion of a small percentage of oxygen molecules (1–2%) that are not reduced to water in the mitochondrial electron transport chain (ETC) [263].
The ability of selenium compounds to scavenge ROS is well documented [51,257,263]. For instance, the dose-dependent free radical scavenging activity (FRS) of water-soluble SeNPs (Aq-SeNPs) using 1,1-diphenyl-2-picrylhydrazyl (DPPH) and 2,2′-azinobis (3-ethylbenzothiazoline-6-sulphonic acid) (ABTS) is revealed by a gradual color change of DPPH from intense purple to light yellow, and of ABTS radical from bluish-green to colorless in the presence of Aq-SeNPs at slightly acidic to neutral pH (6.0, 6.5, 7.0 and buffered 7.0) [215]. Importantly, the scavenging ability of the NPs was found to be stronger than that of sodium selenite. Furthermore, SeNPs of 103 nm in size, stabilized with chitosan of different molecular weights (CTS-SeNPs), exhibited a capability to scavenge free radicals at different levels of DPPH, ABTS, and lipid peroxide models [261]. The efficient penetration of CTS-SeNPs into cells and tissues prevents the accumulation of ROS and lipofuscin (LF), protects GPX activity, and decreases selenium in vitro and in vivo cytotoxicity. In addition, SeNPs of different sizes, ranging from 5 to 200 nm, possess important effects both on scavenging free radicals and protecting DNA from oxidation in a size-dependent fashion; as found in an in vitro model, the smaller, the better [261].
Neuroprotective drugs in tandem with ROS scavenging nanocarriers constitute excellent agents to synergistically protect neurons and help restore nerve function. For instance, multifunctional SeNPs were modified with the soluble polysaccharide–protein complex (PTW) and PG-6 peptide (PLGLAG) and loaded with the therapeutic agents monosialotetrahexosylganglioside (GM1) and tetramethylpyrazine (TMP) to effectively treat spinal cord injury (SCI) (Figure 4) [264]. These SeNPs@GM1/TMP were found to attenuate ROS overproduction, prevent mitochondrial dysfunction by up-regulating the expression of pro-apoptotic proteins Bcl-2 (B-cell lymphoma-2) and Bcl-xl (B-cell lymphoma-extra-large), down-regulating the expression of anti-apoptotic proteins Bax (Bcl-2 associated X) and Bad (Bcl-2 associated agonist of cell death), and inhibit the activation of p53 and mitogen-activated protein kinase (MAPK) pathways. They also display protective effects against tert-butyl hydroperoxide (t-BOOH)-induced G2/M phase arrest and apoptosis. Behavioral assessments in mice demonstrated that SeNPs@GM1/TMP constitute promising therapeutic agents to potentially improve the function recovery of rats after SCI.

4.4.3. Pro-Oxidant Activity

Selenium exerts dual effects: at low concentrations, it possesses antioxidant activity by maintaining the intracellular redox status, whereas at higher concentrations, it acts as a pro-oxidant that generates oxygen radicals and provokes apoptotic cell death [248]. Among all the explored inorganic nanoparticles, much attention has been placed on SeNPs due to their cytotoxic activity since ROS are generated inside malignant cells [265,266]. Furthermore, SeNPs are found to be potential tools in fighting drug resistance, either as effective chemotherapeutic agents or as excellent carriers for gene and drug delivery [174]. A study showed that ROS generation by SeNPs is the initial major cellular event prior to cell cycle arrest and/or apoptosis [267]. Indeed, SeNPs enter malignant cells via receptor mediated endocytosis, mainly due to an acidic pH state with redox imbalance [268,269,270,271,272]. This process leads to an NP pro-oxidant behavior by free radical generation, on one side, which induces mitochondrial membrane disruption and, consequently, the leakage of mitochondrial proteins and endoplasmic reticulum (ER) stress on the other side [61,273,274,275]. Hence, several apoptotic molecular pathways which are regulated by SeNPs are activated or modulated, including TLR4/TRAF3/MFN1 (toll-like receptor-4/TNF receptor associated factor 3/mitofusin-1) [276]; p53, MAPK (mitogen-activated protein kinases) and AKT (protein kinase B) [277,278,279]; Bcl-2 family proteins [280]; ROS/JNK (c-Jun N-terminal kinase) [281]; PI3/Akt/mTOR (mammalian target of rapamycin); mTOR and NF-κB (nuclear factor kappa B) [282]; and caspase apoptotic pathways [268,280,283]. The regulation of these pathways is crucial for oncogenic signaling due to a considerable decrease in cellular proliferation and angiogenic signaling by obstructing the growth-promoting signaling in the vicinity of tumor cells. For instance, Pi et al. reported that SeNPs significantly reduce the adhesion force and Young’s modulus of MCF-7 cells, leading to a diminished expression of trans-membrane CD44 molecules and necrosis of MCF-7 cells [284]. Zeebaree et al. demonstrated that spherical biogenic SeNPs, synthesized using Asteriscus graveolens leaves, enhance the level of ROS and lipid peroxidation while causing the apoptosis of HepG2 by glutathione depletion and a decrease in the mitochondrial membrane potential [285].

4.4.4. Production of Reactive Oxygen Species

ROS have been recognized as signal mediators implicated in cell growth, differentiation, cycle progression, and death [286]. In regards to oxidative stress, ROS overaccumulation in cells leads to a reaction with different cellular components, causing oxidative cellular injury and cell death [287]. The term “oxidative stress” is attributed to the perturbations of the physiological redox homeostasis when the rate of cellular reduction is exceeded by the rate of cellular oxidation [248]. When ROS overwhelm the cellular antioxidant defense system, either through enhanced ROS levels or a decreased cellular antioxidant capacity, oxidative stress occurs. Particularly, selenium compounds possess a high capacity of exerting oxidative stress by oxidizing thiols and generating ROS, thereby termed as redox-active selenium compounds, e.g., selenocysteine, selenite, methyl selenic acid, and MeSeCys [89]. This clearly demonstrates that selenium does not only have antioxidant properties, but also pro-oxidant properties. Thus, wherever applicable, redox-active selenium compounds are not antioxidant by themselves, but only when supplied at dietary dose levels equivalent to physiological optima and incorporated into selenoproteins with oxidoreductase functions. On the other hand, at supra-physiological levels, redox-active selenium compounds can induce oxidative stress, becoming a novel tool in cancer therapy based on ROS-mediated mechanisms [288]. The effectiveness of selenium compounds for in vivo chemoprevention relies on their capability to regulate the cell cycle, stimulate apoptosis, and restrain tumor cell migration and invasion in vitro [90,289].
SeNPs upregulate the activity of selenoenzymes with more efficiency and less toxicity when compared to other selenocompounds, thereby serving as a potential antioxidant and chemopreventive agent [58,96,290]. SeNPs can be reduced into selenide by a thioredoxin (Trx)- or glutaredoxin (Grx)-coupled glutathione system to produce ROS more efficiently than selenite, especially at low levels of NADPH. This is because elemental selenium requires only a single step reduction to selenide anion, thus triggering redox cycling with oxygen [291]. This process leads to a rapid and selective hyper-accumulation of SeNPs in cancer cells, which causes catastrophic oxidative stress and cell death. This underlying concept relies on two observations: firstly, the presence of higher basal levels of ROS in cancer cells compared to normal cells; secondly, cancer cells possess lower tolerance to increased levels of ROS than normal cells [288,292,293]. The process of ROS production, mediated by SeNPs, is illustrated in Figure 5.
The intraperitoneal delivery of SeNPs has emerged as an effective and safe approach in preventing the growth of cancer cells in the peritoneal cavity without remarkable liver toxicity symptoms [294]. For example, Zhao et al. demonstrated that SeNPs delivered to hepatocarcinoma-22 cells in the peritoneal cavity of mice induce ROS production and cause protein degradation and apoptotic response [293]. This study showed that GSH can stimulate a dose-dependent redox biotransformation of SeNPs to generate ROS in a pure enzymatic system, especially given that GSH is the cell’s most abundant thiol-containing small molecule. Another study indicated that a minimal concentration of 2 μg·mL−1 of biogenic SeNPs inhibits the proliferation of prostate adenocarcinoma cell line, PC-3, by a ROS-mediated activation of necroptosis [295]. Furthermore, Y. Wang et al. reported on the inverse relationship between SeNPs size and ROS production mediated by a GSH system encompassing GSH, GSH reductase (GR), and NADPH [296]. They found that the smaller SeNPs (35 nm) were more active than larger SeNPs (91 nm) in inhibiting in vitro and in vivo cancer cell accumulation through an ROS mediated mechanism [296].
The process of Se-induced apoptosis is associated with Se chemical species and their metabolism, altering some cellular morphologies including nuclear breakdown, chromatin condensation, membrane blebbing, cell rounding, and formation of apoptotic bodies that are eliminated via phagocytosis [257]. It is well-known that apoptotic cascades can originate from intrinsic mitochondrial, extrinsic receptor, or endoplasmic reticulum (ER) stress-mediated signaling pathways [79]. Although different mechanisms are proposed to explain the key role of Se in the cell cycle and apoptosis, the complete process is complex and not yet fully understood. It is correlated with the chemical forms and doses of selenium, and encompasses the activation of caspases, protein kinase signaling, p53 phosphorylation, and ROS generation [79,255,293,297]. It is also known that selenium compounds possess caspase modulation activity, causing programmed cell death. For instance, Se-containing heterocycles, such as 1,2-[Bis(1,2-benzisoselenazolone-3-(2H)-ketone)] ethane (BBSKE), fosters the activity of caspase 3 against tongue cancer Tca8113 cell line [298]; 2,5-Bis(5-hydroxymethyl-2-selenienyl)-3-hydroxymethyl-N-methylpyrrole (D-501036) increases the activity of caspases 3 and 9 [299]; and methyl selenic acid activates caspases 8 and 9 in combination with tamoxifen in both tamoxifen-sensitive and tamoxifen-resistant breast cancer cells [289].
The effect of selenium compounds on caspases, also known as cysteine-aspartic-specific proteases, includes the fragmentation of inter-nucleosomal DNA and the induction of the mitochondrial-dependent/independent apoptosis pathway [79]. The intrinsic mitochondrial pathway is the main process for apoptotic caspase activation in mammals, especially owing to the mitochondrial release of cytochrome c (Cyt-c) that creates an apoptosome complex through the oligomerization with Apaf-1 and procaspase-9 [79]. For example, the Se-containing polysaccharide SeGLP-2B-1 disrupts the mitochondrial membrane potential and enhances the cytosolic Cyt-c levels and the activity of caspases 9 and 3 [300,301].

4.5. Virucidal Activity

Nanobiotechnology has enabled the design of smart molecular diagnosis/treatment approaches for viral infections [302,303,304,305]. For example, synthetic NPs exhibit high antiviral activity and can closely mimic the virus and strongly interact with its virulent proteins due to morphological similarities [306]. In addition, nanostructures can deliver viral antigens in a controlled manner, activate follicular dendritic cells or B cells, antigen cross-presentation, as well as induce humoral/cellular immune responses [307].
SeNPs are found to possess much lower acute toxicity and similar or higher bioavailability when compared to other Se species, such as selenite [308,309,310,311], selenate [312], selenium dioxide [313], Se-yeast [314], selenomethionine [290,315], and methylselenocysteine [47]. In addition, SeNPs have significant antiviral, antibacterial, antiparasitic, and cytotoxic activity [54,316,317,318]. For example, SeNPs decorated with amantadine were able to reverse drug resistance caused by the H1N1 influenza virus infection through the inhibition of caspase-3 activity and suppression of the neuraminidase activity [319]. A similar study demonstrated the superior antiviral capability of β-thujaplicin-decorated SeNPs against H1N1 via the regulation of AKT and p53 signaling pathways [320].
At the onset of the COVID-19 pandemic, selenium was proposed as a weapon of choice to fight against SARS-CoV-2 [321,322,323,324,325,326,327]. For instance, Jin et al. reported that a synthetic redox-active selenium compound, ebselen, is a strong inhibitor of the main SARS-CoV-2 protease that enables viral maturation within the host [328]. This study suggested that high Se intake might hypothetically inhibit SARS-CoV-2 proteases and promote a higher cure rate. Knowing that selenium deficiency is linked to severe virulence, intravenous Se therapy and high-dose selenite pharmaconutrition have been proposed to be effective at reducing the occurrence and the progression of acute respiratory distress syndrome (ARDS), multiorgan failure, and new infections in COVID-19 patients [329]. Overall, human Se levels are crucial in antioxidant, anti-inflammatory, and immune effects in COVID-19 patients; thus, it is important to study the impact of Se excess and deficiency in mitigating COVID-19 symptoms, especially in patients with pre-existing comorbidities or long-term diseases [330].
A unique experimental study used lateral flow immunoassay kits (LFIA) relying on SeNPs modified with SARS-CoV-2 nucleoproteins for the combined detection of anti-SARS-CoV-2 IgM and IgG in human sera, and succeeded in exhibiting recent SARS-CoV-2 infection within just 10 min detectable by the naked eye [331]. In addition, the sensitivity and specificity of the kits were clinically examined with real-time polymerase chain reaction (RT-PCR) tests in COVID-19-diagnosed patients and non-infected controls, amounting to 93.33% and 97.34%, respectively. Finally, there were no cross-reactions with rheumatoid factor and positive sera for influenza A, influenza B, and antinuclear antibodies. Similar studies have also designed point-of-care systems based on SeNPs to detect IgG and IgM against SARS-CoV-2 [332,333]. Although little research has been conducted to highlight their potential in mitigating and controlling COVID-19 pandemic, SeNPs clearly constitute superior detection tools and antiviral nanotherapeutics amenable to containing and/or combating viral outbreaks and pandemics. Figure 6 summarizes the biomedical role of SeNPs in diagnosing and curing viral infections.

5. Pharmacokinetics and Cellular Interactions of Selenium Nanoparticles

Nanomedicine is increasingly offering novel nanoparticle-based technologies for therapy and diagnosis [72,334,335,336]. The extraordinary properties of nanomaterials provide a safe and efficient basis for personalized medicine, which promotes tailored therapies considering the patient’s specific characteristics for the best response and highest safety margin [337,338]. In vitro and in vivo pharmacokinetics, i.e., absorption, distribution, metabolism, excretion, and toxicity (ADME-Tox) studies have been the rule for pharmaceutical organic drugs. Likewise, nanomaterials should be subjected to the same in vitro and in vivo ADME-Tox studies [24,339]. NPs need to reach the targeted organ or tissue to accomplish the desired action with both efficacy and safety. Therefore, nanotherapeutics rely on effective NP cellular uptake and tissue or tumor permeability that both depend on various factors, such as the size, shape, and surface chemistry of the NPs, as well as the biological (micro)environment, the specific location, and the targeted tissue [340,341,342].

5.1. Interaction of Selenium Nanoparticles with Cells and Their Components

The interaction of nanomaterials with cells and lipid bilayers is crucial in several biomedical applications, such as drug and gene delivery [280,343,344,345], diagnostics, phototherapy, and imaging [34,219,346,347,348]. Besides interacting with biological entities (i.e., organs and tissues), NPs cross the cellular barriers and are internalized by cells through endocytosis to accumulate in targeted organs and are eventually cleared [349]. The cell internalization of NPs occurs via several pathways, such as direct penetration, phagocytosis, and pinocytosis [350]. On the other hand, the body internalizes the NPs via inhalation, ingestion, and dermal exposure, to name a few, as depicted in Figure 7 [351]. Although this particular capacity leads to beneficial therapeutic applications, some potential adverse effects regarding NP toxicity have been observed [351]. In consequence, the investigation of the NP interactions with their microenvironment, mainly with other nanomaterials and biomolecules, is crucial to determine the efficacy of nanoscale materials [341,349,352].
Several factors determine the success of the NP uptake and interaction with cells, including the intrinsic NP physicochemical properties, such as the shape, size, coating and morphology, and crystalline structure, in addition to the biological environment characteristics and the transformation of NPs during the test, e.g., transformations owing to the formation and adsorption of a protein layer, known as protein corona [353,354]. The size and shape of the NPs directly affect their cellular uptake rate, which is also related to their residence time in the circulatory system [337]. The NP shape plays a special role in their internalization as it impacts their interaction with the cells. In addition, the NP symmetry controls their trajectory within the body since hydrodynamic forces regulate their transport [355]. For example, several studies concur that spheres are the most effective in terms of cellular uptake due mainly to their isotropic shape, which allows a constant distribution of acting forces and a tendency to remain in the blood flow longer [356,357]. Furthermore, spherical NPs must overcome a minimal membrane bending energy barrier, when compared to their non-spherical counterparts.
The importance of NP-cell interactions has been acknowledged by several authors [352,358,359,360]. However, little research has addressed the quantitative analysis of NP–cell interactions, which is essential to fully understand nano-bio mechanisms and nanotoxicology of cell-surface bound and intracellular NPs [361]. Lately, novel analytical mass spectrometry (MS)-based methods, such as laser ablation inductively coupled plasma mass spectroscopy (LA-ICP-MS), time-resolved ICP-MS (TR-ICP-MS), nano secondary ionization mass spectrometry (nano-SIMS), and mass cytometry that fundamentally combines flow cytometry with time of flight mass spectrometry (ToF-MS), have been introduced [353]. Several quantification studies on the interaction of NPs and cells under different experimental conditions (i.e., type of NPs and cell lines) using SC-ICP-MS for a diversity of applications, such as nanotoxicity, drug delivery research, and optimization of techniques for the green synthesis of NPs, have been developed [362,363,364,365,366]. For instance, Hu et al. quantified intracellular amounts of SeNP uptake by γδT cells using ICP-MS and found that the total uptake amount was ~9 × 10−5 μmol of SeNPs per million of cells [97].
The analysis of NP uptake and biodistribution has gained much importance in recent years, mainly to evaluate effective concentrations of clinically administered NPs. According to the literature, the most remarkable biomolecules interacting with the NP surface are nucleic acids and proteins [367,368]. Nucleic acids are convenient receptors for molecular nano-structures, demonstrating potent synergistic activity due to their mechanical rigidity, physicochemical stability, and high specificity of base pairing [369]. On the other hand, proteins possess various binding sites owing to post-translational modifications in addition to specific and non-specific adsorption capability. All these interactions confer stability, sustained enzymatic activity, and immune-biocompatibility to nanomaterials [370,371,372,373].
The main factors that influence the coupling of NPs with biomolecules or other NPs are known as the interaction drivers, and include van der Waals forces, electrostatic or magnetic interactions, and molecular forces, based on complementarity between nanomaterials, their distance, and their geometry [374]. Furthermore, the interaction between surface molecules greatly depends on the surface functionalization with one or different chemical reactive moieties [375]. For instance, NPs in biological milieus are surrounded by interacting biomolecules that are able to change or saturate their surface, resulting in surface coating modification that, in turn, may alter their unique properties, initially designed functionalities, and desired targeting capabilities, in addition to altering their cytotoxicity and influencing their pharmacokinetic features and accumulation [376,377,378,379].
Surface modification of NPs using functional molecules/particles/polymers can increase their cellular interactions and uptake by tuning their overall properties to fit the targeted applications [380]. Multifunctional NPs have various and different interactions with biomolecules, and are embedded within human proximal fluids inside cells and culture media [337,369]. To achieve this multifunctionality, NPs are bioconjugated with several entities that may include the diagnostic imaging domain, the targeting ligand, and therapeutic moiety to yield multifunctional formulations consisting of therapeutic-loaded NPs, also known as theranostic NPs (Figure 8) [381,382,383]. This constitutes a key step forward in nanomedicine towards personalized medicine with promising applications in drug delivery, cancer treatment, and diagnosis, among others [8,24,337].
For instance, monodisperse and homogeneous spherical SeNPs have been successfully modified with a dinuclear luminescent Ru(II) complex resulting in a multifunctional nanocarrier-based delivery system (NDS) that delivers the siRNA targeting tumor-MDR1 gene in cisplatin resistant adenocarcinomic human alveolar basal epithelial cells (A549 cells) [384]. Another study presented amine-terminated generation 5 polyamidoamine (PAMAM) dendrimers (G5.NH2)-modified SeNPs (G5@SeNPs) for the systemic dual delivery of MDR1 siRNA and cisplatin to down-regulate P-glycoprotein and reverse multidrug resistance [385]. Through gel retardation assay, cellular uptake, and transfection studies, these multifunctional G5@SeNPs are found to enhance siRNA loading, release efficiency, and gene-silencing efficacy. In addition, siRNA- and polyethylenimine (PEI)-modified SeNPs are shown to improve the apoptosis of HepG2 cells [386]. Therefore, all these studies demonstrate that multifunctional SeNPs are effective nanosystems for chemotherapy and gene therapy technology.
When NPs interact with plasma proteins, a protein corona forms on their surface. This has been widely studied due to its significant effects on therapeutic NPs [387,388,389,390]. Indeed, it can considerably modify the NP shape, size, surface charge distribution, and susceptibility to aggregation. Its formation also dictates the subsequent biological fate of NPs within the body [342,359] and modulates various biological behaviors, such as cell-uptake, toxicity and immunogenicity [391]. Since this layer is important in the NP-cell interactions, various studies have investigated the parameters affecting the adsorption of proteins on the NP surface in physiological fluids and the role played by the corona in the mechanism of NP uptake by the cells [392]. As a result, the formation of an individualized protein corona might be useful for targeted therapy and, consequently, for personalized medicine approaches [393].
Chakraborty et al. studied the formation of coronas, consisting of serum most abundant proteins, i.e., human serum albumin (HSA), IgG, and transferrin, to encapsulate SeNPs that had been previously functionalized using one of the following surfactants: cetyltriammonium bromide (CTAB, cationic), sodium dodecyl sulfate (SDS, anionic), and Brij-58 (non-ionic) [394]. As a result, HSA was found to increase the antioxidant capacity of SeNPs, whereas the presence of IgG and transferrin reduced their radical scavenging activity. Moreover, the protein corona formation over functionalized SeNPs enhanced their size and decreased their cellular uptake and subsequent toxicity, except for transferrin-coronated NPs that showed increased uptake and cytotoxicity. In addition, protein coronation is significantly influenced by NP functionalization [395]. In fact, protein quantification and densitometry studies showed that cationic SeNPs (CTAB-SeNPs) promote maximum corona formation and possess higher affinity towards the predominantly negative surface potential of serum albumin. This suggests that binding factors, such as electrostatic forces, attachment via cysteine, hydrogen bonding, and entropy-driven binding, govern the process of protein coronation [395]. Moreover, the study of molecular interactions between biogenic SeNPs, synthesized using yeast extract, and HSA by employing a microwave plasma optical emission spectrometry operating in a single-particle mode documented that the potential biomedical application of SeNPs greatly depends on their surface functionalization and capability to form a protein corona [396].
Polysaccharides (PS) have been widely used as functionalizing agents for nanomaterials, especially due to their unique properties, such as excellent biocompatibility, stability, and biodegradability [397,398,399]. In addition, they increase the residence time of PS-functionalized NPs at the target site and the permeation/bioavailability of loaded drugs or biomolecules [400], enhance the specific interaction with biological targets [401], and improve cell-permeability and cancer-targeting ability [398]. For example, SeNPs decorated with mushroom PS–protein complexes (PSP) achieved remarkably enhanced cellular uptake via endocytosis which, in turn, improved their antiproliferative activity [49]. The capping with PSP was carried out through strong physical adsorption of PS hydroxyl groups and protein imino groups to yield PSP–SeNPs, whose main target organelles are lysosomes on MCF-7 human breast carcinoma cells. Moreover, Spirulina PS (SPS) from the food-grade blue-green microalga Spirulina platensis are known to have an essential biological role in free radical scavenging, DNA repair, immunostimulation, and antiviral effect [402]. Thus, SPS could be employed as a surface decorator of NPs to improve their cell-penetrating capabilities, prolong their circulation time, and prevent plasma protein adsorption. For instance, SPS-functionalization of monodisperse spherical SeNPs enhances their cellular uptake capability and cytotoxicity against various human cancer cell lines, including A375 melanoma cells [403]. A comprehensive survey of Se-carbohydrate combinations including PS-decorated SeNPs is provided by Li et al. [404].

5.2. Key Role of Selenoproteins in the Pharmacological Activity of SeNPs

As mentioned earlier, selenium is a unique trace element which presents a pharmacological behavior by its incorporation into selenoproteins. Several selenoproteins are essential enzymes that include at least one SeCys in their active sites to exert catalytic and antioxidant activities [405,406]. Selenoproteins play several physiological roles [250,407,408,409,410], such as regulating the immune functions [408], protecting against oxidative stress [411], preventing cardiovascular disorders [412,413], regulating thyroid hormone metabolism [414,415], influencing the occurrence of type 2 diabetes [416], protecting against cartilage redox homeostasis and the progression of arthropathies [417], chemoprevention and chemotherapy [406], modulating energy metabolism in neural cells [418], and enhancing male fertility [419]. However, there is little information on the effect of SeNPs on the pharmacokinetics and pharmacodynamics of selenoproteins [73]. For example, SeNPs have a better effect on glutathione S-transferase (GST) activity when compared to selenoproteins regardless of supra-nutritional or toxic levels [290]. Moreover, SeNPs are used to design a safe and effective strategy for a highly therapeutic efficacy of cytokine-induced killer (CIK)-based cancer immunotherapy [35]. Since the safety profile of nanomaterials to a large extent determines their biomedical applications, no induced hemolysis is noticed when SeNPs are incubated with human blood, highlighting their hemocompatibility. In addition, SeNPs are gradually metabolized into selenocysteine (SeCys2), which subsequently regulates the expression of multiple selenoproteins and other metabolites in CIK and cancer cells [35]. This unique strategy enables a CIK-SeNPs co-treatment that induces specific immune responses against tumor progression via the production of natural killer cells and the priming of tumor-associated macrophages. All these findings are helpful in translational medicine towards the development of efficient treatments for diseases associated with Se metabolism [32]. Additionally, SeNPs improve selenium bioavailability and facilitate selenoprotein expression when selenium level is low [127]. Indeed, the encapsulation of selenium in chitosan networks (CTS-SeNPs) likely increases its retention and/or delivery to induce selenoprotein expression and prevent Se-induced damage to DNA.

5.3. Pharmacokinetics of Selenium Nanoparticles

Pharmacokinetics investigates the scale and rate of ADME-Tox of drugs in the body through precise and rigorous experimental methods [24,420]. By nanosizing its formulation, the drug dissolution rate can be increased to promote improved absorption and bioavailability [421,422]. Therefore, NPs are useful to deliver drugs and enhance tissue selectivity due to their selective uptake in specific tissues [24,341,422]. Moreover, nanocarriers alter the pharmacokinetic properties of drugs by enhancing their effectiveness and diminishing their adverse effects [423]. The pharmacokinetic profiles of the parent drug and the drug associated with NPs are often different [424]. Thus, studying the pharmacokinetics and biodistribution of drugs formulated as NPs is essential to comprehend and predict their effectiveness and side effects. For delivery purposes, an optimal theranostic NP model should display suitable release kinetics of the drug in specific concentrations at the target site [24].
The physicochemical properties of NPs are essential for pharmacokinetic modulation as they dictate the immediate pharmacological response in the body following their administration [425]. For instance, the shape, size, surface chemistry (PEGylation, ligand conjugation), surface charge, and composition influence the pharmacokinetics, intracellular penetration, and tumor bioavailability [341]. Moreover, NPs prolong the half-life of drugs in blood circulation, decrease their apparent volume of distribution, and significantly reduce their degradation and clearance. Depending on the method of preparation and the desired therapeutic effect, the drug can be dissolved, adsorbed, attached, entrapped, or encapsulated within NPs [426,427]. Regardless of their compositions, all NPs must possess a reasonable half-life in the blood circulation, selective targetability, and efficacious clearance from the body after the drug delivery to target tissues [428,429,430,431,432]. To obtain the adequate NP pharmacokinetic features for clinical applications, it is required to regulate their hydrodynamic diameter, shape, and surface properties. The general process includes: (1) systemic circulation and reticuloendothelial system (RES) interaction, (2) extravasation and tumor penetration, and (3) interaction with target cells [433]. The specific pharmacokinetic parameters include the volume of distribution, half-life, mean residence time, maximum concentration, bioavailability, permeability, clearance, and area under the time–concentration curve [434].
In the case of selenium, Se metabolic cycling and excretion from the human body include both inorganic and organic molecular species of Se present at different oxidation states including 2−, 2+, 4+ and 6+ [435]. Nevertheless, little research has been conducted to show the formation of elemental or metallic Se (Se0) as part of these metabolic processes [435]. For instance, MeSeCys, an organic Se compound with potential anticancer activity, is found to be an advantageous supplementation of Se due to its lowest toxicity among all the selenocompounds while still being greatly bioavailable [86]. The same study also showed that the positive effects of various selenocompounds on the activity of GPX1 and on the generation of the glycoprotein selenoprotein P (SEPP) do not correlate with their toxicity levels but rather with molecule-specific properties [86]. Further, MeSeCys has minor sub-chronic oral toxicity and no genotoxicity at doses far above the daily nutritional Se level (0.5, 0.7, 0.9 mg per kg of body weight) after 90-day oral exposure [436]. Although the pharmacokinetics of sodium selenite has been investigated in various animal models, the biosafety dose of Se is still unclear due to the narrow safe dose range of Se and the distinct animal physiological and pathological conditions [437]. For instance, selenium-enriched yeast (SeY) had higher bioavailability in rats than sodium selenite, while plasma-free SeMet was found to be the ideal biomarker of SeY status in vivo [438]. Furthermore, rapid absorption and slow excretion of sodium selenate in the blood of healthy piglets were observed; this pharmacokinetic process conforms to the two-chamber open model [439]. Additionally, variations in antioxidant systems in piglets as a function of Se levels were noticed, thereby providing a more complete understanding of risk assessment and clinical application of Se supplementation. On the other hand, the safe dose level of intravenously administered sodium selenite, defined as maximum tolerated dose (MTD), was reported to be 10.2 mg·m−2 in terminal cancer patients and implied the importance of kidney function in the excretion of Se from selenite since the pharmacokinetic results demonstrated a linear increase in plasma Se concentration with respect to total Se dose [440]. Lastly, no apparent adverse effects of high dose of repeated selenite administration on physiological selenium homeostasis were reported [440].
The pharmacokinetic and pharmacodynamic profiles of two high doses of parenteral selenite in patients with systemic inflammatory response syndrome (SIRS) showed that the maximum glutathione peroxidase activity appeared only at very high dose [441]. As a result, the very high dose of 2000 µg (25.30 µmol) of selenium supplied as selenous acid by short-term bolus injection, followed by a continuous daily intravenous infusion of 1600 µg (or 20.24 µmol) for 10 days, was very effective in replenishing serum selenium to physiological levels and safely maximizing the antioxidant activity of the selenoenzyme GPX3. Furthermore, it has been shown that three selenium compounds (sodium selenite, methyl selenocysteine, and seleno-1-methionine) were well-tolerated and assessed safe to be used at 400 µg elemental selenium per day in a study population of twenty-four cancer patients, revealing negligible genotoxicity and minor reductions in lymphocyte counts [442].
Although the above therapeutic applications of Se compounds by conventional introduction may seem effective, the delivery of nano-formulated Se has distinct pharmacokinetic advantages such as specific drug delivery, controlled release, high metabolic stability, high membrane permeability, enhanced bioavailability, long duration of action, less toxicity, and therapeutic efficacy [50,61,74,76,90,443]. SeNPs are convenient for administration since they can be delivered by various routes including oral and intravenous ones [444]. At the nanoscale, selenium is also used to improve the pharmacokinetic properties of drugs. For instance, chitosan-decorated SeNPs constitute an excellent carrier of the therapeutic peptide BAY 55-9837 for type 2 diabetes mellitus by decreasing its renal clearance rate [445]. In addition, insulin-loaded SeNPs (Ins-SeNPs), fabricated using ionic cross-linking/in situ reduction, exhibit enhanced antidiabetic effect through a controlled insulin delivery and an outstanding stability in the digestive fluids [446]. Based on ex vivo intestinal imaging and cellular internalization, the transepithelial transport ability of SeNPs to overcome the absorption barrier was assessed, and it was found that Ins-SeNPs not only get into the cytoplasm, but also enter the nuclei. Finally, Ins-SeNPs exhibit excellent hypoglycemic effects after oral administration, requiring a lower oral dose to achieve a long-acting glycemic reduction than insulin alone for the same time and concentration (0.75 μg·mL−1) [446].
In addition, Arg-Gly-Asp (RGD) peptide-decorated and doxorubicin-loaded selenium nanoparticles (RGD-SeNPs) targeting tumor vasculature significantly enhanced the cellular uptake and antiangiogenic activity of SeNPs in vitro and in vivo on human umbilical vein endothelial cells (HUVEC) [447]. The as-designed nanosystem led to bioresponsive triggered drug doxorubicin (DOX) release by disassembly under acidic conditions with the presence of lysozymes and cell lysate [447]. Another study reported on the hybridization of SeNPs with niosomes as lipid nanoparticles (NISM-B@SeNPs) to open a new approach in drug delivery for cancer treatment studies that potentially exhibit good in vivo biocompatibility [343]. Moreover, Se-functionalized liposomes (SeLPs) were developed as a DOX delivery vehicle to prolong the systemic circulation of liposomes by in situ selenium coating and enhance the anticancer effect via the synergy between DOX and Se [339]. A dual-loaded nanocarrier system of the antiretroviral drug Etravirine (TMC-125) and SeNPs was fabricated to evaluate the NP biodistribution in potential human immunodeficiency virus (HIV) reservoirs in vivo in Sprague Dawley rats [448]. The in vivo pharmacokinetic study showed the controlled release potential of the nanocarrier along with high stability, prolonged on-target residence time, low clearance, and a higher accumulation of the dual-loaded nanocarrier in remote HIV reservoir organs such as the brain, ovary, and lymph node.
A nanoselenium-coating biomimetic cytomembrane nanoplatform (BMMP) was prepared as a drug nanocarrier by using P. geniculate for manganese and doxorubicin codelivery and mitochondrial targeted chemotherapy [449]. The pharmacokinetic studies revealed that approximately 50% of BMMP-Mn2+/Se/DOX was retained in the body at 8 h post-injection, implying that BMMP-Mn2+/Se/DOX had a long blood circulation time. In addition, the nanoplatform exhibited excellent biosafety and exerted long-acting effects on tumors, being completely excreted from the body at 96 h post-injection with no obvious side effects from its residue [449]. Se-coated nanostructured lipid carriers (SeNLCs) of around 160 nm were developed for improving the oral bioavailability and the curative effect of berberine, an antidiabetic phytomedicine [450]. The results showed that the berberine-loaded SeNLCs (BB-SeNLCs) had an entrapment efficiency of 90% and greatly enhanced the oral bioavailability of berberine, which was approximately 6.63-fold higher than that of berberine solution. From the abovementioned cases, it is evident that SeNPs are emerging as valid pharmacological tools for further in vitro and in vivo studies on conjugates with drugs, drug candidates, targeting agents, and molecular probes.

6. Green Nanotechnology: A Better Avenue for SeNP Bioapplications

Despite tremendous advances in the application of nanotechnology in the diagnosis and treatment of different diseases, several challenges are yet to be adequately addressed, such as bio- and cyto-compatibility, as well as selectivity and efficiency of NPs [451]. The conventional synthesis of nanomaterials often involves the use and/or the generation of toxic/harmful reagents and substances (i.e., solvents, catalysts, and reducing and capping agents), which affect the environment and patient response. Moreover, bringing nanomaterials from the laboratory to clinical or industrial applications has been slow and challenging due to the poor understanding of the new potential hazards introduced by nanotechnology and the absence of suitable policies to manage emerging risks [452,453,454,455]. Hence, bionanotechnology may hold the solution and constitute a sustainable alternative by offering better and safer processing methods for NP production [65,171,456,457,458,459,460,461]. This growing approach aims at exploring the capabilities of natural, widespread, and renewable resources as part of the starting reagents in the NP production process to eliminate or, at least, reduce the NP hazards to the environment and human health, and, ultimately, to substitute the existing toxic reagents/products with new environmentally friendly products that are benign, eco-friendly, sustainable, biocompatible, and safe [462,463]. Indeed, organic, non-toxic stabilizing agents lead to controlled sizes and shapes, tailored biological responses (e.g., cytotoxicity, inflammation), and enhanced biodistribution of biogenic NPs [451].
Green nanomaterials have demonstrated potential applications in medicine as anticancer, antidiabetic and antioxidant agents, as well as for bio-sensing purposes [464]. For instance, green drug delivery nanosystems display efficient targeted recognition and controlled release, high biocompatibility, and decreased toxicity [465]. Thus, this has triggered a sustained demand for green nanotechnology-driven drug delivery systems fueled by significant developments of diverse delivery devices, such as inorganic NPs, quantum dots, polymeric NPs, dendrimers, nanostructured lipid carriers, solid lipid NPs, etc. [466]. Further, the rise of green nanomaterials is accompanied by sustainable, low-energy, and low-cost procedures for the manufacturing of different tissues that diminish the consumption of toxic materials [467].
Green approaches produce highly stable and biocompatible SeNPs that increase efficiency and minimize side effects [155,468]. Indeed, some metabolites present in plant extracts, such as polyphenols, saponins, vitamins, carbohydrates (including polysaccharides), flavonoids, alkaloids and tannins, are excellent reducing and capping agents of SeNPs that make them safer, more stable [469], and suitable for several potential applications in biomedical sciences [42,60,61,62,63,470]. A recent survey comprehensively addresses the biosynthesis of SeNPs using a variety of plant extracts relating the properties of the synthesized nanostructures (composition, size, shape, stability) with the conditions used for this green route (temperature, time, Se precursor, and extract concentration) [471]. Since biosynthesized SeNPs are less polydisperse and do not aggregate under physiological conditions, they have emerged as effective tools in medical and pharmaceutical sciences to treat different diseases [61,472]. In addition, eco-friendly metallic NPs are safely translated in medicine and serve as safe nanotheranostic agents/platforms [468,473,474]. However, some important concerns need to be considered prior to the use of SeNPs in clinical translational applications, such as the safety profile, pharmacokinetics and pharmacodynamics, and specificity and sensitivity in the biological milieu.

7. Biomedical Applications of Biogenic Selenium Nanoparticles

SeNPs are bioactive entities that might be easily made biologically available to play a crucial role in many oxidoreductive processes [91]. In addition, SeNPs possess a regulative effect to support the correct functioning of the body and offer outstanding health benefits to treat/cure various diseases [32,58,475,476]. The following sections detail the use of biogenic SeNPs for different therapeutic purposes, including anticancer, antimicrobial, and anti-diabetic activities, in addition to gene and drug delivery.

7.1. Antioxidant Activity

Antioxidants are compounds that prevent the generation of free radicals as well as scavenge them when produced during various biochemical reactions in animals and plants, therefore playing a significant role in protecting against oxidative stress and neurodegenerative and cardiovascular diseases [58,477,478]. The antioxidant potential of biogenic NPs relies on the redox potential of phenolic and flavonoid compounds present at their surface [128,479,480]. Since selenium is implicated in antioxidant defense systems and significantly contributes to maintaining the redox homeostasis [481,482,483,484], SeNPs display a protective activity against cellular damage [167,485,486]. To date, an overwhelming number of studies indicate that SeNPs possess great antioxidant ability and free radical scavenging efficiency that potentially protect tissues and cells from oxidative damage [160]. Moreover, SeNPs exhibit antioxidant activity with less toxic effects than zero-valent selenium (Se0) [51] or sodium selenite [155]. In addition, biogenic SeNPs exhibit a higher antioxidant activity with less toxicity to healthy cells than selenium dioxide [260]. It has also been proved that SeNPs, fabricated using Theobroma cacao L. bean shell extract as the reducing and capping agent, are highly stable and possess a better antioxidant activity than the extract itself [487]. Similar behaviors of SeNPs, fabricated using different natural extracts, have been reported [160,488,489,490]. However, Ephedra aphylla extract displays a better antioxidant activity than SeNPs produced using the same extract owing, most likely, to the presence in the extract of phenolics, flavonoids, and tannins in greater amounts [491].
Biogenic SeNPs exhibit numerous biological properties. For instance, SeNPs of 5–200 nm can directly scavenge free radicals in vitro [492]. In addition, their size has a considerable effect on their antioxidant properties since smaller NPs are more efficient in capturing free radicals [262]. This was corroborated by other studies that also established the dependency of antioxidant activity on particle concentration [109,490,493]. Additionally, the capping moieties, such as quercetin and gallic acid, play an important role in the antioxidant potential of ecofriendly SeNPs [494], and of bimetallic Ag-Se NPs [495]. Table 3 summarizes the antioxidant activity of biogenic SeNPs, obtained via different methods.
The most used in vitro technique to measure the free radical scavenging activity (RSA) of SeNPs is 2,2-diphenyl-1-picrylhydrazyl (DPPH) as it is simple, rapid, facile, sensitive, and stable [249]. It is based on the reduction of methanolic DPPH solution by donating an electron or hydrogen atom to form a non-radical, stable, and pale yellow/colorless molecule: 2,2-diphenyl-1-hydrazine [316,490].
Several studies explored the strong antioxidant activity of biogenic SeNPs synthesized by Lactobacillus casei in oxidative stress-caused intestinal epithelial barrier dysfunction [496,497,498]. As a result, SeNPs can alleviate ROS mediated mitochondrial dysfunction via Nrf2-mediated signaling pathway, increase the number of goblet cells, reduce the production of ROS, increase GPX activity, and preserve the mitochondrial functions. Therefore, SeNPs could be employed to treat oxidative stress-related intestinal disorders. Similarly, biogenic SeNPs, mainly coated by polysaccharides, exhibit antioxidative and anti-inflammatory effects in protecting intestinal epithelial cells against H2O2 and ETEC K88-caused injury, and maintaining the intestinal epithelial barrier integrity [499]. This amount of work has led to a consensus stating that biogenic SeNPs present significant antioxidant activity and may serve as a potential antioxidant supplement or ingredient [155,488] and neuroprotective agent [500].

7.2. Antimicrobial Activity

The antimicrobial capability of SeNPs stems from their large surface to volume ratio, which allows them to set better contact with microorganisms, thus leading to improved antimicrobial activity [147]. Therefore, SeNPs can be used in several fields including infectious control, surface treatment of biomedical instruments, pharmaceutical industry, cosmetics, and food manufacturing [155].

7.2.1. Antibacterial Activity

The antibacterial activity of selenium compounds is attributed to the generation of free radicals, including Se oxyanions [501]. These novel active products might constitute a solution to the emerging drug-resistant microorganisms that are considered to be a great current health concern [62,502,503,504]. The unique antibacterial effect of biosynthesized SeNPs has been extensively explored on the basis of morphological and structural changes in the bacterial cells [160,316,490,491,495]. For instance, phytofabricated SeNPs using the aqueous fruit extract of Emblica officinalis were found to possess antimicrobial activity against both Gram-positive and -negative bacteria and fungi [155]. The minimal inhibitory (MIC) and bactericidal (MBC) concentrations were 9.16 ± 0.76 and 19.83 ± 1.25 μg·mL−1, respectively, against S. aureus, and 59.83 ± 2.56 and 97.50 ± 3.27 μg·mL−1, respectively, against E. coli. A similar study provides the MIC of manufactured SeNPs using aqueous extract of fermented lupin against Acinetobacter calcoaceticus (2.343 μg·mL−1) and S. aureus (1.171 μg·mL−1) [149]. In addition, the activity of the same SeNPs against the fungus Aspergillus was strain-selective, as they were effective only against A. flavus while no activity was shown against A. niger.
The NP antimicrobial activity is size-dependent since tiny NPs can easily cross the cell wall and membrane and provoke cell lysis [93,114,155]. For instance, 221.1 nm SeNPs, fabricated using S. maltophilia, exhibit a strong antimicrobial activity with an effective concentration (EC50) of 26.32 mg·L−1 against E. coli, 7.59 mg·L−1 against S. aureus, and 62.37 mg·L−1 against P. aeruginosa [505]. In addition, the large surface area, small size, and spherical shape are probably responsible for the good antimicrobial activity of SeNPs fabricated using the supernatant of Lysinibacillus sp. against E. coli and S. aureus [68]. Moreover, the anti-biofilm activity of different concentrations (0–2 mg·mL−1) of the same biogenic SeNPs against the strong-biofilm producer P. aeruginosa was highlighted.
Besides the dimensions, other important features, such as the elemental structure (purity) and shape of SeNPs, should be considered when studying their antibacterial activity [155]. For example, the antimicrobial and antibiofilm ability of SeNPs, manufactured using S. maltophilia SeITE02, against different pathogenic bacteria seems to be strictly linked to their organic surrounding cap [62]. This characteristic was tested by exposing NPs to progressively stronger protocols that denature their external organic coating; this resulted in increased MIC values with progressive denaturation. Moreover, SeNPs synthesized using C. bulbosa tuber aqueous extract were found to promote inhibitory effects on the growth of certain clinical pathogens, such as B. subtilis and E. coli, as well as a strong larvicidal activity against the dengue vector, Aedes albopictus, with 250 μg·mL−1 as the maximum lethal concentration [506]. In this study, the mortality rate of A. albopictus larvae is caused by the adhesion and penetration of SeNPs across cell membranes to further hinder the function of membrane proteins. On the other hand, biofilm inhibition assays using biogenic SeNPs resulted in a gradual decline of biofilm thickness at 25 μg mL−1 (80–70% reduction) and 50 μg·mL−1 (extreme reduction), thus inhibiting any further bacterial proliferation (Figure 9) [109]. Additionally, SeNPs, derived from B. licheniformis, inhibit S. aureus adherence and microcolony formation on polystyrene, glass, and catheter surfaces [507]. These findings are corroborated by another study showing that SeNPs, fabricated by the whole cell lysate of Bacillus sp., inhibited the biofilm formation by S. aureus, P. aeruginosa, and P. mirabilis by 12.42%, 34.30%, and 53.40%, respectively.
Owing to their interactions with DNA and proteins, biogenic SeNPs totally inhibit S. aureus growth at a 300 μM concentration within 24 h [508]. Spherical SeNPs of 40 to 120 nm in size, fabricated with the non-pathogenic bacterium Ralstonia eutropha, were found to almost totally inhibit (99%) the growth of P. aeruginosa, S. aureus, E. coli, and Streptococcus pyogenes at different concentrations ranging from 10 to 300 μg·mL−1 [55]. In contrast, the growth inhibition of the pathogenic fungus A. clavatus requires 500 μg·mL−1 of the same SeNPs. Moreover, the antimicrobial activity of SeNPs, synthesized by Zingiber officinale root extract, against Gram-positive (S. aureus) and -negative (E. coli, Klebsiella sp., Pseudomonas sp., Serratia sp., and Proteus sp.) bacteria resulted in MIC values of 150–500 μg·mL−1 [501]. On the other hand, SeNPs, produced using E. coli, P. aeruginosa, S. aureus, and methicillin-resistant S. aureus (MRSA), are shown to inhibit the bacterial growth by, most likely, altering the bacterial growth cycle by impacting the synthesis of RNA, enzymes, and/or other molecules involved in the cell division [133]. In addition, these SeNPs may cause a systemic failure of the bacterial metabolism leading to cell death owing to induced ROS generation. Similar results are obtained with biogenic SeNPs, synthesized via a plant-mediated process. This is the case, for instance, of SeNPs produced using Cinnamomum zeylanicum bark extract, which show a biocidal activity against several bacterial foodborne pathogens (E. coli, Salmonella typhimurium, S. aureus, and Listeria monocytogenes), and display a potential edible coating basement [509]. Lastly, interesting antibacterial activities are also obtained using SeNPs fabricated by the aqueous extract of cow urine [510].

7.2.2. Antifungal Activity

Several investigations have reported the antifungal activity of biogenic SeNPs. For example, mycosynthesized spherical SeNPs show an antifungal activity against Pyricularia grisea, Colletotrichum capsica, and Alternaria solani, in addition to the inhibition of the sporulation of P. infestans [147]. Moreover, SeNP-enriched probiotics (L. plantarum and L. johnsonii) combined with extracellular metabolites inhibit the growth of the potent urogenital pathogen C. albicans [511]. Additionally, selenium dioxide in culture supernatants enhances the production of soluble metabolites involved in killing the same yeast. Furthermore, the size and crystallinity of chitosan-stabilized SeNPs greatly influence their synergistic antifungal effect against C. albicans biofilms in a dose–response manner [113].
SeNPs, produced using the fusarium T. harzianum, exhibit excellent antifungal activity and, more specifically, a dramatic deactivation of several synthetic genes (FUM1, PA, TRI5, and TRI6) and toxins of the fungi Alternaria (83% TeA and 79% AOH reduction) and Fusarium (63% FB1 and 76% DON reduction), opening them new avenues as bifunctional nanomaterials for the biocontrol of phytopathogens and mycotoxins in agriculture and food safety [512]. On the other hand, SeNPs, fabricated using the plant extract of E. aphylla, display potent antimicrobial activity against several bacterial and fungal species with an enhanced inhibition zone diameter ranging from 19 to 39 mm [491]. In addition, biosynthesized SeNPs using the leaves of Diospyros montana show the highest inhibition zone against A. niger (12 mm) when compared to the tested bacterial species using a disc diffusion method at different concentrations (10, 20, 30 and 40 μg·mL−1) [490].

7.3. Antiparasitic Activity

Biogenic SeNPs show effective and accurate prophylactic effects on acute toxoplasmosis, thereby offering a potential alternative to pyrimethamine and sulfadiazine, whose treatment presents serious side effects [513]. In addition to their strong larvicidal activity against the dengue vector mosquito, A. albopictus [506], biogenic SeNPs also exert in vitro and in vivo antiparasitic activity against Leishmania major, rendering these nano-objects into novel therapeutic agents for the treatment of localized lesions typical of cutaneous leishmaniasis [514].

7.4. Anticancer Activity

Cancer remains a major global health challenge and embraces nearly 100 types characterized by an uncontrolled division of abnormal cells with the ability to metastasize to other parts of the body [515]. Over the previous decades, nanotechnology has gained much importance as a revolutionary approach to combat cancer owing to the unique properties of NPs, such as their size, shape, large surface-to-volume ratio, tunable surface chemistry, and the ability to encapsulate/carry and deliver various drugs, that confer to them many advantages over their bulk counterparts [516]. A fuller understanding of nano–bio interactions should lead to safer and more efficacious nanotherapeutics by overcoming the physiological barriers posed by the tumor microenvironment, which will eventually facilitate the corresponding clinical developments [517]. On the other hand, green nanomaterials are currently intensively screened for the treatment and diagnosis of cancer owing to their high biocompatibility and effectiveness, among which biogenic SeNPs hold great promise [518,519,520]. Building on previously published data [85], Table 4 provides an update and indicates the anticancer activity displayed by a number of biogenic SeNPs against different cancer cell lines.
For instance, biogenic SeNPs are better anticancer, nontoxic, and biocompatible operators than selenite and selenate compounds [50]. Moreover, B. licheniformis-derived SeNPs emerge as the safest form of selenium supplementation with potent necroptosis activity against LNCaP-FGC cancer cells, without affecting red blood cell integrity [524]. Similar results were obtained against the PC3 prostate adenocarcinoma cell line (Figure 10) [295]. Streptomyces bikiniensis-derived SeNPs demonstrate excellent in vitro anticancer effect against Hep-G2 and MCF-7 malignant cells through a hypothetical mechanism consisting of the increased mobilization of endogenous copper (possibly chromatin-bound copper) of cancer cells and the subsequent pro-oxidant action [521]. Another proposed anticancer mechanism of biogenic SeNPs fabricated using Idiomarina sp. PR58-8 is based on the activation of apoptotic pathways by an increased expression of pro-caspase 3 [526]. This is corroborated by the expression of poly(ADP-ribose) polymerase (PARP), and the cleavage of PARP since the activated form of caspase 3 catalyzes the cleavage of pro-PARP, consistent with the role of caspases, a family of intracellular proteases essential in the initiation and execution of apoptosis or programmed cell death through proteolytic cleavage [536]. A similar study reports that biosynthesized SeNPs in combination with X-rays are involved in caspase 3 activation and downstream targets that inhibit the proliferation of lung cancer cells with high cytotoxic effect [523].
Biogenic SeNPs exert their anticancer activity in a dose-dependent manner [491,525,537] and seem to be less cytotoxic to non-malignant cells when compared to their analogs obtained via chemical routes [522], while their surface functionalization has no impact on their anticancer activity nor on the induction of cell cycle arrest [531]. Besides the cytotoxicity pathways discussed above, Se-induced cytotoxicity against malignant cells might also be the result of ROS generation, as in the case of injection of biogenic SeNPs into the abdominal cavity of mice following their inoculation with highly malignant H22 hepatocarcinoma cells [294]. This is corroborated by another study where biogenic spherical SeNPs, synthesized using Asteriscus graveolens leaf extract, induce ROS overproduction and mitochondrial membrane potential (MMP) disruption, thus evidencing their high antitumor activity against HepG2 cells through the activation of apoptosis pathways (Figure 11) [285]. Via another mechanism, biogenic SeNPs, produced using hawthorn fruit extract, trigger increased apoptosis rates in HepG2 cells by up-regulating caspase 9 and down-regulating Bcl-2 [529].
Overall, nanostructured Se compounds have significant potential in the fight against cancer due to their chemopreventive activity, antioxidant/pro-oxidant activity as modulators of ROS suppression/production, capacity to modulate inflammatory processes, apoptosis, capacity to inhibit cancer metastasis, selective targeting of tumors over healthy adjacent tissue, and, last but not least, capacity to inhibit multidrug efflux pumps and thereby counteract tumor resistance to established chemotherapeutic agents and inhibit cancer metastasis. These emerging insights concerning Se-based small molecules formulated as NPs and quantum dots open the way towards new anticancer therapeutics, or for adjuvants designed to overcome drug resistance associated with current chemotherapeutic protocols used routinely in clinical oncology [538].

7.5. Protective Role of Selenium Nanoparticles against Drug-Induced Toxicity

Notwithstanding the widespread application of chemotherapeutic drugs in clinical tumor treatment, serious toxicity, dose-dependent side effects, and non-specific targeting restrict their therapeutic efficacy [539]. Thus, SeNPs, alone or in formulation, have been considered as a potent agent preventing adverse chemotherapy due to their high bioavailability and low toxicity [90]. For example, SeNPs induce a significant tumor cell apoptosis and an impressive enhancement of the therapeutic effect of irinotecan by a selective modulation of Nrf2-ARE pathway in tumor and healthy tissues [444]. Another report demonstrates that SeNPs, at daily doses of 1 mg per kg body weight, efficiently alleviate bone toxicity caused by the intake of anastrozole—a breast cancer drug—therefore preventing the occurrence of osteoporosis in ovariectomized female SD rats [540]. This study also highlights a process of ossification and mineralization in the femurs of SeNP-treated groups, which probably can be explained by the NP antioxidant and protective action. In addition, biogenic SeNPs protect against adverse effects of antibiotics. For instance, SeNPs, produced using the bacterium Pantoea agglomerans, exhibit a protective role in immunological and oxidative stress generated by enrofloxacin (EFX), a broad-spectrum antibiotic, in broiler chickens at a dose of 0.6 mg kg−1 of feed [541]. Furthermore, SeNPs display protective effects against gentamicin-induced nephro- and hemo-toxicity in female Swiss albino mice [542], and hexavalent chromium-induced thyroid damage in male rats [103].
Cisplatin (CIS) is a commonly used alkylating agent to treat testicular, ovarian, head, and neck cancers, among others [543]. Despite its wide clinical applications, CIS triggers many side effects, including the obstruction of some cellular processes, e.g., DNA replication and transcription by inducing DNA adducts and establishing DNA cross-links [544]. Several articles provide insight on the protective effects of SeNPs against CIS toxicity [385,545,546]. For example, SeNPs possess a strong antioxidant potential to prevent CIS-induced gonadotoxicity [544]. Moreover, SeNPs surface-functionalized by 6-hydroxy-2,5,7,8-tetramethylchroman-2-carboxylic acid (Trolox) (Se@Trolox) block cisplatin-induced ROS accumulation [547]. In addition, SeNPs decorated with amantadine (Se@AM), remarkably, prevent caspase 3 activation and decrease ROS levels to inhibit the ability of the H1N1 influenza virus to infect host cells, thereby overcoming the emergence of drug-resistant viruses [319]. Furthermore, SeNPs display protective effects in the progression of diabetic nephropathy (DN) by increasing the levels of heat shock protein (HSP-70) [548]. Lastly, SeNPs increase the number of neutrophils and prolong their survival duration in healthy sheep, as evidenced through thiobarbituric acid reactive substances (TBARS) assay [549].
Biogenic SeNPs, synthesized using Terminalia arjuna leaf extract, show protective and antigenotoxic effects against arsenite (As3+)-induced genetic damage in isolated human lymphocytes [102]. Selenium has been recognized as an effective chemo-protectant against the toxicity of cadmium, exposure to which produces high ROS levels, regarded as Cd-induced neurotoxicity and nephrotoxicity [550]. For example, SeNPs, sodium selenite, and yeast-Se diets display different protective proficiency in Cd-induced testicular damage by improving the expression and synthesis of selenoproteins via the regulation of numerous related transcription factors [551]. Another study ascribes the chemoprotective effects of SeNPs against neuro- and nephro-toxicity of subchronic exposure to CdCl2, mainly, to the inhibition of lipid peroxidation and regulation of genes encoding numerous detoxifying and antioxidant enzymes [550].
T. harzianum-derived SeNPs exhibit high protective effects in infected maize and pear with Fusarium and Alternaria as the main parasites, suggesting that these NPs may be applicable as bifunctional nanomaterials for biocontrol of phytopathogens and mycotoxins in agriculture and food safety [512]. Furthermore, biologically synthesized SeNPs using the terrestrial actinomycete S. griseobrunneus, at 64 μg·mL−1, can eliminate, under UV irradiation, 94% of diclofenac through hydroxylation, oxidation, and decarboxylation [228].

7.6. Anti-Inflammatory Activity

Biogenic SeNPs, synthesized using Trachyspermum ammi, exhibit anti-rheumatic and immunomodulatory properties in arthritic Balb/c mice, as a SeNP treatment reduces paw edema along with decreased lymphocytic cellular infiltration in liver, kidney, and spleen specimens, as well as improved redox state of inflamed synovium [552]. Moreover, SeNPs, dispersed in phytochemical P-coumaric acid, exert an anti-inflammatory activity by modulating catalase, GPX1, and COX-2 gene expression in a rheumatoid arthritis rat model [553]. In addition, biogenic SeNPs, synthesized using L. casei, possess strong antioxidant and anti-inflammatory activity to effectively protect human colon epithelial cells against H2O2-induced injury [498]. SeNPs are a promising anticonvulsant agent due to their potent antioxidant, anti-inflammatory, and neuromodulatory activities against pentylenetetrazole (PTZ)-mediated epileptic seizures in mice hippocampus [554]. Importantly, SeNPs are more effective than sodium selenite in terms of antioxidant and anti-inflammatory activity against induced eimeriosis in the jejunum of mice; therefore, they could be applied for immunoregulation purposes [555]. The potential anti-inflammatory activity of biogenic SeNPs may be ascribed to their down-regulation of pro-inflammatory genes and mediators (e.g., TNF-α, PGE2, TBAR and NOx) and/or to their further antioxidant activity [556]. Additionally, SeNPs ameliorate the health state of rats with streptozotocin-instigated brain oxidative-inflammatory stress and neurobehavioral alterations by regulating the molecular markers of oxidative stress and tissue damage: Nrf2, caspase 3, and parvalbumin proteins [557].
Macrophages play a vital role in chronic inflammatory diseases (CIDs), and thus regulating their activity is crucial in detecting and reducing chronic inflammation [558]. SeNPs, decorated with Ulva lactuca polysaccharides, display anti-inflammatory effects, and relieve the symptoms of acute colitis through the inhibition of the hyper-activation of NF-κB in colon tissues and macrophages [559]. Biogenic SeNPs lower the amounts of H2O2 produced by the pro-inflammatory-activated macrophages in addition to selectively targeting, imaging, and killing pro-inflammatory-activated ones under photodynamic treatment [560]. Further, SeNPs, stabilized by sulfated Ganoderma lucidum polysaccharides, inhibit the inflammation caused by over-activated macrophages in Raw 264.7 cells in a dose-dependent manner [561].

7.7. Antidiabetic Activity

Diabetes mellitus (DM) is defined as a group of metabolic disorders characterized by the decrease in insulin secretion by pancreatic islet cells leading to high blood glucose levels (hyperglycemia) [562]. Diabetes is classified into two major types: type 1 diabetes mellitus, or insulin-dependent diabetes, and type 2, or non-insulin dependent diabetes. Type 1 diabetes mellitus causes a deficiency of insulin due to autoimmune or genetic disorders, while type 2 diabetes (T2D) generates an insulin resistance or reduced insulin sensitivity as a result of inappropriate diet or lack of physical activity [563]. Advances in nanotechnology, molecular and biomedical imaging tools, and drug delivery systems are offering new opportunities for early diagnosis and monitoring disease progression in patients with type 1 or type 2 diabetes combined with diminished insulin secretion [562].
Several articles have thoroughly explored the association between selenium concentration and diabetes [564,565,566,567,568,569]. Patients with DM are often affected by oxidative stress, requiring more antioxidant species to reduce the oxidative and inflammatory response [570]. Since selenium is known to possess excellent antioxidant and anti-inflammatory effects against DM [571,572], several studies have investigated the pivotal therapeutic role of SeNPs in alleviating most diabetic complications and insulin resistance [53,140,446,573]. Owing to their antidiabetic potency, SeNPs can increment insulin secretion by preserving the pancreatic β cell integrity, repressing oxidative stress, inducing glucose depletion, and inhibiting pancreatic inflammation [574]. For instance, biogenic SeNPs, fabricated using Hibiscus sabdariffa (roselle plant) leaf extract, significantly decreased the oxidative stress indicators of testicular tissue in streptozotocin (STZ)-induced diabetic rats, such as nitric oxide and lipid peroxidation [168]. These findings indicate that these nanocrystals hold great promise in attenuating oxidative damage induced by diabetes. Similar research reveals that the administration of SeNPs prepared with Catathelasma ventricosum polysaccharides (CVPS-SeNPs), remarkably, ameliorate body weight, blood sugar level, antioxidant enzyme activities, and lipid levels in STZ-induced diabetic mice, highlighting, thus, their dramatic antidiabetic activity [140]. Although a synergistic effect of CVPS-SeNPs and vitamin E regarding this antidiabetic activity is suggested, the underlying mechanism is not yet clearly understood.
CTS-SeNPs conjugated with a novel peptide, consisting of a recombinant pituitary adenylate cyclase-activating polypeptide (PACAP)-derived peptide DBAYL capable of specifically activating the vasoactive intestinal peptide receptor 2, also known as the VPAC2 receptor, that influences glucose-dependent insulin secretion, enhances insulin sensitivity, hyperglycemia, and lipid profiles, thus demonstrating the potential of this assembly to become a long-acting anti-T2D therapeutic [573]. Moreover, chitosan-decorated SeNPs (CTS-SeNPs) are used to treat T2D by prolonging the in vivo half-life of the therapeutic peptide BAY 55-9837 and slowing its renal clearance rate, proving that BAY 55-9837-loaded CTS-SeNPs possess a desirable sustained-release profile and high stability that could enhance the half-life of low-molecular-weight therapeutics by increasing their apparent molecular size [445]. In combination with metformin (MF), CTS-SeNPs are also effective in the treatment of T2D by mitigating the diabetic complications in a better way than a monotherapeutic approach, and considerably restrict the T2D-induced sperm abnormalities, such as reduced sperm motility, diminished levels of sexual hormones, testicular oxidative damage, and steroidogenesis-related genes dysregulation [575]. Other studies detail the antidiabetic effect of CTS-SeNPs as a monotherapy or as part of a combined therapy with drugs able to swiftly decrease blood glucose and insulin levels [576], in addition to nanohybrid systems to treat diabetic wound infection at mild stage [577].

7.8. Diagnostic Applications

Nanotechnology has led to the development of various NP formulations for diagnostic applications, thereby revolutionizing treatment strategies of relevant diseases, whose outcomes depend on early and accurate detection, such as cancer [25,336,426,578,579], diabetes [580,581,582], gastrointestinal disorders [583,584,585], infectious diseases [586,587,588,589], and neurodegenerative disorders [590,591,592]. Imaging and point-of-care technologies are two specific fields that could benefit from the utilization of NPs [593,594]. In the case of imaging (vide infra), NPs appear positioned to play a key role in the future of medical diagnostics due to their many advantages over the conventional contrast agents, including high affinity binding, specific molecular targeting abilities, controlled biological clearance pathways, and prolonged residence time, thereby providing a longer time for imaging with even multimodal and stimuli-responsive attributes [595,596,597]. However, nanodiagnostics remain useful in very limited clinical situations due to complex demands on pharmacokinetic activity and clearance [595]. The application of SeNPs is of particular interest as they have high photoconductivity, piezoelectricity, thermoelectricity, and spectral sensitivity properties [210,219,598]. For instance, optical and photoluminescence properties of SeNPs can be exploited in the fabrication of nanosensors and imaging markers, eliminating the requirement for additional fluorescent tags, such as proteins or dyes [219].

7.8.1. Detection, Biosensing and Diagnostics

Before addressing the specific place of nanoscale Se in detection and biosensing, a brief outline of this emerging area is in order. Nanobiosensors have witnessed tremendous developments resulting in innovative and sophisticated devices, due to the ever-increasing demand to efficiently and reliably sense a great variety of molecules at low concentrations with high specificity and selectivity [464,599,600,601,602,603,604]. Compared to bulk materials, nanomaterials possess large surface-to-volume ratios, which enable them to provide a greater surface to anchor biomolecules of interest [605,606,607]. In addition to this feature exploited in achieving a high-density immobilization of bioreceptors, nanobiosensors benefit from the unique NP properties to enhance biological signaling and transduction mechanisms since these nanostructures catalyze the bioreactions, mediate the electron transfer, amplify the mass change, and refine the refractive index changes [599,608,609].
Conventional procedures to immobilize enzymes include physical adsorption, affinity labeling, covalent cross-linking, and entrapment [610,611,612,613,614,615]. However, these methodologies are often multistep procedures and present different shortcomings such as non-specific adsorption, leakage, and/or partial denaturation of immobilized biomolecules [616,617]. Aiming to tackle these challenges, approaches based on nanostructured platforms offer exceptional benefits in enzyme immobilization including high binding capacity, high catalytic activity, long operational and storage stability, high specific surface area for volume-efficient catalysis, low protein unfolding, and minimal mass transport limitations [618,619,620]. Nano-immobilized enzymes can be tailored into diverse sizes and shapes without employing toxic reagents [621].
In the above-described beneficial context, biogenic NPs, including those based on Se, offer significant advantages over their analogs synthesized by conventional methods, such as a better stability of up to several months that leads to simple, rapid, nontoxic, cost-effective, and handy sensing strategies [42,622,623,624,625]. Indeed, different proteins and biomolecules present in the reaction medium bind to the surface on the NPs, preventing their aggregation or flocculation, and conferring to them long-term stability [626,627,628]. Furthermore, green-chemistry techniques can potentially improve biosensing applications, such as transducers or electroactive labels, especially in NP-based electrochemical detection systems [629,630]. Thus, the emerging greener biosensors can be relevant for point-of-care handling due to their biocompatibility [631].
Hydrogen peroxide (H2O2) has received particular attention as an important analyte for human metabolism because any imbalance between its generation and consumption can damage lysosomal membranes and DNA or induce apoptosis [632,633,634]. Therefore, reliable, accurate, and rapid sensing techniques for cellular peroxide detection are of paramount importance [635]. For instance, biogenic Se nanorods (SeNRs), fabricated using citric acid and flavonoids from lemon juice, serve as an H2O2 spectrometric sensor with interfering ions and a visual color change technique from reddish to faint pink [636]. This study also demonstrates a morphology change by chemical surface leaching from nano-rod to nano-oval, proving the selectivity of Se nanomaterials to preferentially detect peroxide over other cellular cationic ions based on surface plasmon resonance. Moreover, spherical monoclinic SeNPs, synthesized using B. subtilis, exhibit high electrocatalytic activity in detecting H2O2 with a detection limit of 8 × 10−8 M, and show good adhesive ability and biocompatibility towards its substrates such as heme-containing proteins/enzymes [637]. Furthermore, colloidally stable SeNPs, produced using B. pumilus cell-free extract, enable the design of a low cost, sensitive, and reproducible H2O2 biosensor [638]. In addition, biogenic SeNPs, fabricated using the bacterial isolate P. aeruginosa, are used to design a biosensor for the visual assessment of the relative toxicity of a variety of chemicals, involving the inhibition of the bioreduction process of SeO32− in NP-treated bacterial culture supernatant, as a toxicity end-point [639]. This novel Se-bioassay could be easily applied to prescreen a plethora of environmental toxicants including nanostructures prior to intensive toxicity investigations. This adds up to the improved oxidase-like activity of CTS-SeNPs that provides a low-cost colorimetric method for Hg2+ detection with a detection limit of 0.12 μM, broadening, thus, the application of biogenic SeNPs in chemical sensor systems [640]. Lastly, Zhao et al. performed a green and controlled synthesis of SeNPs through a self-assembled method on molecular imprinting sites of zeolite-chitosan-TiO2 microspheres by coupling chitosan biosorption and TiO2 photocatalysis [641]. These NPs were successfully employed for dot-blot immunoassays with multiple native antigens for the prompt serodiagnosis of human lung cancer.

7.8.2. Cellular Imaging

SeNPs have become one of the most prospective and potential tools for cancer diagnosis and therapy. For example, Korany et al. fabricated SeNPs capped with glutathione as a novel radio-platform for tumor imaging by studying the radiochemical yield of radioactive technetium-99m (99mTc) in intravenous and intratumoral routes [346]. Sun et al. showed that luminescent Ru(II)-thiol SeNPs possess high tumor-targeted fluorescence imaging in HepG2 and HUVEC malignant cells while displaying improved antitumor efficacy and decreased systemic toxicity [642]. These functional SeNPs exhibit a well-defined, time-dependent increase in fluorescence intensity from 115 ± 17 a.u. after 0.5 h to 1171 ± 127 a.u. after 4 h. Huang et al. used SeNPs to build a smart drug-delivery nanoplatform to achieve simultaneous diagnosis, real-time monitoring, and therapy of cancer [643]. In this platform, epidermal growth factor receptor (EGFR) is used as the targeting molecule, gadolinium chelate as the magnetic resonance imaging contrast agent, polyamidoamine (PAMAM) and 3,3′-dithiobis (sulfosuccinimidyl propionate) as the response agents of intratumoral glutathione, 5-fluorouracil (5Fu) and cetuximab as drug payloads, and the pH as the release stimulus for the combined diagnosis and treatment of nasopharyngeal carcinoma (NPC) using SeNPs.
Vitamin C is a notable antioxidant human vitamin widely employed as a coating in NPs to prevent aggregation and achieve enhanced size and shape control and stabilization. Vitamin C-stabilized SeNPs (Vit C-SeNPs), labeled with 99mTc for further in vivo studies on normal and solid tumors induced in mice, exhibit an enhanced antioxidant activity leading to improved uptake and retention by tumor cells [34]. Moreover, functional SeNPs are used to develop a siRNA-delivery system for vascular endothelial growth factor (VEGF), a known signaling molecule involved in cancer [644]. This design comprises two nanostructures, SeNPs@siRNA and G2/PAH-Cit/SeNPs@siRNA; the latter is a pH-sensitive delivery system able to improve siRNA loading. Importantly, the utilization of pH-sensitive functional SeNPs results in no lesions in major target organs, thereby offering a novel, safe, and promising cancer treatment. Furthermore, photodynamic SeNPs, modified with photosensitive and macrophage-targeting bilayers, control activated macrophages and quench the intracellular H2O2 and NO that are associated with chronic inflammation diseases [560]. The first layer of the photosensitive system consists of, principally, a conjugate of a photosensitizer (rose bengal, RB) and a thiolated chitosan (chitosan-glutathione), while the second layer is made by conjugating hyaluronic acid with folic acid using an ethylenediamine linker. In addition, the intense and long-lasting intrinsic fluorescence of individual SeNPs in the visible to near infrared range enables their application for real-time tracking and imaging in cells, without the need of any chemical tags or dyes [645]. Further, biogenic SeNPs, produced using a Se-tolerant strain of S. maltophilia, exhibit a higher ability to emit light (photoluminescence) than organic fluorophores, thus constituting potential markers for bioimaging and fluorescence lifetime imaging microscopy (FLIM) [219].

8. Translational Nanomedicine: Recent Progress, Emerging Challenges, and Future Prospects for Biogenic Selenium Nanoparticles

Nanotechnology has triggered tremendous developments that benefit different fields of science, especially biomedicine [646,647,648,649,650]. To improve human health, scientific discoveries, which start at “the bench” as a result of fundamental research, must be translated into practical applications by progressing to the clinical stage, named “bedside” [651]. The use of nanotechnology in molecular medicine offers a plethora of advantages, such as local and ultrafast strategies at the nanometer length scale (i.e., diffusion, intermixing, and sensor response), controlled and intensified physical and chemical processes, direct access to biomarkers, and real time studies [652]. Successfully translating nanomedicine agents from pre-clinical proof-of-concept to demonstrated therapeutic value in the clinic remains challenging since it is of paramount importance to develop more precise and better translatable nanodevices towards a patient-focused and disease-specific targeting from the outset [357,653]. Hence, nanosystems must grow substantially in safety and sophistication before focused smart nanomedicine can become a reality in which a single platform performs seamless processes ranging from ultrasensitive diagnosis to pinpoint therapy [654]. To enhance nanomedicine translation and performance, targeted therapies should employ a specific decision-making framework: correct tissue/exposure, correct target/efficacy, correct safety, correct patient, and correct commercial potential [655]. Moreover, translation strategies usually require innovation in the laboratory that must be supported by the pillars of evidence-based medicine for predictable regulatory outcomes.
Extensive research on engineered nanomaterials has led to the design of numerous nano-based formulations, i.e., nanomedicines, for theranostic applications with one of the significant benefits lying in the ability to formulate a drug without using dose-limiting toxic excipients present in many current marketed formulations, often enhancing tolerability and containing more active molecules to be administered to patients [655,656]. To date, SeNPs have demonstrated great preclinical applications in diagnosis and gene and drug delivery in cancer therapy owing, particularly, to their selective and effective accumulation in tumors through the enhanced permeability and retention (EPR) effect [119,657]. Nevertheless, there is a lack of research about the benefits of SeNPs in clinical settings, although they constitute a good candidate for advanced-stage clinical research due to their low toxicity and excellent biocompatibility [61,67].
The development of safer SeNPs with enhanced therapeutic efficacy in clinical settings requires a better understanding of the toxicity, possible side effects, and interaction with the biological environment. Before envisaging their application in humans, it is important to test their biosafety, degradation rate, long-term metabolic activity, pharmacokinetics and pharmacodynamics, and interaction with cells, organoids, etc. [658]. Nanomaterials usually interact with biomolecules in the physiological milieu, such as plasma proteins, which results in corona formation [656]. This corona can alter the NM stability, targeting ability, bio-identity, cellular uptake, dissolution properties, and change their biodistribution and in vivo toxicity [24,342,393,659,660]. Therefore, the nano–bio interface of SeNPs needs further assessments in terms of biomedical safety following rigorous methodologies, such as ADME-Tox as pointed out earlier.
The design and delivery of therapeutics to the brain has been an ongoing challenge in the treatment of brain tumors, especially due to the blood–brain barrier (BBB) that impedes reaching a proper local drug concentration [661,662,663,664]. Nanotechnology advances, such as the adequate surface functionalization, have improved the NP penetration across the BBB by receptor-mediated transcytosis [391,665]. For example, SeNPs coated with B6 peptide and functionalized with sialic acid (B6-SA-SeNPs) inhibit Aβ aggregation and pass the BBB, becoming a potential therapeutic nanovehicle to treat Alzheimer’s disease [666].
Although biogenic NPs are inexpensive, clean, non-toxic, and facile to produce via scalable processes, there has not yet been much work on their application in the fields of diagnostics and therapeutics, notably at the clinical stage [667,668,669,670,671]. The promising clinical applications of plant-synthesized metallic NPs have fueled a swelling interest switching from routine antioxidant and antimicrobial studies on trivial microbial lines to antibiotic-resistant pathogens and antitumor studies [672]. Furthermore, biogenic colloidal metallic NPs, especially those made of silver and gold, are found to be multifunctional theranostic agents [673]. However, more research should be directed towards developing facile and greener techniques for the large-scale fabrication of SeNPs using biological resources, as they hold potential applications in different aspects of nanomedicine, especially in combating cancer (Table 4). Indeed, selenium has been drawing great interest in the following clinical aspects: radioprotection of normal tissues, radiosensitizing in malignant tumors, antiedematous activity, prognostic impact, and effects on primary and secondary cancer prevention [674].
In a comprehensive review on strategies to enhance the success of clinical translation of cancer nanomedicines, van der Meel et al. listed four important factors: (i) stratification and selection of patients likely to respond to nanomedicine-based therapy, (ii) rational drug choice rather than opportunistic preferences, (iii) combination and multimodal therapies for synergistic effects, and (iv) empowering immunotherapy [653]. Hence, the engineering strategy of NPs, including those made of selenium, needs to reach important goals to boost the desirable effects and achieve transformation from formulation-driven research to disease-driven rational development. These goals include highly stable association of drug and carrier in systemic circulation, enhanced drug delivery to cancer cells, and controlled and prolonged release of active drugs in affected tissues [675]. Furthermore, the benefits of nanostructure-based diagnostics lie in their potentially higher sensitivity and selectivity compared to classical methods, thus enabling the earlier diagnosis of diseases resulting in enhanced resolution and sensitivity that will, ultimately, lead to novel, fast, convenient, and inexpensive screening, diagnostic, and therapeutic tools in medicine [652,676,677].
Another challenge to overcome for the “next-generation nanomedicines” resides in the regulatory classification, because the lack of knowledge regarding biosafety and long-term effects of nanomaterials leads to a regulatory uncertainty and deficient standards and protocols for scale-up manufacturing and safe clinical uses [678,679]. This twilight zone in the nanomedicine market affects the effective collaborative work between stakeholders from industrial/academic R&D, professionals in the health system, regulatory bodies, and society (Figure 12). Moreover, classical methods of drug development and (non-)clinical assessments are expensive and encompass some uncertainties, delaying the advent of innovative approaches and the elucidation of safety concerns. For instance, the wrong information or the lack of full characterization of current nanoparticle-drug products leads to failures in follow-on versions, making further evaluation and manufacturing processes tedious before even envisaging any regulatory approval and marketing [679,680].
Advances in nanoscience and nanobiotechnology, combined with the call for personalized medicine based on nanotheranostics, have given rise to a cutting-edge, exciting, and fast-growing research area, where nanoscale Se formulations, including biogenic ones, are occupying a significant niche. It is necessary to precisely utilize engineered, nanotechnology-enabled solutions to face the challenges of the current drug delivery, imaging, therapeutics, and diagnostics procedures. Therefore, nanomedicine can lead to the next generation of biomedical breakthroughs by accelerating the translation of therapies with greater efficacy and reduced side effects into personalized/precision nanomedicine since innovative and clinically effective nanotherapeutics hold the potential to revolutionize nanoscale healthcare and pharmaceutical products and applications [681].

9. Conclusions and Perspectives

Selenium is a significant trace element which, in its elemental form or in the form of its various chemical species and Se-containing biomolecules, offers unique biological properties for the proper function of the body. Due to its high bioavailability, low toxicity, and affordability, nano-sized selenium has proven to be a proper nutritional supplement and an efficient theranostic agent. Based on these properties and the great versatility in their control thanks to green, biogenic routes for their synthesis and modification, SeNPs are emerging as a research hotspot in nanomedicine with the potential for promising benefits in clinical settings. Since biogenic SeNPs are free from toxic/hazardous components, they are well suited in biomedicine and therapeutics. Current findings highlight the outstanding physicochemical and biological properties of biogenic SeNPs that pave the way for extensive biomedical contributions, including pharmaceutical, therapeutic, imaging, and diagnostic applications. As such, SeNPs have been shown to combat cancer, pathogenic infections, diabetes, inflammatory syndromes, cardiovascular and neurological diseases, and drug-induced cyto- and geno-toxicity, among others. In addition to their antimicrobial and antiparasitic applications, SeNPs possess a potential utility in curbing viral outbreaks including the ongoing Covid-19 pandemic. Importantly, SeNPs constitute valuable nanoplatforms with multiple desired features for clinical translation. Thanks to the possibility of precise calibration and rational modification of their physicochemical properties, novel SeNPs are particularly attractive as therapeutic agents easily transportable in the organism and offering stability in the physiological microenvironment of target tissues. However, researchers are still in need of investigation into possible side effects due to the relatively narrow therapeutic window of Se compounds. Se nanoparticle-based diagnosis and therapy are in their early stages and preparing to progress into clinical trials. At this stage, many of these compounds could offer new mechanistic insights and pave the way towards the rational design of novel therapeutics following extensive analyses of structure–activity relationships (SARs). Thus, it is still necessary to accomplish further preclinical safety and selectivity studies before these new eco-friendly SeNPs are safely translated into clinical practice.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Silva, G.A. Introduction to nanotechnology and its applications to medicine. Surg. Neurol. 2004, 61, 216–220. [Google Scholar] [CrossRef] [PubMed]
  2. Jamkhande, P.G.; Ghule, N.W.; Bamer, A.H.; Kalaskar, M.G. Metal nanoparticles synthesis: An overview on methods of preparation, advantages and disadvantages, and applications. J. Drug Deliv. Sci. Technol. 2019, 53, 101174. [Google Scholar] [CrossRef]
  3. Morigi, V.; Tocchio, A.; Pellegrini, C.B.; Sakamoto, J.H.; Arnone, M.; Tasciotti, E. Nanotechnology in medicine: From inception to market domination. J. Drug Deliv. 2012, 2012, 389485. [Google Scholar] [CrossRef] [PubMed]
  4. Ventola, C.L. Progress in nanomedicine: Approved and investigational nanodrugs. Pharm. Ther. 2017, 42, 742–755. [Google Scholar]
  5. Andraos, C.; Gulumian, M. Intracellular and extracellular targets as mechanisms of cancer therapy by nanomaterials in relation to their physicochemical properties. WIREs Nanomed. Nanobiotechnol. 2021, 13, e1680. [Google Scholar] [CrossRef]
  6. Wei, G.; Wang, Y.; Yang, G.; Wang, Y.; Ju, R. Recent progress in nanomedicine for enhanced cancer chemotherapy. Theranostics 2021, 11, 6370–6392. [Google Scholar] [CrossRef]
  7. Zhang, C.; Yan, L.; Wang, X.; Zhu, S.; Chen, C.; Gu, Z.; Zhao, Y. Progress, challenges, and future of nanomedicine. Nano Today 2020, 35, 101008. [Google Scholar] [CrossRef]
  8. Mura, S.; Couvreur, P. Nanotheranostics for personalized medicine. Adv. Drug Deliv. Rev. 2012, 64, 1394–1416. [Google Scholar] [CrossRef]
  9. Zhang, X.-Q.; Xu, X.; Bertrand, N.; Pridgen, E.; Swami, A.; Farokhzad, O.C. Interactions of nanomaterials and biological systems: Implications to personalized nanomedicine. Adv. Drug Deliv. Rev. 2012, 64, 1363–1384. [Google Scholar] [CrossRef]
  10. Kim, T.H.; Lee, S.; Chen, X. Nanotheranostics for personalized medicine. Expert Rev. Mol. Diagn. 2013, 13, 257–269. [Google Scholar] [CrossRef]
  11. Caster, J.; Patel, A.N.; Zhang, T.; Wang, A. Investigational nanomedicines in 2016: A review of nanotherapeutics currently undergoing clinical trials. WIREs Nanomed. Nanobiotechnol. 2017, 9, e1416. [Google Scholar] [CrossRef]
  12. Xie, J.; Lee, S.; Chen, X. Nanoparticle-based theranostic agents. Adv. Drug Deliv. Rev. 2010, 62, 1064–1079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Farjadian, F.; Ghasemi, A.; Gohari, O.; Roointan, A.; Karimi, M.; Hamblin, M.R. Nanopharmaceuticals and nanomedicines currently on the market: Challenges and opportunities. Nanomedicine 2019, 14, 93–126. [Google Scholar] [CrossRef] [PubMed]
  14. Siddique, S.; Chow, J.C.L. Recent Advances in Functionalized Nanoparticles in Cancer Theranostics. Nanomaterials 2022, 12, 2826. [Google Scholar] [CrossRef]
  15. Zhou, H.; Ge, J.; Miao, Q.; Zhu, R.; Wen, L.; Zeng, J.; Gao, M. Biodegradable inorganic nanoparticles for cancer theranostics: Insights into the degradation behavior. Bioconjug. Chem. 2020, 31, 315–331. [Google Scholar] [CrossRef]
  16. Shete, M.B.; Patil, T.S.; Deshpande, A.S.; Saraogi, G.; Vasdev, N.; Deshpande, M.; Rajpoot, K.; Tekade, R.K. Current trends in theranostic nanomedicines. J. Drug Deliv. Sci. Technol. 2022, 71, 103280. [Google Scholar] [CrossRef]
  17. Hu, X.; Ha, E.; Ai, F.; Huang, X.; Yan, L.; He, S.; Ruan, S.; Hu, J. Stimulus-responsive inorganic semiconductor nanomaterials for tumor-specific theranostics. Co-ord. Chem. Rev. 2022, 473, 214821. [Google Scholar] [CrossRef]
  18. Satalkar, P.; Elger, B.S.; Shaw, D.M. Defining nano, nanotechnology and nanomedicine: Why should it matter? Sci. Eng. Ethics 2016, 22, 1255–1276. [Google Scholar] [CrossRef]
  19. Roco, M.C.; Harthorn, B.; Guston, D.; Shapira, P. Innovative and responsible governance of nanotechnology for societal development. J. Nanopart. Res. 2011, 13, 3557–3590. [Google Scholar] [CrossRef]
  20. McNamara, K.; Tofail, S.A.M. Nanoparticles in biomedical applications. Adv. Phys. X 2016, 2, 54–88. [Google Scholar] [CrossRef]
  21. Khan, I.; Saeed, K.; Khan, I. Nanoparticles: Properties, applications and toxicities. Arab. J. Chem. 2019, 12, 908–931. [Google Scholar] [CrossRef]
  22. Pereira, L.; Mehboob, F.; Stams, A.J.M.; Mota, M.M.; Rijnaarts, H.; Alves, M.M. Metallic nanoparticles: Microbial synthesis and unique properties for biotechnological applications, bioavailability and biotransformation. Crit. Rev. Biotechnol. 2015, 35, 114–128. [Google Scholar] [CrossRef] [PubMed]
  23. Doane, T.L.; Burda, C. The unique role of nanoparticles in nanomedicine: Imaging, drug delivery and therapy. Chem. Soc. Rev. 2012, 41, 2885–2911. [Google Scholar] [CrossRef] [PubMed]
  24. Kang, H.; Mintri, S.; Menon, A.V.; Lee, H.Y.; Choi, H.S.; Kim, J. Pharmacokinetics, pharmacodynamics and toxicology of theranostic nanoparticles. Nanoscale 2015, 7, 18848–18862. [Google Scholar] [CrossRef] [Green Version]
  25. Laroui, H.; Rakhya, P.; Xiao, B.; Viennois, E.; Merlin, D. Nanotechnology in diagnostics and therapeutics for gastrointestinal disorders. Dig. Liver Dis. 2013, 45, 995–1002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Dykman, L.; Khlebtsov, N. Gold nanoparticles in biomedical applications: Recent advances and perspectives. Chem. Soc. Rev. 2012, 41, 2256–2282. [Google Scholar] [CrossRef] [PubMed]
  27. Moore, J.A.; Chow, J.C.L. Recent progress and applications of gold nanotechnology in medical biophysics using artificial intelligence and mathematical modeling. Nano Express 2021, 2, 022001. [Google Scholar] [CrossRef]
  28. AlShehri, S.; Imam, S.S.; Rizwanullah; Akhter, S.; Mahdi, W.; Kazi, M.; Ahmad, J. Progress of Cancer Nanotechnology as Diagnostics, Therapeutics, and Theranostics Nanomedicine: Preclinical Promise and Translational Challenges. Pharmaceutics 2020, 13, 24. [Google Scholar] [CrossRef]
  29. Ahmad, M.Z.; Rizwanullah; Ahmad, J.; Alasmary, M.Y.; Akhter, H.; Abdel-Wahab, B.A.; Warsi, M.H.; Haque, A. Progress in nanomedicine-based drug delivery in designing of chitosan nanoparticles for cancer therapy. Int. J. Polym. Mater. Polym. Biomater. 2021, 71, 602–623. [Google Scholar] [CrossRef]
  30. Song, Y.; Huang, Y.; Zhou, F.; Ding, J.; Zhou, W. Macrophage-targeted nanomedicine for chronic diseases immunotherapy. Chin. Chem. Lett. 2022, 33, 597–612. [Google Scholar] [CrossRef]
  31. Emerich, D.F. Nanomedicine—Prospective therapeutic and diagnostic applications. Expert Opin. Biol. Ther. 2005, 5, 1–5. [Google Scholar] [CrossRef] [PubMed]
  32. Liu, H.; Mei, C.; Deng, X.; Lin, W.; He, L.; Chen, T. Rapid visualizing and pathological grading of bladder tumor tissues by simple nanodiagnostics. Biomaterials 2021, 264, 120434. [Google Scholar] [CrossRef] [PubMed]
  33. Yuan, P.; Ding, X.; Yang, Y.Y.; Xu, Q.-H. Metal Nanoparticles for Diagnosis and Therapy of Bacterial Infection. Adv. Health Mater. 2018, 7, e1701392. [Google Scholar] [CrossRef] [PubMed]
  34. Korany, M.; Mahmoud, B.; Ayoub, S.M.; Sakr, T.M.; Ahmed, S.A. Synthesis and radiolabeling of vitamin C-stabilized selenium nanoparticles as a promising approach in diagnosis of solid tumors. J. Radioanal. Nucl. Chem. 2020, 325, 237–244. [Google Scholar] [CrossRef]
  35. Liu, T.; Xu, L.; He, L.; Zhao, J.; Zhang, Z.; Chen, Q.; Chen, T. Selenium nanoparticles regulates selenoprotein to boost cytokine-induced killer cells-based cancer immunotherapy. Nano Today 2020, 35, 100975. [Google Scholar] [CrossRef]
  36. Mukherjee, A.; Paul, M.; Mukherjee, S. Recent Progress in the Theranostics Application of Nanomedicine in Lung Cancer. Cancers 2019, 11, 597. [Google Scholar] [CrossRef] [Green Version]
  37. Augustine, R.; Hasan, A. Emerging applications of biocompatible phytosynthesized metal/metal oxide nanoparticles in healthcare. J. Drug Deliv. Sci. Technol. 2020, 56, 101516. [Google Scholar] [CrossRef]
  38. Dimkpa, C.O.; McLean, J.E.; Britt, D.W.; Anderson, A.J. Nano-CuO and interaction with nano-ZnO or soil bacterium provide evidence for the interference of nanoparticles in metal nutrition of plants. Ecotoxicology 2015, 24, 119–129. [Google Scholar] [CrossRef]
  39. Kah, M.; Tufenkji, N.; White, J.C. Nano-enabled strategies to enhance crop nutrition and protection. Nat. Nanotechnol. 2019, 14, 532–540. [Google Scholar] [CrossRef]
  40. Baetke, S.C.; Lammers, T.; Kiessling, F. Applications of nanoparticles for diagnosis and therapy of cancer. Br. J. Radiol. 2015, 88, 20150207. [Google Scholar] [CrossRef]
  41. Kieliszek, M.; Błażejak, S. Current knowledge on the importance of selenium in food for living organisms: A Review. Molecules 2016, 21, 609. [Google Scholar] [CrossRef] [PubMed]
  42. 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] [PubMed]
  43. Mehdi, Y.; Hornick, J.-L.; Istasse, L.; Dufrasne, I. Selenium in the environment, metabolism and involvement in body functions. Molecules 2013, 18, 3292–3311. [Google Scholar] [CrossRef] [Green Version]
  44. Li, D.-B.; Cheng, Y.-Y.; Wu, C.; Li, W.-W.; Li, N.; Yang, Z.-C.; Tong, Z.-H.; Yu, H.-Q. Selenite reduction by Shewanella oneidensis MR-1 is mediated by fumarate reductase in periplasm. Sci. Rep. 2014, 4, 3735. [Google Scholar] [CrossRef] [Green Version]
  45. Biswas, K.C.; Barton, L.L.; Tsui, W.L.; Shuman, K.; Gillespie, J.; Eze, C.S. A novel method for the measurement of elemental selenium produced by bacterial reduction of selenite. J. Microbiol. Methods 2011, 86, 140–144. [Google Scholar] [CrossRef] [PubMed]
  46. Bhattacharjee, A.; Basu, A.; Bhattacharya, S. Selenium nanoparticles are less toxic than inorganic and organic selenium to mice in vivo. Nucleus 2019, 62, 259–268. [Google Scholar] [CrossRef]
  47. Zhang, J.S.; Wang, X.F.; Xu, T.W. Elemental selenium at nano size (nano-Se) as a potential chemopreventive agent with reduced risk of selenium toxicity: Comparison with Se-methylselenocysteine in mice. Toxicol. Sci. 2008, 101, 22–31. [Google Scholar] [CrossRef] [Green Version]
  48. Tang, S.; Wang, T.; Jiang, M.; Huang, C.; Lai, C.; Fan, Y.; Yong, Q. Construction of arabinogalactans/selenium nanoparticles composites for enhancement of the antitumor activity. Int. J. Biol. Macromol. 2019, 128, 444–451. [Google Scholar] [CrossRef]
  49. Wu, H.; Li, X.; Liu, W.; Chen, T.; Li, Y.; Zheng, W.; Man, C.W.-Y.; Wong, M.-K.; Wong, K.-H. Surface decoration of selenium nanoparticles by mushroom polysaccharides–protein complexes to achieve enhanced cellular uptake and antiproliferative activity. J. Mater. Chem. 2012, 22, 9602–9610. [Google Scholar] [CrossRef]
  50. Menon, S.; Ks, S.D.; Santhiya, R.; Rajeshkumar, S.; Kumar, V. Selenium nanoparticles: A potent chemotherapeutic agent and an elucidation of its mechanism. Colloids Surf. B Biointerfaces 2018, 170, 280–292. [Google Scholar] [CrossRef]
  51. Zhai, X.; Zhang, C.; Zhao, G.; Stoll, S.; Ren, F.; Leng, X. Antioxidant capacities of the selenium nanoparticles stabilized by chitosan. J. Nanobiotechnol. 2017, 15, 4. [Google Scholar] [CrossRef] [PubMed]
  52. Moneim, A.E.A.; Al-Quraishy, S.; Dkhil, M.A. Anti-hyperglycemic activity of selenium nanoparticles in streptozotocin-induced diabetic rats. Int. J. Nanomed. 2015, 10, 6741–6756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Abdulmalek, S.A.; Balbaa, M. Synergistic effect of nano-selenium and metformin on type 2 diabetic rat model: Diabetic complications alleviation through insulin sensitivity, oxidative mediators and inflammatory markers. PLoS ONE 2019, 14, e0220779. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Webster, T.J.; Tran, P. Selenium nanoparticles inhibit Staphylococcus aureus growth. Int. J. Nanomed. 2011, 6, 1553–1558. [Google Scholar] [CrossRef] [Green Version]
  55. Srivastava, N.; Mukhopadhyay, M. Green synthesis and structural characterization of selenium nanoparticles and assessment of their antimicrobial property. Bioprocess Biosyst. Eng. 2015, 38, 1723–1730. [Google Scholar] [CrossRef]
  56. Khiralla, G.M.; El-Deeb, B.A. Antimicrobial and antibiofilm effects of selenium nanoparticles on some foodborne pathogens. LWT—Food Sci. Technol. 2015, 63, 1001–1007. [Google Scholar] [CrossRef]
  57. Chaudhary, S.; Mehta, S.K. Selenium nanomaterials: Applications in electronics, catalysis and sensors. J. Nanosci. Nanotechnol. 2014, 14, 1658–1674. [Google Scholar] [CrossRef]
  58. Hosnedlova, B.; Kepinska, M.; Skalickova, S.; Fernandez, C.; Ruttkay-Nedecky, B.; Peng, Q.; Baron, M.; Melcova, M.; Opatrilova, R.; Zidkova, J.; et al. Nano-selenium and its nanomedicine applications: A critical review. Int. J. Nanomed. 2018, 13, 2107–2128. [Google Scholar] [CrossRef] [Green Version]
  59. Bisht, N.; Phalswal, P.; Khanna, P.K. Selenium nanoparticles: A review on synthesis and biomedical applications. Mater. Adv. 2022, 3, 1415–1431. [Google Scholar] [CrossRef]
  60. Zambonino, M.; Quizhpe, E.; Jaramillo, F.; Rahman, A.; Vispo, N.S.; Jeffryes, C.; Dahoumane, S. Green synthesis of selenium and tellurium nanoparticles: Current trends, biological properties and biomedical applications. Int. J. Mol. Sci. 2021, 22, 989. [Google Scholar] [CrossRef]
  61. Ikram, M.; Javed, B.; Raja, N.I.; Mashwani, Z.-U. Biomedical potential of plant-based selenium nanoparticles: A comprehensive review on therapeutic and mechanistic aspects. Int. J. Nanomed. 2021, 16, 249–268. [Google Scholar] [CrossRef] [PubMed]
  62. Cremonini, E.; Zonaro, E.; Donini, M.; Lampis, S.; Boaretti, M.; Dusi, S.; Melotti, P.; Lleo, M.M.; Vallini, G. Biogenic selenium nanoparticles: Characterization, antimicrobial activity and effects on human dendritic cells and fibroblasts. Microb. Biotechnol. 2016, 9, 758–771. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Yazdi, M.H.; Mahdavi, M.; Varastehmoradi, B.; Faramarzi, M.A.; Shahverdi, A.R. The immunostimulatory effect of biogenic selenium nanoparticles on the 4T1 breast cancer model: An in vivo study. Biol. Trace Elem. Res. 2012, 149, 22–28. [Google Scholar] [CrossRef] [PubMed]
  64. Chandra, H.; Kumari, P.; Bontempi, E.; Yadav, S. Medicinal plants: Treasure trove for green synthesis of metallic nanoparticles and their biomedical applications. Biocatal. Agric. Biotechnol. 2020, 24, 101518. [Google Scholar] [CrossRef]
  65. Yanchatuña Aguayo, O.P.; Mouheb, L.; Villota Revelo, K.; Vásquez-Ucho, P.A.; Pawar, P.P.; Rahman, A.; Jeffryes, C.; Terencio, T.; Dahoumane, S.A. Biogenic sulfur-based chalcogenide nanocrystals: Methods of fabrication, mechanistic aspects, and bio-applications. Molecules 2022, 27, 458. [Google Scholar] [PubMed]
  66. Verma, A.; Gautam, S.P.; Bansal, K.K.; Prabhakar, N.; Rosenholm, J.M. Green nanotechnology: Advancement in phytoformulation research. Medicines 2019, 6, 39. [Google Scholar] [CrossRef] [Green Version]
  67. Yan, J.-K.; Qiu, W.-Y.; Wang, Y.-Y.; Wang, W.-H.; Yang, Y.; Zhang, H.-N. Fabrication and stabilization of biocompatible selenium nanoparticles by carboxylic curdlans with various molecular properties. Carbohydr. Polym. 2018, 179, 19–27. [Google Scholar] [CrossRef]
  68. San Keskin, N.O.; Akbal Vural, O.; Abaci, S. Biosynthesis of noble selenium nanoparticles from Lysinibacillus sp. NOSK for antimicrobial, antibiofilm activity, and biocompatibility. Geomicrobiol. J. 2020, 37, 919–928. [Google Scholar] [CrossRef]
  69. Nancharaiah, Y.V.; Lens, P.N. Selenium biomineralization for biotechnological applications. Trends Biotechnol. 2015, 33, 323–330. [Google Scholar] [CrossRef]
  70. Salama, L.; Pastor, E.R.; Stone, T.; Mousa, S.A. Emerging nanopharmaceuticals and nanonutraceuticals in cancer management. Biomedicines 2020, 8, 347. [Google Scholar] [CrossRef]
  71. Janjic, J.M.; Gorantla, V.S. Novel nanoimaging strategies for noninvasive graft monitoring in vascularized composite allotransplantation. Curr. Transplant. Rep. 2018, 5, 369–372. [Google Scholar] [CrossRef]
  72. Patra, J.K.; Das, G.; Fraceto, L.F.; Campos, E.V.R.; del Pilar Rodriguez-Torres, M.; Acosta-Torres, L.S.; Diaz-Torres, L.A.; Grillo, R.; Swamy, M.K.; Sharma, S.; et al. Nano based drug delivery systems: Recent developments and future prospects. J. Nanobiotechnol. 2018, 16, 71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Khurana, A.; Tekula, S.; Saifi, M.A.; Venkatesh, P.; Godugu, C. Therapeutic applications of selenium nanoparticles. Biomed. Pharmacother. 2019, 111, 802–812. [Google Scholar] [CrossRef] [PubMed]
  74. Guan, B.; Yan, R.; Li, R.; Zhang, X. Selenium as a pleiotropic agent for medical discovery and drug delivery. Int. J. Nanomed. 2018, 13, 7473–7490. [Google Scholar] [CrossRef] [Green Version]
  75. Steinbrenner, H.; Speckmann, B.; Klotz, L.-O. Selenoproteins: Antioxidant selenoenzymes and beyond. Arch. Biochem. Biophys. 2016, 595, 113–119. [Google Scholar] [CrossRef] [PubMed]
  76. Skalickova, S.; Milosavljevic, V.; Cihalova, K.; Horky, P.; Richtera, L.; Adam, V. Selenium nanoparticles as a nutritional supplement. Nutrition 2017, 33, 83–90. [Google Scholar] [CrossRef]
  77. Kieliszek, M.; Błażejak, S. Selenium: Significance, and outlook for supplementation. Nutrition 2013, 29, 713–718. [Google Scholar] [CrossRef]
  78. Fernandes, A.P.; Gandin, V. Selenium compounds as therapeutic agents in cancer. Biochim. Biophys. Acta 2015, 1850, 1642–1660. [Google Scholar] [CrossRef]
  79. Sanmartin, C.; Plano, D.; Sharma, A.K.; Palop, J.A. Selenium compounds, apoptosis and other types of cell death: An overview for cancer therapy. Int. J. Mol. Sci. 2012, 13, 9649–9672. [Google Scholar] [CrossRef]
  80. Duntas, L.H.; Benvenga, S. Selenium: An element for life. Endocrine 2015, 48, 756–775. [Google Scholar] [CrossRef]
  81. Sun, H.-J.; Rathinasabapathi, B.; Wu, B.; Luo, J.; Pu, L.-P.; Ma, L.Q. Arsenic and selenium toxicity and their interactive effects in humans. Environ. Int. 2014, 69, 148–158. [Google Scholar] [CrossRef] [PubMed]
  82. Kuršvietienė, L.; Mongirdienė, A.; Bernatonienė, J.; Šulinskienė, J.; Stanevičienė, I. Selenium anticancer properties and impact on cellular redox status. Antioxidants 2020, 9, 80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Gandin, V.; Khalkar, P.; Braude, J.; Fernandes, A.P. Organic selenium compounds as potential chemotherapeutic agents for improved cancer treatment. Free Radic. Biol. Med. 2018, 127, 80–97. [Google Scholar] [CrossRef] [PubMed]
  84. Kipp, A.P. Selenium in colorectal and differentiated thyroid cancer. Hormones 2020, 19, 41–46. [Google Scholar] [CrossRef] [PubMed]
  85. Tan, H.-W.; Mo, H.-J.; Lau, A.T.Y.; Xu, Y.-M. Selenium species: Current status and potentials in cancer prevention and therapy. Int. J. Mol. Sci. 2019, 20, 75. [Google Scholar] [CrossRef] [Green Version]
  86. Hoefig, C.S.; Renko, K.; Köhrle, J.; Birringer, M.; Schomburg, L. Comparison of different selenocompounds with respect to nutritional value vs. toxicity using liver cells in culture. J. Nutr. Biochem. 2011, 22, 945–955. [Google Scholar] [CrossRef]
  87. Cao, S.; A Durrani, F.; Tóth, K.; Rustum, Y.M. Se-methylselenocysteine offers selective protection against toxicity and potentiates the antitumour activity of anticancer drugs in preclinical animal models. Br. J. Cancer 2014, 110, 1733–1743. [Google Scholar] [CrossRef] [Green Version]
  88. Fan, C.; Chen, J.; Wang, Y.; Wong, Y.-S.; Zhang, Y.; Zheng, W.; Cao, W.; Chen, T. Selenocystine potentiates cancer cell apoptosis induced by 5-fluorouracil by triggering reactive oxygen species-mediated DNA damage and inactivation of the ERK pathway. Free Radic. Biol. Med. 2013, 65, 305–316. [Google Scholar] [CrossRef]
  89. Misra, S.; Boylan, M.; Selvam, A.; Spallholz, J.E.; Björnstedt, M. Redox-active aelenium compounds—From toxicity and cell death to cancer treatment. Nutrients 2015, 7, 3536–3556. [Google Scholar] [CrossRef] [Green Version]
  90. Zhang, J.; Wang, D.; Taylor, E.W.; Wang, Y.; Wan, X. Encapsulated nanoepigallocatechin-3-gallate and elemental selenium nanoparticles as paradigms for nanochemoprevention. Int. J. Nanomed. 2012, 7, 1711–1721. [Google Scholar] [CrossRef] [Green Version]
  91. Kumar, A.; Prasad, K.S. Role of nano-selenium in health and environment. J. Biotechnol. 2021, 325, 152–163. [Google Scholar] [CrossRef] [PubMed]
  92. Zhang, C.; Zhai, X.; Zhao, G.; Ren, F.; Leng, X. Synthesis, characterization, and controlled release of selenium nanoparticles stabilized by chitosan of different molecular weights. Carbohydr. Polym. 2015, 134, 158–166. [Google Scholar] [CrossRef] [PubMed]
  93. Muthu, S.; Raju, V.; Gopal, V.B.; Gunasekaran, A.; Narayan, K.S.; Malairaj, S.; Lakshmikanthan, M.; Duraisamy, N.; Krishnan, K.; Perumal, P. A rapid synthesis and antibacterial property of selenium nanoparticles using egg white lysozyme as a stabilizing agent. SN Appl. Sci. 2019, 1, 1543. [Google Scholar] [CrossRef] [Green Version]
  94. Zhang, J.; Teng, Z.; Yuan, Y.; Zeng, Q.-Z.; Lou, Z.; Lee, S.-H.; Wang, Q. Development, physicochemical characterization and cytotoxicity of selenium nanoparticles stabilized by beta-lactoglobulin. Int. J. Biol. Macromol. 2018, 107, 1406–1413. [Google Scholar] [CrossRef] [PubMed]
  95. Zhang, S.-Y.; Zhang, J.; Wang, H.-Y.; Chen, H.-Y. Synthesis of selenium nanoparticles in the presence of polysaccharides. Mater. Lett. 2004, 58, 2590–2594. [Google Scholar] [CrossRef]
  96. Gao, S.; Li, T.; Guo, Y.; Sun, C.; Xianyu, B.; Xu, H. Selenium-containing nanoparticles combine the NK Cells mediated immunotherapy with radiotherapy and chemotherapy. Adv. Mater. 2020, 32, e1907568. [Google Scholar] [CrossRef] [PubMed]
  97. Hu, Y.; Liu, T.; Li, J.; Mai, F.; Li, J.; Chen, Y.; Jing, Y.; Dong, X.; Lin, L.; He, J.; et al. Selenium nanoparticles as new strategy to potentiate gd T cell anti-tumor cytotoxicity through upregulation of tubulin-a acetylation. Biomaterials 2019, 222, 119397. [Google Scholar] [CrossRef]
  98. Neamat-Allah, A.N.; Mahmoud, E.; El Hakim, Y.A. Efficacy of dietary nano-selenium on growth, immune response, antioxidant, transcriptomic profile and resistance of Nile tilapia, Oreochromis niloticus against Streptococcus iniae infection. Fish Shellfish Immunol. 2019, 94, 280–287. [Google Scholar] [CrossRef]
  99. Dawood, M.A.O.; Zommara, M.; Eweedah, N.M.; Helal, A.I.; Aboel-Darag, M.A. The potential role of nano-selenium and vitamin C on the performances of Nile tilapia (Oreochromis niloticus). Environ. Sci. Pollut. Res. 2020, 27, 9843–9852. [Google Scholar] [CrossRef]
  100. Zhu, M.L.; Wang, G.; Wang, H.; Guo, Y.M.; Song, P.; Xu, J.; Li, P.; Wang, S.; Yang, L. Amorphous nano-selenium quantum dots improve endothelial dysfunction in rats and prevent atherosclerosis in mice through Na+/H+ exchanger 1 inhibition. Vascul. Pharmacol. 2019, 115, 26–32. [Google Scholar] [CrossRef]
  101. Shi, L.-G.; Yang, R.-J.; Yue, W.-B.; Xun, W.-J.; Zhang, C.-X.; Ren, Y.-S.; Shi, L.; Lei, F.-L. Effect of elemental nano-selenium on semen quality, glutathione peroxidase activity, and testis ultrastructure in male Boer goats. Anim. Reprod. Sci. 2010, 118, 248–254. [Google Scholar] [CrossRef] [PubMed]
  102. Prasad, K.S.; Selvaraj, K. Biogenic Synthesis of Selenium Nanoparticles and Their Effect on As(III)-Induced Toxicity on Human Lymphocytes. Biol. Trace Elem. Res. 2014, 157, 275–283. [Google Scholar] [CrossRef] [PubMed]
  103. Hashem, K.; Hassnin, K.M.A.; AbdEl-Kawi, S.H. The prospective protective effect of selenium nanoparticles against chromium-induced oxidative and cellular damage in rat thyroid. Int. J. Nanomed. 2013, 8, 1713–1720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Trabelsi, H.; Azzouz, I.; Ferchichi, S.; Tebourbi, O.; Sakly, M.; Abdelmelek, H. Nanotoxicological evaluation of oxidative responses in rat nephrocytes induced by cadmium. Int. J. Nanomed. 2013, 8, 3447–3453. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Sheiha, A.M.; Abdelnour, S.A.; El-Hack, M.E.A.; Khafaga, A.F.; Metwally, K.A.; Ajarem, J.S.; Maodaa, S.N.; Allam, A.A.; El-Saadony, M.T. Effects of dietary biological or chemical-synthesized nano-selenium supplementation on growing rabbits exposed to thermal stress. Animals 2020, 10, 430. [Google Scholar] [CrossRef] [Green Version]
  106. Tran, P.A.; Sarin, L.; Hurt, R.H.; Webster, T.J. Differential effects of nanoselenium doping on healthy and cancerous osteoblasts in coculture on titanium. Int. J. Nanomed. 2010, 5, 351–358. [Google Scholar]
  107. Bartůněk, V.; Vokatá, B.; Kolářová, K.; Ulbrich, P. Preparation of amorphous nano-selenium-PEG composite network with selective antimicrobial activity. Mater. Lett. 2019, 238, 51–53. [Google Scholar] [CrossRef]
  108. El-Sayyad, G.S.; El-Bastawisy, H.S.; Gobara, M.; El-Batal, A.I. Gentamicin-assisted mycogenic selenium nanoparticles synthesized under gamma irradiation for robust reluctance of resistant urinary tract infection-causing pathogens. Biol. Trace Elem. Res. 2020, 195, 323–342. [Google Scholar] [CrossRef]
  109. Yazhiniprabha, M.; Vaseeharan, B. In vitro and in vivo toxicity assessment of selenium nanoparticles with significant larvicidal and bacteriostatic properties. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 103, 109763. [Google Scholar] [CrossRef]
  110. Yu, B.; You, P.; Song, M.; Zhou, Y.; Yu, F.; Zheng, W. A facile and fast synthetic approach to create selenium nanoparticles with diverse shapes and their antioxidation ability. New J. Chem. 2016, 40, 1118–1123. [Google Scholar] [CrossRef]
  111. Panahi-Kalamuei, M.; Salavati-Niasari, M.; Hosseinpour-Mashkani, S.M. Facile microwave synthesis, characterization, and solar cell application of selenium nanoparticles. J. Alloys Compd. 2014, 617, 627–632. [Google Scholar] [CrossRef]
  112. El-Batal, A.I.; Mosallam, F.M.; Ghorab, M.; Hanora, A.; Gobara, M.; Baraka, A.; Elsayed, M.A.; Pal, K.; Fathy, R.M.; Elkodous, M.A.; et al. Factorial design-optimized and gamma irradiation-assisted fabrication of selenium nanoparticles by chitosan and Pleurotus ostreatus fermented fenugreek for a vigorous in vitro effect against carcinoma cells. Int. J. Biol. Macromol. 2020, 156, 1584–1599. [Google Scholar] [CrossRef] [PubMed]
  113. Lara, H.H.; Guisbiers, G.; Mendoza, J.; Mimun, L.C.; A Vincent, B.; Lopez-Ribot, J.L.; Nash, K.L. Synergistic antifungal effect of chitosan-stabilized selenium nanoparticles synthesized by pulsed laser ablation in liquids against Candida albicans biofilms. Int. J. Nanomed. 2018, 13, 2697–2708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Menazea, A.; Ismail, A.; Awwad, N.S.; Ibrahium, H.A. Physical characterization and antibacterial activity of PVA/Chitosan matrix doped by selenium nanoparticles prepared via one-pot laser ablation route. J. Mater. Res. Technol. 2020, 9, 9598–9606. [Google Scholar] [CrossRef]
  115. Guisbiers, G.; Lara, H.H.; Mendoza-Cruz, R.; Naranjo, G.; Vincent, B.A.; Peralta, X.G.; Nash, K.L. Inhibition of Candida albicans biofilm by pure selenium nanoparticles synthesized by pulsed laser ablation in liquids. Nanomed. Nanotechnol. Biol. Med. 2017, 13, 1095–1103. [Google Scholar] [CrossRef] [Green Version]
  116. Guisbiers, G.; Wang, Q.; Khachatryan, E.; Mimun, L.C.; Mendoza-Cruz, R.; Larese-Casanova, P.; Webster, T.J.; Nash, K.L. Inhibition of E. coli and S. aureus with selenium nanoparticles synthesized by pulsed laser ablation in deionized water. Int. J. Nanomed. 2016, 11, 3731–3736. [Google Scholar] [CrossRef] [Green Version]
  117. Luesakul, U.; Komenek, S.; Puthong, S.; Muangsin, N. Shape-controlled synthesis of cubic-like selenium nanoparticles via the self-assembly method. Carbohydr. Polym. 2016, 153, 435–444. [Google Scholar] [CrossRef]
  118. Panahi-Kalamuei, M.; Mousavi-Kamazani, M.; Salavati-Niasari, M.; Hosseinpour-Mashkani, S.M. A simple sonochemical approach for synthesis of selenium nanostructures and investigation of its light harvesting application. Ultrason. Sonochem. 2015, 23, 246–256. [Google Scholar] [CrossRef]
  119. Maiyo, F.; Singh, M. Selenium nanoparticles: Potential in cancer gene and drug delivery. Nanomedicine 2017, 12, 1075–1089. [Google Scholar] [CrossRef]
  120. Kumar, A.; Sevonkaev, I.; Goia, D.V. Synthesis of selenium particles with various morphologies. J. Colloid Interface Sci. 2014, 416, 119–123. [Google Scholar] [CrossRef]
  121. Yu, B.; Zhang, Y.; Zheng, W.; Fan, C.; Chen, T. Positive surface charge enhances selective cellular uptake and anticancer efficacy of selenium nanoparticles. Inorg. Chem. 2012, 51, 8956–8963. [Google Scholar] [CrossRef] [PubMed]
  122. Chung, S.; Zhou, R.; Webster, T.J. Green synthesized BSA-coated selenium nanoparticles inhibit bacterial growth while promoting mammalian cell growth. Int. J. Nanomed. 2020, 15, 115–124. [Google Scholar] [CrossRef] [PubMed]
  123. Webster, T.J.; Ramos, J.F. Cytotoxicity of selenium nanoparticles in rat dermal fibroblasts. Int. J. Nanomed. 2012, 7, 3907–3914. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Gangadoo, S.; Stanley, D.; Hughes, R.J.; Moore, R.J.; Chapman, J. The synthesis and characterisation of highly stable and reproducible selenium nanoparticles. Inorg. Nano-Metal Chem. 2017, 47, 1568–1576. [Google Scholar] [CrossRef]
  125. Ye, X.; Chen, L.; Liu, L.; Bai, Y. Electrochemical synthesis of selenium nanoparticles and formation of sea urchin-like selenium nanoparticles by electrostatic assembly. Mater. Lett. 2017, 196, 381–384. [Google Scholar] [CrossRef]
  126. Langi, B.; Shah, C.; Singh, K.; Chaskar, A.; Kumar, M.; Bajaj, P.N. Ionic liquid-induced synthesis of selenium nanoparticles. Mater. Res. Bull. 2010, 45, 668–671. [Google Scholar] [CrossRef]
  127. Zhang, S.; Luo, Y.; Zeng, H.; Wang, Q.; Tian, F.; Song, J.; Cheng, W.-H. Encapsulation of selenium in chitosan nanoparticles improves selenium availability and protects cells from selenium-induced DNA damage response. J. Nutr. Biochem. 2011, 22, 1137–1142. [Google Scholar] [CrossRef]
  128. Mittal, H.; Ray, S.S.; Kaith, B.S.; Bhatia, J.K.; Sukriti; Sharma, J.; Alhassan, S.M. Recent progress in the structural modification of chitosan for applications in diversified biomedical fields. Eur. Polym. J. 2018, 109, 402–434. [Google Scholar] [CrossRef]
  129. Bai, K.; Hong, B.; He, J.; Hong, Z.; Tan, R. Preparation and antioxidant properties of selenium nanoparticles-loaded chitosan microspheres. Int. J. Nanomed. 2017, 12, 4527–4539. [Google Scholar] [CrossRef] [Green Version]
  130. Luo, Y.; Zhang, B.; Cheng, W.-H.; Wang, Q. Preparation, characterization and evaluation of selenite-loaded chitosan/TPP nanoparticles with or without zein coating. Carbohydr. Polym. 2010, 82, 942–951. [Google Scholar] [CrossRef]
  131. Cui, D.; Yan, C.; Miao, J.; Zhang, X.; Chen, J.; Sun, L.; Meng, L.; Liang, T.; Li, Q. Synthesis, characterization and antitumor properties of selenium nanoparticles coupling with ferulic acid. Mater. Sci. Eng. C 2018, 90, 104–112. [Google Scholar] [CrossRef] [PubMed]
  132. Modrzejewska-Sikorska, A.; Konował, E.; Klapiszewski, Ł.; Nowaczyk, G.; Jurga, S.; Jesionowski, T.; Milczarek, G. Lignosulfonate-stabilized selenium nanoparticles and their deposition on spherical silica. Int. J. Biol. Macromol. 2017, 103, 403–408. [Google Scholar] [CrossRef] [PubMed]
  133. Medina Cruz, D.; Mi, G.; Webster, T.J. Synthesis and characterization of biogenic selenium nanoparticles with antimicrobial properties made by Staphylococcus aureus, methicillin-resistant Staphylococcus aureus (MRSA), Escherichia coli, and Pseudomonas aeruginos. J. Biomed. Mater. Res. A 2018, 106, 1400–1412. [Google Scholar] [CrossRef]
  134. Lampis, S.; Zonaro, E.; Bertolini, C.; Cecconi, D.; Monti, F.; Micaroni, M.; Turner, R.J.; Butler, C.S.; Vallini, G. Selenite biotransformation and detoxification by Stenotrophomonas maltophilia SeITE02: Novel clues on the route to bacterial biogenesis of selenium nanoparticles. J. Hazard. Mater. 2017, 324, 3–14. [Google Scholar] [CrossRef] [PubMed]
  135. Kora, A.J. Bacillus cereus, selenite-reducing bacterium from contaminated lake of an industrial area: A renewable nanofactory for the synthesis of selenium nanoparticles. Bioresour. Bioprocess. 2018, 5, 30. [Google Scholar] [CrossRef] [Green Version]
  136. Khoei, N.S.; Lampis, S.; Zonaro, E.; Yrjälä, K.; Bernardi, P.; Vallini, G. Insights into selenite reduction and biogenesis of elemental selenium nanoparticles by two environmental isolates of Burkholderia fungorum. New Biotechnol. 2017, 34, 1–11. [Google Scholar] [CrossRef] [PubMed]
  137. Kamnev, A.A.; Mamchenkova, P.V.; Dyatlova, Y.A.; Tugarova, A.V. FTIR spectroscopic studies of selenite reduction by cells of the rhizobacterium Azospirillum brasilense Sp7 and the formation of selenium nanoparticles. J. Mol. Struct. 2017, 1140, 106–112. [Google Scholar] [CrossRef]
  138. Espinosa-Ortiz, E.J.; Gonzalez-Gil, G.; Saikaly, P.; van Hullebusch, E.D.; Lens, P.N.L. Effects of selenium oxyanions on the white-rot fungus Phanerochaete chrysosporium. Appl. Microbiol. Biotechnol. 2015, 99, 2405–2418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Chasteen, T.G.; Bentley, R. Biomethylation of selenium and tellurium: Microorganisms and plants. Chem. Rev. 2003, 103, 1–26. [Google Scholar] [CrossRef]
  140. Liu, Y.; Zeng, S.; Liu, Y.; Wu, W.; Shen, Y.; Zhang, L.; Li, C.; Chen, H.; Liu, A.; Shen, L.; et al. Synthesis and antidiabetic activity of selenium nanoparticles in the presence of polysaccharides from Catathelasma ventricosum. Int. J. Biol. Macromol. 2018, 114, 632–639. [Google Scholar] [CrossRef]
  141. Gao, X.; Li, X.; Mu, J.; Ho, C.-T.; Su, J.; Zhang, Y.; Lin, X.; Chen, Z.; Li, B.; Xie, Y. Preparation, physicochemical characterization, and anti-proliferation of selenium nanoparticles stabilized by Polyporus umbellatus polysaccharide. Int. J. Biol. Macromol. 2020, 152, 605–615. [Google Scholar] [CrossRef] [PubMed]
  142. Vetchinkina, E.; Loshchinina, E.; Kursky, V.; Nikitina, V. Reduction of organic and inorganic selenium compounds by the edible medicinal basidiomycete Lentinula edodes and the accumulation of elemental selenium nanoparticles in its mycelium. J. Microbiol. 2013, 51, 829–835. [Google Scholar] [CrossRef] [PubMed]
  143. Vetchinkina, E.; Loshchinina, E.; Kupryashina, M.; Burov, A.; Nikitina, V. Shape and size diversity of gold, silver, selenium, and silica nanoparticles prepared by green synthesis using fungi and bacteria. Ind. Eng. Chem. Res. 2019, 58, 17207–17218. [Google Scholar] [CrossRef]
  144. Liang, X.; Perez, M.A.M.-J.; Nwoko, K.C.; Egbers, P.; Feldmann, J.; Csetenyi, L.; Gadd, G.M. Fungal formation of selenium and tellurium nanoparticles. Appl. Microbiol. Biotechnol. 2019, 103, 7241–7259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Liao, W.; Yu, Z.; Lin, Z.; Lei, Z.; Ning, Z.; Regenstein, J.M.; Yang, J.; Ren, J. Biofunctionalization of selenium nanoparticle with Dictyophora Indusiata polysaccharide and its antiproliferative activity through death-receptor and mitochondria-mediated apoptotic pathways. Sci. Rep. 2015, 5, 18629. [Google Scholar] [CrossRef] [Green Version]
  146. Zhang, H.; Zhou, H.; Bai, J.; Li, Y.; Yang, J.; Ma, Q.; Qu, Y. Biosynthesis of selenium nanoparticles mediated by fungus Mariannaea sp. HJ and their characterization. Colloids Surf. A Physicochem. Eng. Asp. 2019, 571, 9–16. [Google Scholar] [CrossRef]
  147. Joshi, S.; De Britto, S.; Jogaiah, S.; Ito, S.-I. Mycogenic selenium nanoparticles as potential new generation broad spectrum antifungal molecules. Biomolecules 2019, 9, 419. [Google Scholar] [CrossRef] [Green Version]
  148. Xiao, Y.; Huang, Q.; Zheng, Z.; Guan, H.; Liu, S. Construction of a Cordyceps sinensis exopolysaccharide-conjugated selenium nanoparticles and enhancement of their antioxidant activities. Int. J. Biol. Macromol. 2017, 99, 483–491. [Google Scholar] [CrossRef]
  149. Mosallam, F.M.; El-Sayyad, G.S.; Fathy, R.M.; El-Batal, A.I. Biomolecules-mediated synthesis of selenium nanoparticles using Aspergillus oryzae fermented Lupin extract and gamma radiation for hindering the growth of some multidrug-resistant bacteria and pathogenic fungi. Microb. Pathog. 2018, 122, 108–116. [Google Scholar] [CrossRef]
  150. Faramarzi, S.; Anzabi, Y.; Jafarizadeh-Malmiri, H. Nanobiotechnology approach in intracellular selenium nanoparticle synthesis using Saccharomyces cerevisiae—Fabrication and characterization. Arch. Microbiol. 2020, 202, 1203–1209. [Google Scholar] [CrossRef]
  151. Bartosiak, M.; Giersz, J.; Jankowski, K. Analytical monitoring of selenium nanoparticles green synthesis using photochemical vapor generation coupled with MIP-OES and UV–Vis spectrophotometry. Microchem. J. 2019, 145, 1169–1175. [Google Scholar] [CrossRef]
  152. Zhang, L.; Li, D.; Gao, P. Expulsion of selenium/protein nanoparticles through vesicle-like structures by Saccharomyces cerevisiae under microaerophilic environment. World J. Microbiol. Biotechnol. 2012, 28, 3381–3386. [Google Scholar] [CrossRef] [PubMed]
  153. Elahian, F.; Reiisi, S.; Shahidi, A.; Mirzaei, S.A. High-throughput bioaccumulation, biotransformation, and production of silver and selenium nanoparticles using genetically engineered Pichia pastoris. Nanomedicine 2017, 13, 853–861. [Google Scholar] [CrossRef] [PubMed]
  154. Rasouli, M. Biosynthesis of selenium nanoparticles using yeast Nematospora coryli and examination of their anti-candida and anti-oxidant activities. IET Nanobiotechnol. 2019, 13, 214–218. [Google Scholar] [CrossRef] [PubMed]
  155. Gunti, L.; Dass, R.S.; Kalagatur, N.K. Phytofabrication of selenium nanoparticles from Emblica officinalis fruit extract and exploring its biopotential applications: Antioxidant, antimicrobial, and biocompatibility. Front. Microbiol. 2019, 10, 931. [Google Scholar] [CrossRef] [Green Version]
  156. Hassanien, R.; Abed-Elmageed, A.A.I.; Husein, D.Z. Eco-friendly approach to synthesize selenium nanoparticles: Photocatalytic degradation of sunset yellow azo dye and anticancer activity. ChemistrySelect 2019, 4, 9018–9026. [Google Scholar] [CrossRef]
  157. Hu, T.; Li, H.; Li, J.; Zhao, G.; Wu, W.; Liu, L.; Wang, Q.; Guo, Y. Absorption and bio-transformation of selenium nanoparticles by wheat seedlings (Triticum aestivum L.). Front. Plant Sci. 2018, 9, 597. [Google Scholar] [CrossRef] [Green Version]
  158. Fardsadegh, B.; Vaghari, H.; Mohammad-Jafari, R.; Najian, Y.; Jafarizadeh-Malmiri, H. Biosynthesis, characterization and antimicrobial activities assessment of fabricated selenium nanoparticles using Pelargonium zonale leaf extract. Green Process. Synth. 2019, 8, 191–198. [Google Scholar] [CrossRef]
  159. Liang, T.; Qiu, X.; Ye, X.; Liu, Y.; Li, Z.; Tian, B.; Yan, D. Biosynthesis of selenium nanoparticles and their effect on changes in urinary nanocrystallites in calcium oxalate stone formation. 3 Biotech 2020, 10, 23. [Google Scholar] [CrossRef]
  160. Alagesan, V.; Venugopal, S. Green synthesis of selenium sanoparticle using leaves extract of Withania somnifera and its biological applications and photocatalytic activities. Bionanoscience 2019, 9, 105–116. [Google Scholar] [CrossRef] [Green Version]
  161. Alam, H.; Khatoon, N.; Raza, M.; Ghosh, P.C.; Sardar, M. Synthesis and characterization of nano selenium using plant biomolecules and their potential applications. Bionanoscience 2018, 9, 96–104. [Google Scholar] [CrossRef]
  162. Anu, K.; Devanesan, S.; Prasanth, R.; AlSalhi, M.S.; Ajithkumar, S.; Singaravelu, G. Biogenesis of selenium nanoparticles and their anti-leukemia activity. J. King Saud Univ. Sci. 2020, 32, 2520–2526. [Google Scholar] [CrossRef]
  163. Prasad, K.S.; Patel, H.; Patel, T.; Patel, K.; Selvaraj, K. Biosynthesis of Se nanoparticles and its effect on UV-induced DNA damage. Colloids Surf. B Biointerfaces 2013, 103, 261–266. [Google Scholar] [CrossRef]
  164. Hashem, A.H.; Selim, T.A.; Alruhaili, M.H.; Selim, S.; Alkhalifah, D.H.M.; Al Jaouni, S.K.; Salem, S.S. Unveiling antimicrobial and insecticidal activities of biosynthesized selenium nanoparticles using prickly pear peel waste. J. Funct. Biomater. 2022, 13, 112. [Google Scholar] [CrossRef]
  165. Ghaderi, R.S.; Adibian, F.; Sabouri, Z.; Davoodi, J.; Kazemi, M.; Jamehdar, S.A.; Meshkat, Z.; Soleimanpour, S.; Daroudi, M. Green synthesis of selenium nanoparticle by Abelmoschus esculentus extract and assessment of its antibacterial activity. Mater. Technol. 2021, 37, 1289–1297. [Google Scholar] [CrossRef]
  166. Salem, S.S.; Badawy, M.S.E.M.; Al-Askar, A.A.; Arishi, A.A.; Elkady, F.M.; Hashem, A.H. Green biosynthesis of selenium nanoparticles using orange peel waste: Characterization, antibacterial and antibiofilm activities against multidrug-resistant bacteria. Life 2022, 12, 893. [Google Scholar] [CrossRef] [PubMed]
  167. Liu, C.; Li, Y.; Li, H.; Wang, Y.; Zhao, K. Nano-selenium and Macleaya cordata extracts improved immune functions of intrauterine growth retardation piglets under maternal oxidation stress. Biol. Trace Elem. Res. 2022, 200, 3975–3982. [Google Scholar] [CrossRef]
  168. Fan, D.; Li, L.; Li, Z.; Zhang, Y.; Ma, X.; Wu, L.; Zhang, H.; Guo, F. Biosynthesis of selenium nanoparticles and their protective, antioxidative effects in streptozotocin induced diabetic rats. Sci. Technol. Adv. Mater. 2020, 21, 505–514. [Google Scholar] [CrossRef]
  169. Wang, L.; Li, C.; Huang, Q.; Fu, X. Biofunctionalization of selenium nanoparticles with a polysaccharide from Rosa roxburghii fruit and their protective effect against H2O2-induced apoptosis in INS-1 cells. Food Funct. 2019, 10, 539–553. [Google Scholar] [CrossRef]
  170. Sowndarya, P.; Ramkumar, G.; Shivakumar, M.S. Green synthesis of selenium nanoparticles conjugated Clausena dentata plant leaf extract and their insecticidal potential against mosquito vectors. Artif. Cells Nanomed. Biotechnol. 2017, 45, 1490–1495. [Google Scholar] [CrossRef] [Green Version]
  171. Dahoumane, S.A.; Jeffryes, C.; Mechouet, M.; Agathos, S.N. Biosynthesis of inorganic nanoparticles: A fresh look at the control of shape, size and composition. Bioengineering 2017, 4, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Wu, Z.; Yang, S.; Wu, W. Shape control of inorganic nanoparticles from solution. Nanoscale 2016, 8, 1237–1259. [Google Scholar] [CrossRef] [PubMed]
  173. Meija, J.; Coplen, T.B.; Berglund, M.; Brand, W.A.; De Bièvre, P.; Gröning, M.; Holden, N.E.; Irrgeher, J.; Loss, R.D.; Walczyk, T.; et al. Atomic weights of the elements 2013 (IUPAC Technical Report). Pure Appl. Chem. 2016, 88, 265–291. [Google Scholar] [CrossRef] [Green Version]
  174. Nayak, V.; Singh, K.R.; Singh, A.K.; Singh, R.P. Potentialities of selenium nanoparticles in biomedical science. New J. Chem. 2021, 45, 2849–2878. [Google Scholar] [CrossRef]
  175. Gates, B.; Yin, Y.; Xia, Y. A solution-phase approach to the synthesis of uniform nanowires of crystalline selenium with lateral dimensions in the range of 10−30 nm. J. Am. Chem. Soc. 2000, 122, 12582–12583. [Google Scholar] [CrossRef]
  176. Johnson, J.A.; Saboungi, M.-L.; Thiyagarajan, P.; Csencsits, R.; Meisel, D. Selenium nanoparticles: A small-angle neutron scattering study. J. Phys. Chem. B 1999, 103, 59–63. [Google Scholar] [CrossRef]
  177. Xiong, S.; Xi, B.; Wang, W.; Wang, C.; Fei, L.; Zhou, H.; Qian, Y. The Fabrication and characterization of single-crystalline selenium nanoneedles. Cryst. Growth Des. 2006, 6, 1711–1716. [Google Scholar] [CrossRef]
  178. Zhou, X.; Gao, P.; Sun, S.; Bao, D.; Wang, Y.; Li, X.; Wu, T.; Chen, Y.; Yang, P. Amorphous, crystalline and crystalline/amorphous selenium nanowires and their different (de)lithiation mechanisms. Chem. Mater. 2015, 27, 6730–6736. [Google Scholar] [CrossRef]
  179. Di, Y.; Zeng, Q.; Huang, C.; Tang, D.; Sun, K.; Yan, C.; Wang, Y.; Ke, S.; Jiang, L.; Hao, X.; et al. Thermal-evaporated selenium as a hole-transporting material for planar perovskite solar cells. Sol. Energy Mater. Sol. Cells 2018, 185, 130–135. [Google Scholar] [CrossRef]
  180. Cui, Y.; Abouimrane, A.; Lu, J.; Bolin, T.; Ren, Y.; Weng, W.; Sun, C.; Maroni, V.A.; Heald, S.M.; Amine, K. (De)lithiation mechanism of Li/SeSx (x = 0–7) batteries determined by in situ synchrotron X-ray diffraction and X-ray absorption spectroscopy. J. Am. Chem. Soc. 2013, 135, 8047–8056. [Google Scholar] [CrossRef]
  181. Jiang, Y.; Ma, X.; Feng, J.; Xiong, S. Selenium in nitrogen-doped microporous carbon spheres for high-performance lithium–selenium batteries. J. Mater. Chem. A 2015, 3, 4539–4546. [Google Scholar] [CrossRef]
  182. Kubota, M.; Kato, T.; Suzuki, S.; Maruyama, H.; Shidara, K.; Tanioka, K.; Sameshima, K.; Makishima, T.; Tsuji, K.; Hirai, T.; et al. Ultrahigh-sensitivity new super-HARP camera. IEEE Trans. Broadcast. 1996, 42, 251–258. [Google Scholar] [CrossRef]
  183. Yamazaki, J.; Maruyama, H.; Andoh, F.; Tanioka, K.; Araki, S.; Tsuji, K.; Kawamura, T. A solid-state imager joined to an avalanche multiplier film with micro-bump electrodes. In Proceedings of the International Electron Devices Meeting, Washington, DC, USA, 10–13 December 1995; pp. 175–178. [Google Scholar] [CrossRef]
  184. Imura, S.; Watabe, T.; Miyakawa, K.; Hagiwara, K.; Ohtake, H.; Kubota, M. Effects of grain refinement on surface enhancement of thin-film chlorine-doped crystalline selenium. J. Mater. Sci. Mater. Electron. 2017, 28, 7064–7069. [Google Scholar] [CrossRef]
  185. Takiguchi, Y.; Maruyama, H.; Kosugi, M.; Andoh, F.; Kato, T.; Tanioka, K.; Yamazaki, J.; Tsuji, K.; Kawamura, T. A CMOS imager hybridized to an avalanche multiplied film. IEEE Trans. Electron Devices 1997, 44, 1783–1788. [Google Scholar] [CrossRef]
  186. Liu, P.; Ma, Y.; Cai, W.; Wang, Z.; Wang, J.; Qi, L.; Chen, D. Photoconductivity of single-crystalline selenium nanotubes. Nanotechnology 2007, 18, 205704. [Google Scholar] [CrossRef]
  187. Chen, W.; Li, Y.; Yang, S.; Yue, L.; Jiang, Q.; Xia, W. Synthesis and antioxidant properties of chitosan and carboxymethyl chitosan-stabilized selenium nanoparticles. Carbohydr. Polym. 2015, 132, 574–581. [Google Scholar] [CrossRef]
  188. Hageman, S.P.; van der Weijden, R.D.; Stams, A.J.; Buisman, C.J. Bio-production of selenium nanoparticles with diverse physical properties for recovery from water. Int. J. Miner. Process. 2017, 169, 7–15. [Google Scholar] [CrossRef]
  189. Jiang, F.; Cai, W.; Tan, G. Facile synthesis and optical properties of small selenium nanocrystals and nanorods. Nanoscale Res. Lett. 2017, 12, 401. [Google Scholar] [CrossRef] [Green Version]
  190. Frey, J.B.; Belev, G.; Tousignant, O.; Mani, H.; Laperriere, L.; Kasap, S.O. Dark current in multilayer stabilized amorphous selenium based photoconductive x-ray detectors. J. Appl. Phys. 2012, 112, 14502. [Google Scholar] [CrossRef]
  191. Kasap, S.; Frey, J.B.; Belev, G.; Tousignant, O.; Mani, H.; Laperriere, L.; Reznik, A.; Rowlands, J.A. Amorphous selenium and its alloys from early xeroradiography to high resolution X-ray image detectors and ultrasensitive imaging tubes. Phys. Status solidi (B) 2009, 246, 1794–1805. [Google Scholar] [CrossRef]
  192. Kannan, H.; Stavro, J.; Mukherjee, A.; Léveillé, S.; Kisslinger, K.; Guan, L.; Zhao, W.; Sahu, A.; Goldan, A.H. Ultralow dark currents in avalanche amorphous selenium photodetectors using solution-processed quantum dot blocking layer. ACS Photon. 2020, 7, 1367–1374. [Google Scholar] [CrossRef]
  193. Imura, S.; Mineo, K.; Miyakawa, K.; Nanba, M.; Ohtake, H.; Kubota, M. Low-dark-current photodiodes comprising highly (100)-oriented hexagonal selenium with crystallinity-enhanced tellurium nucleation layers. IEEE Sens. J. 2018, 18, 3108–3113. [Google Scholar] [CrossRef]
  194. Wronski, M.M.; Rowlands, J.A. Direct-conversion flat-panel imager with avalanche gain: Feasibility investigation for HARP-AMFPI. Med. Phys. 2008, 35, 5207–5218. [Google Scholar] [CrossRef] [Green Version]
  195. Zhao, W.; Li, D.; Reznik, A.; Lui, B.J.M.; Hunt, D.C.; Rowlands, J.A.; Ohkawa, Y.; Tanioka, K. Indirect flat-panel detector with avalanche gain: Fundamental feasibility investigation for SHARP-AMFPI (scintillator HARP active matrix flat panel imager). Med. Phys. 2005, 32, 2954–2966. [Google Scholar] [CrossRef]
  196. Okano, K.; Saito, I.; Mine, T.; Suzuki, Y.; Yamada, T.; Rupesinghe, N.; Amaratunga, G.; Milne, W.; Zahn, D. Characterizations of a-Se based photodetectors using X-ray photoelectron spectroscopy and Raman spectroscopy. J. Non-Cryst. Solids 2007, 353, 308–312. [Google Scholar] [CrossRef]
  197. Oonuki, K.; Suzuki, Y.; Yamaguchi, H.; Okano, K.; Okamura, Y. Diode structure amorphous selenium photodetector with nitrogen (N)-doped diamond cold cathode. J. Vac. Sci. Technol. B Microelectron. Nanometer Struct. 2003, 21, 1586. [Google Scholar] [CrossRef]
  198. Suzuki, Y.; Yamaguchi, H.; Oonuki, K.; Okamura, Y.; Okano, K. Amorphous selenium photodetector driven by diamond cold cathode. IEEE Electron Device Lett. 2003, 24, 16–18. [Google Scholar] [CrossRef]
  199. Hadar, I.; Hu, X.; Luo, Z.-Z.; Dravid, V.P.; Kanatzidis, M.G. Nonlinear band gap tunability in selenium–tellurium alloys and its utilization in solar cells. ACS Energy Lett. 2019, 4, 2137–2143. [Google Scholar] [CrossRef]
  200. Platzer-Björkman, C.; Zabierowski, P.; Pettersson, J.; Törndahl, T.; Edoff, M. Improved fill factor and open circuit voltage by crystalline selenium at the Cu(In,Ga)Se2/buffer layer interface in thin film solar cells. Prog. Photovolt. Res. Appl. 2010, 18, 249–256. [Google Scholar] [CrossRef]
  201. Sharma, T.; Sharma, R.; Tamboli, R.A.; Kanhere, D.G. Ab initio investigation of structural and electronic properties of selenium and tellurium clusters. Eur. Phys. J. B 2019, 92, 51. [Google Scholar] [CrossRef]
  202. Wang, L.; Kafshgari, M.H.; Meunier, M. Optical properties and applications of plasmonic-metal nanoparticles. Adv. Funct. Mater. 2020, 30, 2005400. [Google Scholar] [CrossRef]
  203. Huang, X.; Zhu, Y.; Kianfar, E. Nano Biosensors: Properties, applications and electrochemical techniques. J. Mater. Res. Technol. 2021, 12, 1649–1672. [Google Scholar] [CrossRef]
  204. Jara, N.; Milán, N.; Rahman, A.; Mouheb, L.; Boffito, D.; Jeffryes, C.; Dahoumane, S. Photochemical synthesis of gold and silver nanoparticles—A review. Molecules 2021, 26, 4585. [Google Scholar] [CrossRef]
  205. Huang, J.; Liu, J.; Wang, J. Optical properties of biomass-derived nanomaterials for sensing, catalytic, biomedical and environmental applications. TrAC Trends Anal. Chem. 2020, 124, 115800. [Google Scholar] [CrossRef]
  206. Lesnichaya, M.; Shendrik, R.; Sukhov, B. Relation between excitation dependent luminescence and particle size distributions for the selenium nanoparticles in κ-carrageenan shell. J. Lumin. 2019, 211, 305–313. [Google Scholar] [CrossRef]
  207. Rajalakshmi, M.; Arora, A. Optical properties of selenium nanoparticles dispersed in polymer. Solid State Commun. 1999, 110, 75–80. [Google Scholar] [CrossRef]
  208. Singh, S.C.; Mishra, S.K.; Srivastava, R.K.; Gopal, R. Optical properties of selenium quantum dots produced with laser irradiation of water suspended Se nanoparticles. J. Phys. Chem. C 2010, 114, 17374–17384. [Google Scholar] [CrossRef]
  209. Sharma, G.; Sharma, A.R.; Bhavesh, R.; Park, J.; Ganbold, B.; Nam, J.-S.; Lee, S.-S. Biomolecule-mediated synthesis of selenium nanoparticles using dried Vitis vinifera (raisin) extract. Molecules 2014, 19, 2761–2770. [Google Scholar] [CrossRef]
  210. Piacenza, E.; Presentato, A.; Zonaro, E.; Lampis, S.; Vallini, G.; Turner, R.J. Selenium and tellurium nanomaterials. Phys. Sci. Rev. 2018, 3, 20170100. [Google Scholar] [CrossRef]
  211. Zhang, W.; Chen, Z.; Liu, H.; Zhang, L.; Gao, P.; Li, D. Biosynthesis and structural characteristics of selenium nanoparticles by Pseudomonas alcaliphila. Colloids Surf. B Biointerfaces 2011, 88, 196–201. [Google Scholar] [CrossRef]
  212. Colmenares, L.; Jusys, Z.; Behm, R.J. Activity, selectivity, and methanol tolerance of Se-modified Ru/C cathode catalysts. J. Phys. Chem. C 2007, 111, 1273–1283. [Google Scholar] [CrossRef]
  213. Zehl, G.; Bogdanoff, P.; Dorbandt, I.; Fiechter, S.; Wippermann, K.; Hartnig, C. Carbon supported Ru–Se as methanol tolerant catalysts for DMFC cathodes. Part I: Preparation and characterization of catalysts. J. Appl. Electrochem. 2007, 37, 1475–1484. [Google Scholar] [CrossRef]
  214. Chaudhary, S.; Umar, A.; Mehta, S.K. Surface functionalized selenium nanoparticles for biomedical applications. J. Biomed. Nanotechnol. 2014, 10, 3004–3042. [Google Scholar] [CrossRef]
  215. Dumore, N.S.; Mukhopadhyay, M. Antioxidant properties of aqueous selenium nanoparticles (ASeNPs) and its catalysts activity for 1, 1-diphenyl-2-picrylhydrazyl (DPPH) reduction. J. Mol. Struct. 2019, 1205, 127637. [Google Scholar] [CrossRef]
  216. Jin, Z.; Zhang, M.; Wang, M.; Feng, C.; Wang, Z.-S. Metal selenides as efficient counter electrodes for dye-sensitized solar cells. Accounts Chem. Res. 2017, 50, 895–904. [Google Scholar] [CrossRef]
  217. Verjulio, R.W.; Santander, J.; Ma, J.; Alonso-Vante, N. Selective CoSe2/C cathode catalyst for passive air-breathing alkaline anion exchange membrane μ-direct methanol fuel cell (AEM-μDMFC). Int. J. Hydrogen Energy 2016, 41, 19595–19600. [Google Scholar] [CrossRef]
  218. Gurkan, Y.Y.; Kasapbasi, E.; Cinar, Z. Enhanced solar photocatalytic activity of TiO2 by selenium(IV) ion-doping: Characterization and DFT modeling of the surface. Chem. Eng. J. 2013, 214, 34–44. [Google Scholar] [CrossRef]
  219. Piacenza, E.; Presentato, A.; Heyne, B.; Turner, R.J. Tunable photoluminescence properties of selenium nanoparticles: Biogenic versus chemogenic synthesis. Nanophotonics 2020, 9, 3615–3628. [Google Scholar] [CrossRef]
  220. Srivastava, N.; Mukhopadhyay, M. Biosynthesis of SnO2 nanoparticles using bacterium Erwinia herbicola and their photocatalytic activity for degradation of dyes. Ind. Eng. Chem. Res. 2014, 53, 13971–13979. [Google Scholar] [CrossRef]
  221. Kumari, C.; Sharma, P.; Katyal, S.; Tanwar, M.; Bamola, P.; Sharma, H.; Kumar, R.; Chhoker, S. Photocatalytic activity of GeSbSeEr quaternary chalcogenide for efficient methylene blue degradation in visible light. Results Surf. Interfaces 2022, 9, 100088. [Google Scholar] [CrossRef]
  222. Al Jahdaly, B.A.; Al-Radadi, N.S.; Eldin, G.M.; Almahri, A.; Ahmed, M.; Shoueir, K.; Janowska, I. Selenium nanoparticles synthesized using an eco-friendly method: Dye decolorization from aqueous solutions, cell viability, antioxidant, and antibacterial effectiveness. J. Mater. Res. Technol. 2021, 11, 85–97. [Google Scholar] [CrossRef]
  223. Altaf, S.; Ajaz, H.; Imran, M.; Ul-Hamid, A.; Naz, M.; Aqeel, M.; Shahzadi, A.; Shahbaz, A.; Ikram, M. Synthesis and characterization of binary selenides of transition metals to investigate its photocatalytic, antimicrobial and anticancer efficacy. Appl. Nanosci. 2020, 10, 2113–2127. [Google Scholar] [CrossRef]
  224. Talukdar, S.; Dutta, R.K. A mechanistic approach for superoxide radicals and singlet oxygen mediated enhanced photocatalytic dye degradation by selenium doped ZnS nanoparticles. RSC Adv. 2016, 6, 928–936. [Google Scholar] [CrossRef]
  225. Kaviyarasu, K.; Ayeshamariam, A.; Manikandan, E.; Kennedy, J.; Ladchumananandasivam, R.; Gomes, U.U.; Jayachandran, M.; Maaza, M. Solution processing of CuSe quantum dots: Photocatalytic activity under RhB for UV and visible-light solar irradiation. Mater. Sci. Eng. B 2016, 210, 1–9. [Google Scholar] [CrossRef]
  226. Karthikeyan, N.; Sivaranjani, T.; Dhanavel, S.; Gupta, V.; Narayanan, V.; Stephen, A. Visible light degradation of textile effluent by electrodeposited multiphase CuInSe2 semiconductor photocatalysts. J. Mol. Liq. 2017, 227, 194–201. [Google Scholar] [CrossRef]
  227. Li, J.; Yang, M.; Feng, Z.; Lan, F.; Teng, C.; Li, D.; Tang, J. A new HgSe photocatalyst for degradation of organic dyes under visible light irradiation. Mater. Lett. 2015, 161, 591–594. [Google Scholar] [CrossRef]
  228. Ameri, A.; Shakibaie, M.; Pournamdari, M.; Ameri, A.; Foroutanfar, A.; Doostmohammadi, M.; Forootanfar, H. Degradation of diclofenac sodium using UV/biogenic selenium nanoparticles/H2O2: Optimization of process parameters. J. Photochem. Photobiol. A Chem. 2020, 392, 112382. [Google Scholar] [CrossRef]
  229. Chiou, Y.-D.; Hsu, Y.-J. Room-temperature synthesis of single-crystalline Se nanorods with remarkable photocatalytic properties. Appl. Catal. B Environ. 2011, 105, 211–219. [Google Scholar] [CrossRef]
  230. Tripathi, R.M.; Hameed, P.; Rao, R.P.; Shrivastava, N.; Mittal, J.; Mohapatra, S. Biosynthesis of highly stable fluorescent selenium nanoparticles and the evaluation of their photocatalytic degradation of dye. Bionanoscience 2020, 10, 389–396. [Google Scholar] [CrossRef]
  231. Zhang, R.; Tian, X.; Ma, L.; Yang, C.; Zhou, Z.; Wang, Y.; Wang, S. Visible-light-responsive t-Se nanorod photocatalysts: Synthesis, properties, and mechanism. RSC Adv. 2015, 5, 45165–45171. [Google Scholar] [CrossRef]
  232. Che, L.; Dong, Y.; Wu, M.; Zhao, Y.; Liu, L.; Zhou, H. Characterization of selenite reduction by Lysinibacillus sp. ZYM-1 and photocatalytic performance of biogenic selenium nanospheres. ACS Sustain. Chem. Eng. 2017, 5, 2535–2543. [Google Scholar] [CrossRef]
  233. Song, L.; Chen, C.; Zhang, S. Preparation and photocatalytic activity of visible light-sensitive selenium-doped bismuth sulfide. Powder Technol. 2011, 207, 170–174. [Google Scholar] [CrossRef]
  234. Sharma, A.; Dutta, R.K. Se-doped CuO NPs/H2O2/UV as a highly efficient and sustainable photo-fenton catalytic system for enhanced degradation of 4-bromophenol. J. Clean. Prod. 2018, 185, 464–475. [Google Scholar] [CrossRef]
  235. Huang, L.; Tong, X.; Li, Y.; Teng, J.; Bai, Y. Preparation of a novel supported selenium nanoparticles adsorbent and its application for copper removal from aqueous solution. J. Chem. Eng. Data 2014, 60, 151–160. [Google Scholar] [CrossRef]
  236. Bai, Y.; Rong, F.; Wang, H.; Zhou, Y.; Xie, X.; Teng, J. Removal of copper from aqueous solutions by adsorption on elemental selenium nanoparticles. J. Chem. Eng. Data 2011, 56, 2563–2568. [Google Scholar] [CrossRef]
  237. Jain, R.; Jordan, N.; Schild, D.; van Hullebusch, E.D.; Weiss, S.; Franzen, C.; Farges, F.; Hübner, R.; Lens, P.N. Adsorption of zinc by biogenic elemental selenium nanoparticles. Chem. Eng. J. 2015, 260, 855–863. [Google Scholar] [CrossRef] [Green Version]
  238. Yuan, F.; Song, C.; Sun, X.; Tan, L.; Wang, Y.; Wang, S. Adsorption of Cd(II) from aqueous solution by biogenic selenium nanoparticles. RSC Adv. 2016, 6, 15201–15209. [Google Scholar] [CrossRef]
  239. Inukai, J.; Cao, D.; Wieckowski, A.; Chang, K.-C.; Menzel, A.; Komanicky, V.; You, H. In situ synchrotron X-ray spectroscopy of ruthenium nanoparticles modified with selenium for an oxygen reduction reaction. J. Phys. Chem. C 2007, 111, 16889–16894. [Google Scholar] [CrossRef]
  240. Kieliszek, M. Selenium-fascinating microelement, properties and sources in food. Molecules 2019, 24, 1298. [Google Scholar] [CrossRef] [Green Version]
  241. Barchielli, G.; Capperucci, A.; Tanini, D. The role of selenium in pathologies: An updated review. Antioxidants 2022, 11, 251. [Google Scholar] [CrossRef]
  242. Bjørklund, G.; Shanaida, M.; Lysiuk, R.; Antonyak, H.; Klishch, I.; Shanaida, V.; Peana, M. Selenium: An antioxidant with a critical role in anti-aging. Molecules 2022, 27, 6613. [Google Scholar] [CrossRef]
  243. Hu, W.; Zhao, C.; Hu, H.; Yin, S. Food sources of selenium and its relationship with chronic diseases. Nutrients 2021, 13, 1739. [Google Scholar] [CrossRef]
  244. Kohrle, J.; Jakob, F.; Contempre, B.; Dumont, J.E. Selenium, the thyroid, and the endocrine system. Endocr. Rev. 2005, 26, 944–984. [Google Scholar] [CrossRef]
  245. Gereben, B.; McAninch, E.A.; Ribeiro, M.O.; Bianco, A.C. Scope and limitations of iodothyronine deiodinases in hypothyroidism. Nat. Rev. Endocrinol. 2015, 11, 642–652. [Google Scholar] [CrossRef] [Green Version]
  246. Guillin, O.M.; Vindry, C.; Ohlmann, T.; Chavatte, L. Selenium, selenoproteins and viral infection. Nutrients 2019, 11, 2101. [Google Scholar] [CrossRef] [Green Version]
  247. Orozco, A.; Valverde, R.C.; Olvera, A.; Garcia, G.C. Iodothyronine deiodinases: A functional and evolutionary perspective. J. Endocrinol. 2012, 215, 207–219. [Google Scholar] [CrossRef] [Green Version]
  248. Jeong, D.; Lee, K.H. Bimodal actions of selenium essential for antioxidant and toxic pro-oxidant activities: The selenium paradox (Review). Mol. Med. Rep. 2012, 5, 299–304. [Google Scholar] [CrossRef]
  249. Alam, M.N.; Bristi, N.J.; Rafiquzzaman, M. Review on in vivo and in vitro methods evaluation of antioxidant activity. Saudi Pharm. J. 2013, 21, 143–152. [Google Scholar] [CrossRef] [Green Version]
  250. Hariharan, S.; Dharmaraj, S. Selenium and selenoproteins: It’s role in regulation of inflammation. Inflammopharmacology 2020, 28, 667–695. [Google Scholar] [CrossRef]
  251. Qin, S.; Huang, B.; Ma, J.; Wang, X.; Zhang, J.; Li, L.; Chen, F. Effects of selenium-chitosan on blood selenium concentration, antioxidation status, and cellular and humoral immunity in mice. Biol. Trace Elem. Res. 2015, 165, 145–152. [Google Scholar] [CrossRef]
  252. Bermingham, E.N.; Hesketh, J.E.; Sinclair, B.R.; Koolaard, J.P.; Roy, N.C. Selenium-enriched foods are more effective at increasing glutathione peroxidase (GPx) activity compared with selenomethionine: A meta-analysis. Nutrients 2014, 6, 4002–4031. [Google Scholar] [CrossRef] [PubMed]
  253. Singh, A.; Kukreti, R.; Saso, L.; Kukreti, S. Oxidative stress: A key modulator in neurodegenerative diseases. Molecules 2019, 24, 1583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Valko, M.; Leibfritz, D.; Moncol, J.; Cronin, M.T.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]
  255. Cole-Ezea, P.; Swan, D.; Shanley, D.; Hesketh, J. Glutathione peroxidase 4 has a major role in protecting mitochondria from oxidative damage and maintaining oxidative phosphorylation complexes in gut epithelial cells. Free Radic. Biol. Med. 2012, 53, 488–497. [Google Scholar] [CrossRef] [PubMed]
  256. Weekley, C.M.; Aitken, J.B.; Finney, L.; Vogt, S.; Witting, P.K.; Harris, H.H. Selenium metabolism in cancer cells: The combined application of XAS and XFM techniques to the problem of selenium speciation in biological systems. Nutrients 2013, 5, 1734–1756. [Google Scholar] [CrossRef] [Green Version]
  257. Yeo, J.E.; Kang, S.K. Selenium effectively inhibits ROS-mediated apoptotic neural precursor cell death in vitro and in vivo in traumatic brain injury. Biochim. Biophys. Acta 2007, 1772, 1199–1210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  258. Ganther, H.E. Selenium metabolism, selenoproteins and mechanisms of cancer prevention: Complexities with thioredoxin reductase. Carcinogenesis 1999, 20, 1657–1666. [Google Scholar] [CrossRef] [Green Version]
  259. Tobe, R.; Mihara, H. Delivery of selenium to selenophosphate synthetase for selenoprotein biosynthesis. Biochim. Biophys. Acta Gen. Subj. 2018, 1862, 2433–2440. [Google Scholar] [CrossRef]
  260. Forootanfar, H.; Adeli-Sardou, M.; Nikkhoo, M.; Mehrbani, M.; Amirheidari, B.; Shahverdi, A.R.; Shakibaie, M. Antioxidant and cytotoxic effect of biologically synthesized selenium nanoparticles in comparison to selenium dioxide. J. Trace Elem. Med. Biol. 2014, 28, 75–79. [Google Scholar] [CrossRef]
  261. Huang, B.; Zhang, J.; Hou, J.; Chen, C. Free radical scavenging efficiency of nano-Se in vitro. Free Radic. Biol. Med. 2003, 35, 805–813. [Google Scholar] [CrossRef]
  262. Torres, S.K.; Campos, V.L.; León, C.G.; Rodríguez-Llamazares, S.M.; Rojas, S.M.; González, M.; Smith, C.; Mondaca, M.A. Biosynthesis of selenium nanoparticles by Pantoea agglomerans and their antioxidant activity. J. Nanopart. Res. 2012, 14, 1236. [Google Scholar] [CrossRef]
  263. Ferreira, C.A.; Ni, D.; Rosenkrans, Z.T.; Cai, W. Scavenging of reactive oxygen and nitrogen species with nanomaterials. Nano Res. 2018, 11, 4955–4984. [Google Scholar] [CrossRef]
  264. Rao, S.; Lin, Y.; Du, Y.; He, L.; Huang, G.; Chen, B.; Chen, T. Designing multifunctionalized selenium nanoparticles to reverse oxidative stress-induced spinal cord injury by attenuating ROS overproduction and mitochondria dysfunction. J. Mater. Chem. B 2019, 7, 2648–2656. [Google Scholar] [CrossRef]
  265. Lee, S.-C.; Lee, N.-H.; Patel, K.D.; Jun, S.-K.; Park, J.-H.; Knowles, J.C.; Kim, H.-W.; Lee, H.-H.; Lee, J.-H. A study on myogenesis by regulation of reactive oxygen species and cytotoxic activity by selenium nanoparticles. Antioxidants 2021, 10, 1727. [Google Scholar] [CrossRef]
  266. Varlamova, E.; Goltyaev, M.; Mal’Tseva, V.; Turovsky, E.; Sarimov, R.; Simakin, A.; Gudkov, S. Mechanisms of the cytotoxic effect of selenium nanoparticles in different human cancer cell lines. Int. J. Mol. Sci. 2021, 22, 7798. [Google Scholar] [CrossRef]
  267. Bhattacharjee, A.; Basu, A.; Biswas, J.; Sen, T.; Bhattacharya, S. Chemoprotective and chemosensitizing properties of selenium nanoparticle (Nano-Se) during adjuvant therapy with cyclophosphamide in tumor-bearing mice. Mol. Cell Biochem. 2017, 424, 13–33. [Google Scholar] [CrossRef]
  268. Luesakul, U.; Puthong, S.; Neamati, N.; Muangsin, N. pH-responsive selenium nanoparticles stabilized by folate-chitosan delivering doxorubicin for overcoming drug-resistant cancer cells. Carbohydr. Polym. 2018, 181, 841–850. [Google Scholar] [CrossRef]
  269. Huang, Y.; He, L.; Liu, W.; Fan, C.; Zheng, W.; Wong, Y.-S.; Chen, T. Selective cellular uptake and induction of apoptosis of cancer-targeted selenium nanoparticles. Biomaterials 2013, 34, 7106–7116. [Google Scholar] [CrossRef]
  270. Pi, J.; Jiang, J.; Cai, H.; Yang, F.; Jin, H.; Yang, P.; Cai, J.; Chen, Z.W. GE11 peptide conjugated selenium nanoparticles for EGFR targeted oridonin delivery to achieve enhanced anticancer efficacy by inhibiting EGFR-mediated PI3K/AKT and Ras/Raf/MEK/ERK pathways. Drug Deliv. 2017, 24, 1549–1564. [Google Scholar] [CrossRef] [Green Version]
  271. Chen, T.; Zheng, S.; Li, X.; Zhang, Y.; Wong, Y.-S. PEG-nanolized ultrasmall selenium nanoparticles overcome drug resistance in hepatocellular carcinoma HepG2 cells through induction of mitochondria dysfunction. Int. J. Nanomed. 2012, 7, 3939–3949. [Google Scholar] [CrossRef] [Green Version]
  272. Wei, D.; Yu, Y.; Zhang, X.; Wang, Y.; Chen, H.; Zhao, Y.; Wang, F.; Rong, G.; Wang, W.; Kang, X.; et al. Breaking the intracellular redox balance with diselenium nanoparticles for maximizing chemotherapy efficacy on patient-derived xenograft models. ACS Nano 2020, 14, 16984–16996. [Google Scholar] [CrossRef] [PubMed]
  273. Zhang, H.; Wang, T.; Liu, H.; Han, Y.; Zheng, Q.; Xu, Q.; Bao, B.; Xing, W.; Li, Z. Boost therapy of hepatocellular carcinoma by amplifying vicious cycle between mitochondrial oxidative stress and endoplasmic reticulum stress via biodegradable ultrasmall nanoparticles and old drug. Nano Today 2022, 46, 101601. [Google Scholar] [CrossRef]
  274. Qiao, L.; Yan, S.; Dou, X.; Song, X.; Chang, J.; Pi, S.; Zhang, X.; Xu, C. Biogenic selenium nanoparticles alleviate intestinal epithelial barrier damage through regulating endoplasmic reticulum stress-mediated mitophagy. Oxidative Med. Cell Longev. 2022, 2022, 3982613. [Google Scholar] [CrossRef] [PubMed]
  275. Jiang, Y.; Yu, X.; Su, C.; Zhao, L.; Shi, Y. Chitosan nanoparticles induced the antitumor effect in hepatocellular carcinoma cells by regulating ROS-mediated mitochondrial damage and endoplasmic reticulum stress. Artif. Cells Nanomed. Biotechnol. 2019, 47, 747–756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  276. Liu, H.J.; Qin, Y.; Zhao, Z.H.; Zhang, Y.; Yang, J.H.; Zhai, D.H.; Cui, F.; Luo, C.; Lu, M.X.; Liu, P.P.; et al. Lentinan-functionalized selenium nanoparticles target tumor cell mitochondria via TLR4/TRAF3/MFN1 pathway. Theranostics 2020, 10, 9083–9099. [Google Scholar] [CrossRef]
  277. Jiang, W.; Fu, Y.; Yang, F.; Yang, Y.; Liu, T.; Zheng, W.; Zeng, L.; Chen, T. Gracilaria lemaneiformis Polysaccharide as sntegrin-targeting surface decorator of selenium nanoparticles to achieve enhanced anticancer efficacy. ACS Appl. Mater. Interfaces 2014, 6, 13738–13748. [Google Scholar] [CrossRef]
  278. Guo, M.; Li, Y.; Lin, Z.; Zhao, M.; Xiao, M.; Wang, C.; Xu, T.; Xia, Y.; Zhu, B. Surface decoration of selenium nanoparticles with curcumin induced HepG2 cell apoptosis through ROS mediated p53 and AKT signaling pathways. RSC Adv. 2017, 7, 52456–52464. [Google Scholar] [CrossRef] [Green Version]
  279. Song, Z.; Liu, T.; Chen, T. Overcoming blood-brain barrier by HER2-targeted nanosystem to suppress glioblastoma cell migration, invasion and tumor growth. J. Mater. Chem. B 2018, 6, 568–579. [Google Scholar] [CrossRef]
  280. Xia, Y.; Zhong, J.; Zhao, M.; Tang, Y.; Han, N.; Hua, L.; Xu, T.; Wang, C.; Zhu, B. Galactose-modified selenium nanoparticles for targeted delivery of doxorubicin to hepatocellular carcinoma. Drug Deliv. 2019, 26, 1–11. [Google Scholar] [CrossRef] [Green Version]
  281. Shao, C.; Yu, Z.; Luo, T.; Zhou, B.; Song, Q.; Li, Z.; Yu, X.; Jiang, S.; Zhou, Y.; Dong, W.; et al. Chitosan-coated selenium nanoparticles attenuate PRRSV replication and ROS/JNK-mediated apoptosis in vitro. Int. J. Nanomed. 2022, 17, 3043–3054. [Google Scholar] [CrossRef]
  282. Al-Noshokaty, T.M.; Mesbah, N.M.; Abo-Elmatty, D.M.; Abulsoud, A.I.; Abdel-Hamed, A.R. Selenium nanoparticles overcomes sorafenib resistance in thioacetamide induced hepatocellular carcinoma in rats by modulation of mTOR, NF-κB pathways and LncRNA-AF085935/GPC3 axis. Life Sci. 2022, 303, 120675. [Google Scholar] [CrossRef]
  283. Vundela, S.R.; Kalagatur, N.K.; Nagaraj, A.; Kadirvelu, K.; Chandranayaka, S.; Kondapalli, K.; Hashem, A.; Abd_Allah, E.F.; Poda, S. Multi-biofunctional properties of phytofabricated selenium nanoparticles from Carica papaya fruit extract: Antioxidant, antimicrobial, antimycotoxin, anticancer, and biocompatibility. Front. Microbiol. 2021, 12, 769891. [Google Scholar] [CrossRef]
  284. Pi, J.; Yang, F.; Jin, H.; Huang, X.; Liu, R.; Yang, P.; Cai, J. Selenium nanoparticles induced membrane bio-mechanical property changes in MCF-7 cells by disturbing membrane molecules and F-actin. Bioorg. Med. Chem. Lett. 2013, 23, 6296–6303. [Google Scholar] [CrossRef]
  285. Zeebaree, S.Y.S.; Zebari, O.I.H. Diagnosis of the multiple effect of selenium nanoparticles decorated by Asteriscus graveolens components in inhibiting HepG2 cell proliferation. Sustain. Chem. Pharm. 2020, 15, 1002101. [Google Scholar] [CrossRef]
  286. Juan, C.; de la Lastra, J.P.; Plou, F.; Pérez-Lebeña, E. The Chemistry of reactive oxygen species (ROS) revisited: Outlining their role in biological macromolecules (DNA, lipids and proteins) and induced pathologies. Int. J. Mol. Sci. 2021, 22, 4642. [Google Scholar] [CrossRef] [PubMed]
  287. Mittler, R. ROS are good. Trend Plant Sci. 2017, 22, 11–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  288. Trachootham, D.; Alexandre, J.; Huang, P. Targeting cancer cells by ROS-mediated mechanisms: A radical therapeutic approach? Nat. Rev. Drug Discov. 2009, 8, 579–591. [Google Scholar] [CrossRef] [PubMed]
  289. Li, Z.; Carrier, L.; Rowan, B.G. Methylseleninic acid synergizes with tamoxifen to induce caspase-mediated apoptosis in breast cancer cells. Mol. Cancer Ther. 2008, 7, 3056–3063. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  290. Wang, H.; Zhang, J.; Yu, H. Elemental selenium at nano size possesses lower toxicity without compromising the fundamental effect on selenoenzymes: Comparison with selenomethionine in mice. Free Radic. Biol. Med. 2007, 42, 1524–1533. [Google Scholar] [CrossRef] [PubMed]
  291. Zhao, G.; Wu, X.; Chen, P.; Zhang, L.; Yang, C.S.; Zhang, J. Selenium nanoparticles are more efficient than sodium selenite in producing reactive oxygen species and hyper-accumulation of selenium nanoparticles in cancer cells generates potent therapeutic effects. Free Radic. Biol. Med. 2018, 126, 55–66. [Google Scholar] [CrossRef]
  292. Kong, Q.; Beel, J.; Lillehei, K. A threshold concept for cancer therapy. Med. Hypotheses 2000, 55, 29–35. [Google Scholar] [CrossRef] [PubMed]
  293. Zhao, G.; Dong, R.; Teng, J.; Yang, L.; Liu, T.; Wu, X.; He, Y.; Wang, Z.; Pu, H.; Wang, Y. N-Acetyl-l-cysteine enhances the effect of selenium nanoparticles on cancer cytotoxicity by increasing the production of selenium-induced reactive oxygen species. ACS Omega 2020, 5, 11710–11720. [Google Scholar] [CrossRef] [PubMed]
  294. Wang, X.; Sun, K.; Tan, Y.; Wu, S.; Zhang, J. Efficacy and safety of selenium nanoparticles administered intraperitoneally for the prevention of growth of cancer cells in the peritoneal cavity. Free Radic. Biol. Med. 2014, 72, 1–10. [Google Scholar] [CrossRef] [PubMed]
  295. Sonkusre, P.; Cameotra, S.S. Biogenic selenium nanoparticles induce ROS-mediated necroptosis in PC-3 cancer cells through TNF activation. J. Nanobiotechnol. 2017, 15, 43. [Google Scholar] [CrossRef] [Green Version]
  296. Wang, Y.; Chen, P.; Zhao, G.; Sun, K.; Li, D.; Wan, X.; Zhang, J. Inverse relationship between elemental selenium nanoparticle size and inhibition of cancer cell growth in vitro and in vivo. Food Chem. Toxicol. 2015, 85, 71–77. [Google Scholar] [CrossRef]
  297. Valdiglesias, V.; Pasaro, E.; Méndez, J.; Laffon, B. In vitro evaluation of selenium genotoxic, cytotoxic, and protective effects: A review. Arch. Toxicol. 2010, 84, 337–351. [Google Scholar] [CrossRef]
  298. Xing, F.; Li, S.; Ge, X.; Wang, C.; Zeng, H.; Li, D.; Dong, L. The inhibitory effect of a novel organoselenium compound BBSKE on the tongue cancer Tca8113 in vitro and in vivo. Oral Oncol. 2008, 44, 963–969. [Google Scholar] [CrossRef]
  299. Shiah, H.-S.; Lee, W.-S.; Juang, S.-H.; Hong, P.-C.; Lung, C.-C.; Chang, C.-J.; Chou, K.-M.; Chang, J.-Y. Mitochondria-mediated and p53-associated apoptosis induced in human cancer cells by a novel selenophene derivative, D-501036. Biochem. Pharmacol. 2007, 73, 610–619. [Google Scholar] [CrossRef]
  300. Shang, D.; Li, Y.; Wang, C.; Wang, X.; Yu, Z.; Fu, X. A novel polysaccharide from Se-enriched Ganoderma lucidum induces apoptosis of human breast cancer cells. Oncol. Rep. 2010, 25, 267–272. [Google Scholar] [CrossRef] [Green Version]
  301. Shang, D.; Zhang, J.; Wen, L.; Li, Y.; Cui, Q. Preparation, characterization, and antiproliferative activities of the Se-containing polysaccharide SeGLP-2B-1 from Se-enriched Ganoderma lucidum. J. Agric. Food Chem. 2009, 57, 7737–7742. [Google Scholar] [CrossRef]
  302. Zamora-Ledezma, C.; Clavijo, C.D.F.; Medina, E.; Sinche, F.; Vispo, N.S.; Dahoumane, S.A.; Alexis, F. Biomedical science to tackle the COVID-19 pandemic: Current status and future perspectives. Molecules 2020, 25, 4620. [Google Scholar] [CrossRef] [PubMed]
  303. Vazquez-Munoz, R.; Lopez-Ribot, J.L. Nanotechnology as an alternative to reduce the spread of COVID-19. Challenges 2020, 11, 15. [Google Scholar] [CrossRef]
  304. Rai, P.K.; Usmani, Z.; Thakur, V.K.; Gupta, V.K.; Mishra, Y.K. Tackling COVID-19 pandemic through nanocoatings: Confront and exactitude. Curr. Res. Green Sustain. Chem. 2020, 3, 100011. [Google Scholar] [CrossRef]
  305. Rao, L.; Xia, S.; Xu, W.; Tian, R.; Yu, G.; Gu, C.; Pan, P.; Meng, Q.-F.; Cai, X.; Qu, D.; et al. Decoy nanoparticles protect against COVID-19 by concurrently adsorbing viruses and inflammatory cytokines. Proc. Natl. Acad. Sci. USA 2020, 117, 27141–27147. [Google Scholar] [CrossRef]
  306. Tharayil, A.; Rajakumari, R.; Kumar, A.; Choudhary, M.D.; Palit, P.; Thomas, S. New insights into application of nanoparticles in the diagnosis and screening of novel coronavirus (SARS-CoV-2). Emergent Mater. 2021, 4, 101–117. [Google Scholar] [CrossRef]
  307. Nasrollahzadeh, M.; Sajjadi, M.; Soufi, G.J.; Iravani, S.; Varma, R.S. Nanomaterials and nanotechnology-associated innovations against viral infections with a focus on coronaviruses. Nanomaterials 2020, 10, 1072. [Google Scholar] [CrossRef]
  308. Zhang, J.-S.; Gao, X.-Y.; Zhang, L.-D.; Bao, Y.-P. Biological effects of a nano red elemental selenium. Biofactors 2001, 15, 27–38. [Google Scholar] [CrossRef]
  309. Zhang, J.; Wang, H.; Yan, X.; Zhang, L. Comparison of short-term toxicity between nano-Se and selenite in mice. Life Sci. 2005, 76, 1099–1109. [Google Scholar] [CrossRef]
  310. Jia, X.; Li, N.; Chen, J. A subchronic toxicity study of elemental nano-Se in Sprague-Dawley rats. Life Sci. 2005, 76, 1989–2003. [Google Scholar] [CrossRef]
  311. Hu, C.; Li, Y.; Xiong, L.; Zhang, H.; Song, J.; Xia, M. Comparative effects of nano elemental selenium and sodium selenite on selenium retention in broiler chickens. Anim. Feed. Sci. Technol. 2012, 177, 204–210. [Google Scholar] [CrossRef]
  312. Benko, I.; Nagy, G.; Tanczos, B.; Ungvari, E.; Sztrik, A.; Eszenyi, P.; Prokisch, J.; Banfalvi, G. Subacute toxicity of nano-selenium compared to other selenium species in mice. Environ. Toxicol. Chem. 2013, 31, 2812–2820. [Google Scholar] [CrossRef] [PubMed]
  313. Shakibaie, M.; Shahverdi, A.R.; Faramarzi, M.A.; Hassanzadeh, G.R.; Rahimi, H.R.; Sabzevari, O. Acute and subacute toxicity of novel biogenic selenium nanoparticles in mice. Pharm. Biol. 2012, 51, 58–63. [Google Scholar] [CrossRef] [PubMed]
  314. Shi, L.; Xun, W.; Yue, W.; Zhang, C.; Ren, Y.; Shi, L.; Wang, Q.; Yang, R.; Lei, F. Effect of sodium selenite, Se-yeast and nano-elemental selenium on growth performance, Se concentration and antioxidant status in growing male goats. Small Rumin. Res. 2011, 96, 49–52. [Google Scholar] [CrossRef]
  315. Liu, T.; Zeng, L.; Jiang, W.; Fu, Y.; Zheng, W.; Chen, T. Rational design of cancer-targeted selenium nanoparticles to antagonize multidrug resistance in cancer cells. Nanomed. Nanotechnol. Biol. Med. 2015, 11, 947–958. [Google Scholar] [CrossRef]
  316. Ramya, S.; Shanmugasundaram, T.; Balagurunathan, R. Biomedical potential of actinobacterially synthesized selenium nanoparticles with special reference to anti-biofilm, anti-oxidant, wound healing, cytotoxic and anti-viral activities. J. Trace Elem. Med. Biol. 2015, 32, 30–39. [Google Scholar] [CrossRef] [PubMed]
  317. Soflaei, S.; Dalimi, A.; Abdoli, A.; Kamali, M.; Nasiri, V.; Shakibaie, M.; Tat, M. Anti-leishmanial activities of selenium nanoparticles and selenium dioxide on Leishmania infantum. Comp. Clin. Pathol. 2012, 23, 15–20. [Google Scholar] [CrossRef]
  318. Li, Y.; He, J.; Shen, X. Effects of nano-selenium poisoning on immune function in the Wumeng semi-fine wool sSheep. Biol. Trace Elem. Res. 2021, 199, 2919–2924. [Google Scholar] [CrossRef]
  319. Li, Y.; Lin, Z.; Guo, M.; Zhao, M.; Xia, Y.; Wang, C.; Xu, T.; Zhu, B. Inhibition of H1N1 influenza virus-induced apoptosis by functionalized selenium nanoparticles with amantadine through ROS-mediated AKT signaling pathways. Int. J. Nanomed. 2018, 13, 2005–2016. [Google Scholar] [CrossRef] [Green Version]
  320. Wang, C.; Chen, H.; Chen, D.; Zhao, M.; Lin, Z.; Guo, M.; Xu, T.; Chen, Y.; Hua, L.; Lin, T.; et al. The inhibition of H1N1 influenza virus-induced apoptosis by surface decoration of selenium nanoparticles with b-thujaplicin through reactive oxygen species-mediated AKT and p53 signaling pathways. ACS Omega 2020, 5, 30633–30642. [Google Scholar] [CrossRef]
  321. Alexander, J.; Tinkov, A.; Strand, T.A.; Alehagen, U.; Skalny, A.; Aaseth, J. Early nutritional interventions with zinc, selenium and vitamin D for raising anti-viral resistance against progressive COVID-19. Nutrients 2020, 12, 2358. [Google Scholar] [CrossRef]
  322. Liu, Q.; Zhao, X.; Ma, J.; Mu, Y.; Wang, Y.; Yang, S.; Wu, Y.; Wu, F.; Zhou, Y. Selenium (Se) plays a key role in the biological effects of some viruses: Implications for COVID-19. Environ. Res. 2021, 196, 110984. [Google Scholar] [CrossRef] [PubMed]
  323. Kieliszek, M.; Lipinski, B. Selenium supplementation in the prevention of coronavirus infections (COVID-19). Med. Hypotheses 2020, 143, 109878. [Google Scholar] [CrossRef] [PubMed]
  324. Bae, M.; Kim, H. Mini-review on the roles of vitamin C, vitamin D, and selenium in the immune system against COVID-19. Molecules 2020, 25, 5346. [Google Scholar] [CrossRef] [PubMed]
  325. Hiffler, L.; Rakotoambinina, B. Selenium and RNA virus interactions: Potential implications for SARS-CoV-2 infection (COVID-19). Front. Nutr. 2020, 7, 164. [Google Scholar] [CrossRef]
  326. Majeed, M.; Nagabhushanam, K.; Gowda, S.; Mundkur, L. An exploratory study of selenium status in healthy individuals and in patients with COVID-19 in a South Indian population: The case for adequate selenium status. Nutrition 2021, 82, 111053. [Google Scholar] [CrossRef]
  327. Dhama, K.; Sharun, K.; Tiwari, R.; Dadar, M.; Malik, Y.S.; Singh, K.P.; Chaicumpa, W. COVID-19, an emerging coronavirus infection: Advances and prospects in designing and developing vaccines, immunotherapeutics, and therapeutics. Hum. Vaccines Immunother. 2020, 16, 1232–1238. [Google Scholar] [CrossRef] [Green Version]
  328. Jin, Z.; Du, X.; Xu, Y.; Deng, Y.; Liu, M.; Zhao, Y.; Zhang, B.; Li, X.; Zhang, L.; Peng, C.; et al. Structure of M(pro) from SARS-CoV-2 and discovery of its inhibitors. Nature 2020, 582, 289–293. [Google Scholar] [CrossRef] [Green Version]
  329. Manzanares, W.; Moreira, E.; Hardy, G. Pharmaconutrition revisited for critically ill patients with coronavirus disease 2019 (COVID-19): Does selenium have a place? Nutrition 2021, 81, 110989. [Google Scholar] [CrossRef]
  330. Moghaddam, A.; Heller, R.A.; Sun, Q.; Seelig, J.; Cherkezov, A.; Seibert, L.; Hackler, J.; Seemann, P.; Diegmann, J.; Pilz, M.; et al. Selenium deficiency is associated with mortality risk from COVID-19. Nutrients 2020, 12, 2098. [Google Scholar] [CrossRef]
  331. Wang, Z.; Zheng, Z.; Hu, H.; Zhou, Q.; Liu, W.; Li, X.; Liu, Z.; Wang, Y.; Ma, Y. A point-of-care selenium nanoparticle-based test for the combined detection of anti-SARS-CoV-2 IgM and IgG in human serum and blood. Lab Chip 2020, 20, 4255–4261. [Google Scholar] [CrossRef]
  332. Peng, T.; Sui, Z.; Huang, Z.; Xie, J.; Wen, K.; Zhang, Y.; Huang, W.; Mi, W.; Peng, K.; Dai, X.; et al. Point-of-care test system for detection of immunoglobulin-G and -M against nucleocapsid protein and spike glycoprotein of SARS-CoV-2. Sens. Actuator B Chem. 2021, 331, 129415. [Google Scholar] [CrossRef] [PubMed]
  333. Bayin, Q.; Huang, L.; Ren, C.; Fu, Y.; Ma, X.; Guo, J. Anti-SARS-CoV-2 IgG and IgM detection with a GMR based LFIA system. Talanta 2021, 227, 122207. [Google Scholar] [CrossRef] [PubMed]
  334. Germain, M.; Caputo, F.; Metcalfe, S.; Tosi, G.; Spring, K.; Åslund, A.K.; Pottier, A.; Schiffelers, R.; Ceccaldi, A.; Schmid, R. Delivering the power of nanomedicine to patients today. J. Control. Release 2020, 326, 164–171. [Google Scholar] [CrossRef] [PubMed]
  335. Wu, P.-H.; Opadele, A.E.; Onodera, Y.; Nam, J.-M. Targeting integrins in cancer nanomedicine: Applications in cancer diagnosis and therapy. Cancers 2019, 11, 1783. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  336. Bajpai, S.; Tiwary, S.K.; Sonker, M.; Joshi, A.; Gupta, V.; Kumar, Y.; Shreyash, N.; Biswas, S. Recent advances in nanoparticle-based cancer treatment: A review. ACS Appl. Nano Mater. 2021, 4, 6441–6470. [Google Scholar] [CrossRef]
  337. Cole, J.T.; Holland, N.B. Multifunctional nanoparticles for use in theranostic applications. Drug Deliv. Transl. Res. 2015, 5, 295–309. [Google Scholar] [CrossRef] [PubMed]
  338. Ryu, J.H.; Lee, S.; Son, S.; Kim, S.H.; Leary, J.F.; Choi, K.; Kwon, I.C. Theranostic nanoparticles for future personalized medicine. J. Control. Release 2014, 190, 477–484. [Google Scholar] [CrossRef]
  339. Xie, Q.; Deng, W.; Yuan, X.; Wang, H.; Ma, Z.; Wu, B.; Zhang, X. Selenium-functionalized liposomes for systemic delivery of doxorubicin with enhanced pharmacokinetics and anticancer effect. Eur. J. Pharm. Biopharm. 2018, 122, 87–95. [Google Scholar] [CrossRef]
  340. Treuel, L.; Eslahian, K.A.; Docter, D.; Lang, T.; Zellner, R.; Nienhaus, K.; Nienhaus, G.U.; Stauber, R.H.; Maskos, M. Physicochemical characterization of nanoparticles and their behavior in the biological environment. Phys. Chem. Chem. Phys. 2014, 16, 15053–15067. [Google Scholar] [CrossRef]
  341. Ernsting, M.J.; Murakami, M.; Roy, A.; Li, S.-D. Factors controlling the pharmacokinetics, biodistribution and intratumoral penetration of nanoparticles. J. Control. Release 2013, 172, 782–794. [Google Scholar] [CrossRef] [Green Version]
  342. Foroozandeh, P.; Aziz, A.A. Merging worlds of nanomaterials and biological environment: Factors governing protein corona formation on nanoparticles and its biological consequences. Nanoscale Res. Lett. 2015, 10, 221. [Google Scholar] [CrossRef] [PubMed]
  343. Gharbavi, M.; Johari, B.; Mousazadeh, N.; Rahimi, B.; Leilan, M.P.; Eslami, S.S.; Sharafi, A. Hybrid of niosomes and bio-synthesized selenium nanoparticles as a novel approach in drug delivery for cancer treatment. Mol. Biol. Rep. 2020, 47, 6517–6529. [Google Scholar] [CrossRef] [PubMed]
  344. Naidoo, S.; Daniels, A.; Habib, S.; Singh, M. Poly-L-lysine–lactobionic acid-capped selenium nanoparticles for liver-targeted gene delivery. Int. J. Mol. Sci. 2022, 23, 1492. [Google Scholar] [CrossRef] [PubMed]
  345. Maiyo, F.; Singh, M. Polymerized selenium nanoparticles for folate-receptor-targeted delivery of anti-luc-siRNA: Potential for gene silencing. Biomedicines 2020, 8, 76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  346. Korany, M.; Marzook, F.; Mahmoud, B.; Ahmed, S.A.; Ayoub, S.M.; Sakr, T.M. Exhibiting the diagnostic face of selenium nanoparticles as a radio-platform for tumor imaging. Bioorg. Chem. 2020, 100, 103910. [Google Scholar] [CrossRef] [PubMed]
  347. Liu, X.; Wang, Y.; Yu, Q.; Deng, G.; Wang, Q.; Ma, X.; Wang, Q.; Lu, J. Selenium nanocomposites as multifunctional nanoplatform for imaging guiding synergistic chemo-photothermal therapy. Colloids Surf. B Biointerfaces 2018, 166, 161–169. [Google Scholar] [CrossRef]
  348. Yu, C.; Wo, F.; Shao, Y.; Dai, X.; Chu, M. Bovine serum albumin nanospheres synchronously encapsulating “gold selenium/gold” nanoparticles and photosensitizer for high-efficiency cancer phototherapy. Appl. Biochem. Biotechnol. 2013, 169, 1566–1578. [Google Scholar] [CrossRef]
  349. Auría-Soro, C.; Nesma, T.; Juanes-Velasco, P.; Landeira-Viñuela, A.; Fidalgo-Gomez, H.; Acebes-Fernandez, V.; Gongora, R.; Almendral Parra, M.J.; Manzano-Roman, R.; Fuentes, M. Interactions of nanoparticles and biosystems: Microenvironment of nanoparticles and biomolecules in nanomedicine. Nanomaterials 2019, 9, 1365. [Google Scholar] [CrossRef] [Green Version]
  350. Kumari, A.; Kumar Yadav, S. Cellular interactions of therapeutically delivred nanoparticules. Expert Opin. Drug Deliv. 2011, 8, 141–151. [Google Scholar] [CrossRef]
  351. Cai, X.; Liu, X.; Jiang, J.; Gao, M.; Wang, W.; Zheng, H.; Xu, S.; Li, R. Molecular mechanisms, characterization methods, and utilities of nanoparticle biotransformation in nanosafety assessments. Small 2020, 16, e1907663. [Google Scholar] [CrossRef]
  352. Chiappi, M.; Conesa, J.J.; Pereiro, E.; Sorzano, C.O.S.; Rodríguez, M.J.; Henzler, K.; Schneider, G.; Chichón, F.J.; Carrascosa, J.L. Cryo-soft X-ray tomography as a quantitative three-dimensional tool to model nanoparticle:cell interaction. J. Nanobiotechnol. 2016, 14, 15. [Google Scholar] [CrossRef] [PubMed]
  353. Ivask, A.; Mitchell, A.J.; Malysheva, A.; Voelcker, N.H.; Lombi, E. Methodologies and approaches for the analysis of cell–nanoparticle interactions. WIREs Nanomed. Nanobiotechnol. 2018, 10, e1486. [Google Scholar] [CrossRef] [PubMed]
  354. Kamaly, N.; Xiao, Z.; Valencia, P.M.; Radovic-Moreno, A.F.; Farokhzad, O.C. Targeted polymeric therapeutic nanoparticles: Design, development and clinical translation. Chem. Soc. Rev. 2012, 41, 2971–3010. [Google Scholar] [CrossRef] [PubMed]
  355. Liu, Y.; Tan, J.; Thomas, A.; Ou-Yang, D.; Muzykantov, V.R. The shape of things to come: Importance of design in nanotechnology for drug delivery. Ther. Deliv. 2012, 3, 181–194. [Google Scholar] [CrossRef] [Green Version]
  356. Verma, A.; Stellacci, F. Effect of surface properties on nanoparticle-cell interactions. Small 2010, 6, 12–21. [Google Scholar] [CrossRef]
  357. Toy, R.; Bauer, L.; Hoimes, C.; Ghaghada, K.B.; Karathanasis, E. Targeted nanotechnology for cancer imaging. Adv. Drug Deliv. Rev. 2014, 76, 79–97. [Google Scholar] [CrossRef] [Green Version]
  358. Fleischer, C.C.; Payne, C.K. Nanoparticle–cell interactions: Molecular structure of the protein corona and cellular outcomes. Acc. Chem. Res. 2014, 47, 2651–2659. [Google Scholar] [CrossRef]
  359. Gräfe, C.; Weidner, A.; Lühe, M.V.; Bergemann, C.; Schacher, F.H.; Clement, J.H.; Dutz, S. Intentional formation of a protein corona on nanoparticles: Serum concentration affects protein corona mass, surface charge, and nanoparticle–cell interaction. Int. J. Biochem. Cell Biol. 2016, 75, 196–202. [Google Scholar] [CrossRef]
  360. Runa, S.; Hussey, M.; Payne, C.K. Nanoparticle–cell interactions: Relevance for public health. J. Phys. Chem. B 2017, 122, 1009–1016. [Google Scholar] [CrossRef] [Green Version]
  361. Ivask, A.; Mitchell, A.J.; Hope, C.M.; Barry, S.C.; Lombi, E.; Voelcker, N.H. Single cell level quantification of nanoparticle–cell interactions using mass cytometry. Anal. Chem. 2017, 89, 8228–8232. [Google Scholar] [CrossRef]
  362. Rasmussen, L.; Shi, H.; Liu, W.; Shannon, K.B. Quantification of silver nanoparticle interactions with yeast Saccharomyces cerevisiae studied using single-cell ICP-MS. Anal. Bioanal. Chem. 2022, 414, 3077–3086. [Google Scholar] [CrossRef] [PubMed]
  363. Shim, Y.J.; Ha, M.K.; Yoon, T.H. Quantitative estimation of cell-associated silver nanoparticles using the normalized side scattering intensities of flow cytometry. Nanomaterials 2021, 11, 3079. [Google Scholar] [CrossRef] [PubMed]
  364. Korzeniowska, B.; Fonseca, M.P.; Gorshkov, V.; Skytte, L.; Rasmussen, K.L.; Schrøder, H.D.; Kjeldsen, F. The cytotoxicity of metal nanoparticles depends on their synergistic interactions. Part. Part. Syst. Charact. 2020, 37, 2000135. [Google Scholar] [CrossRef]
  365. Khanehzar, A.; Fraire, J.C.; Xi, M.; Feizpour, A.; Xu, F.; Wu, L.; Coronado, E.A.; Reinhard, B.M. Nanoparticle–cell interactions induced apoptosis: A case study with nanoconjugated epidermal growth factor. Nanoscale 2018, 10, 6712–6723. [Google Scholar] [CrossRef] [PubMed]
  366. Xie, X.; Hu, X.; Li, Q.; Yin, M.; Song, H.; Hu, J.; Wang, L.; Fan, C.; Chen, N. Unraveling cell-type-specific targeted delivery of membrane-camouflaged nanoparticles with plasmonic imaging. Nano Lett. 2020, 20, 5228–5235. [Google Scholar] [CrossRef]
  367. Mutalik, S.P.; Pandey, A.; Mutalik, S. Nanoarchitectronics: A versatile tool for deciphering nanoparticle interaction with cellular proteins, nucleic acids and phospholipids at biological interfaces. Int. J. Biol. Macromol. 2020, 151, 136–158. [Google Scholar] [CrossRef]
  368. Nash, J.A.; Kwansa, A.L.; Peerless, J.S.; Kim, H.S.; Yingling, Y.G. Advances in molecular modeling of nanoparticle–nucleic acid interfaces. Bioconjug. Chem. 2017, 28, 3–10. [Google Scholar] [CrossRef]
  369. Samanta, A.; Medintz, I.L. Nanoparticles and DNA—A powerful and growing functional combination in bionanotechnology. Nanoscale 2016, 8, 9037–9095. [Google Scholar] [CrossRef] [Green Version]
  370. Arsalan, A.; Younus, H. Enzymes and nanoparticles: Modulation of enzymatic activity via nanoparticles. Int. J. Biol. Macromol. 2018, 118, 1833–1847. [Google Scholar] [CrossRef]
  371. Fusco, L.; Avitabile, E.; Armuzza, V.; Orecchioni, M.; Istif, A.; Bedognetti, D.; Da Ros, T.; Delogu, L.G. Impact of the surface functionalization on nanodiamond biocompatibility: A comprehensive view on human blood immune cells. Carbon 2020, 160, 390–404. [Google Scholar] [CrossRef]
  372. Suk, J.S.; Xu, Q.; Kim, N.; Hanes, J.; Ensign, L.M. PEGylation as a strategy for improving nanoparticle-based drug and gene delivery. Adv. Drug Deliv. Rev. 2016, 99, 28–51. [Google Scholar] [CrossRef] [PubMed]
  373. Guerrini, L.; Alvarez-Puebla, R.A.; Pazos-Perez, N. Surface modifications of nanoparticles for stability in biological fluids. Materials 2018, 11, 1154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  374. Yan, C.; Wang, T. A new view for nanoparticle assemblies: From crystalline to binary cooperative complementarity. Chem. Soc. Rev. 2017, 46, 1483–1509. [Google Scholar] [CrossRef] [PubMed]
  375. Dyawanapelly, S.; Mehrotra, P.; Ghosh, G.; Jagtap, D.D.; Dandekar, P.; Jain, R. How the surface functionalized nanoparticles affect conformation and activity of proteins: Exploring through protein-nanoparticle interactions. Bioorg. Chem. 2019, 82, 17–25. [Google Scholar] [CrossRef]
  376. Kolosnjaj-Tabi, J.; Lartigue, L.; Javed, Y.; Luciani, N.; Pellegrino, T.; Wilhelm, C.; Alloyeau, D.; Gazeau, F. Biotransformations of magnetic nanoparticles in the body. Nano Today 2016, 11, 280–284. [Google Scholar] [CrossRef]
  377. Ruiz, A.; Hernández, Y.; Cabal, C.; González, E.; Veintemillas-Verdaguer, S.; Martínez, E.; Morales, M.P. Biodistribution and pharmacokinetics of uniform magnetite nanoparticles chemically modified with polyethylene glycol. Nanoscale 2013, 5, 11400–11408. [Google Scholar] [CrossRef] [Green Version]
  378. Vlasova, I.I.; Kapralov, A.A.; Michael, Z.P.; Burkert, S.C.; Shurin, M.R.; Star, A.; Shvedova, A.A.; Kagan, V.E. Enzymatic oxidative biodegradation of nanoparticles: Mechanisms, significance and applications. Toxicol. Appl. Pharmacol. 2016, 299, 58–69. [Google Scholar] [CrossRef] [Green Version]
  379. Yang, G.; Phua, S.Z.F.; Bindra, A.K.; Zhao, Y. Degradability and clearance of inorganic nanoparticles for biomedical applications. Adv. Mater. 2019, 31, e1805730. [Google Scholar] [CrossRef]
  380. de Dios, A.S.; Diaz-Garcia, M.E. Multifunctional nanoparticles: Analytical prospects. Anal. Chim. Acta 2010, 666, 1–22. [Google Scholar] [CrossRef]
  381. Murar, M.; Albertazzi, L.; Pujals, S. Advanced optical imaging-guided nanotheranostics towards personalized cancer drug delivery. Nanomaterials 2022, 12, 399. [Google Scholar] [CrossRef]
  382. Russell, P.; Hagemeyer, C.E.; Esser, L.; Voelcker, N.H. Theranostic nanoparticles for the management of thrombosis. Theranostics 2022, 12, 2773–2800. [Google Scholar] [CrossRef] [PubMed]
  383. Beh, C.Y.; Prajnamitra, R.P.; Chen, L.-L.; Hsieh, P.C.-H. Advances in biomimetic nanoparticles for targeted cancer therapy and diagnosis. Molecules 2021, 26, 5052. [Google Scholar] [CrossRef] [PubMed]
  384. Chen, Q.; Yu, Q.; Liu, Y.; Bhavsar, D.; Yang, L.; Ren, X.; Sun, D.; Zheng, W.; Liu, J.; Chen, L.M. Multifunctional selenium nanoparticles: Chiral selectivity of delivering MDR-siRNA for reversal of multidrug resistance and real-time biofluorescence imaging. Nanomedicine 2015, 11, 1773–1784. [Google Scholar] [CrossRef] [PubMed]
  385. Zheng, W.; Cao, C.; Liu, Y.; Yu, Q.; Zheng, C.; Sun, D.; Ren, X.; Liu, J. Multifunctional polyamidoamine-modified selenium nanoparticles dual-delivering siRNA and cisplatin to A549/DDP cells for reversal multidrug resistance. Acta Biomater. 2015, 11, 368–380. [Google Scholar] [CrossRef]
  386. Zhu, B.; Lin, Z.; Zhao, M.; Xu, T.; Wang, C.; Xia, H.; Wang, H.; Li, Y. Multifunctional selenium nanoparticles as carriers of HSP70 siRNA to induce apoptosis of HepG2 cells. Int. J. Nanomed. 2016, 11, 3065–3076. [Google Scholar] [CrossRef] [Green Version]
  387. Siemer, S.; Westmeier, D.; Barz, M.; Eckrich, J.; Wünsch, D.; Seckert, C.; Thyssen, C.; Schilling, O.; Hasenberg, M.; Pang, C.; et al. Biomolecule-corona formation confers resistance of bacteria to nanoparticle-induced killing: Implications for the design of improved nanoantibiotics. Biomaterials 2019, 192, 551–559. [Google Scholar] [CrossRef]
  388. Wang, Y.; Li, M.; Xu, X.; Tang, W.; Xiong, L.; Sun, Q. Formation of protein corona on nanoparticles with digestive enzymes in simulated gastrointestinal fluids. J. Agric. Food Chem. 2019, 67, 2296–2306. [Google Scholar] [CrossRef]
  389. Peng, Q.; Liu, J.; Zhang, T.; Zhang, T.-X.; Zhang, C.-L.; Mu, H. Digestive enzyme corona formed in the gastrointestinal tract and its impact on epithelial cell uptake of nanoparticles. Biomacromolecules 2019, 20, 1789–1797. [Google Scholar] [CrossRef]
  390. Obst, K.; Yealland, G.; Balzus, B.; Miceli, E.; Dimde, M.; Weise, C.; Eravci, M.; Bodmeier, R.; Haag, R.; Calderón, M.; et al. Protein corona formation on colloidal polymeric nanoparticles and polymeric nanogels: Impact on cellular uptake, toxicity, immunogenicity, and drug release properties. Biomacromolecules 2017, 18, 1762–1771. [Google Scholar] [CrossRef]
  391. Gupta, M.N.; Roy, I. How corona formation impacts nanomaterials as drug carriers. Mol. Pharm. 2020, 17, 725–737. [Google Scholar] [CrossRef]
  392. Pyrgiotakis, G.; Blattmann, C.O.; Demokritou, P. Real-time nanoparticle–cell interactions in physiological media by atomic force microscopy. ACS Sustain. Chem. Eng. 2014, 2, 1681–1690. [Google Scholar] [CrossRef] [PubMed]
  393. Corbo, C.; Molinaro, R.; Tabatabaei, M.; Farokhzad, O.C.; Mahmoudi, M. Personalized protein corona on nanoparticles and its clinical implications. Biomater. Sci. 2017, 5, 378–387. [Google Scholar] [CrossRef] [PubMed]
  394. Chakraborty, D.; Chauhan, P.; Alex, S.A.; Chaudhary, S.; Ethiraj, K.; Chandrasekaran, N.; Mukherjee, A. Comprehensive study on biocorona formation on functionalized selenium nanoparticle and its biological implications. J. Mol. Liq. 2018, 268, 335–342. [Google Scholar] [CrossRef]
  395. Chakraborty, D.; Chauhan, P.; Kumar, S.; Chaudhary, S.; Chandrasekaran, N.; Mukherjee, A.; Ethiraj, K. Utilizing corona on functionalized selenium nanoparticles for loading and release of doxorubicin payload. J. Mol. Liq. 2019, 296, 111864. [Google Scholar] [CrossRef]
  396. Borowska, M.; Pawlik, E.; Jankowski, K. Investigation of interaction between biogenic selenium nanoparticles and human serum albumin using microwave plasma optical emission spectrometry operating in a single-particle mode. Monatsh. Chem. 2020, 151, 1283–1290. [Google Scholar] [CrossRef]
  397. Swierczewska, M.; Han, H.; Kim, K.; Park, J.; Lee, S. Polysaccharide-based nanoparticles for theranostic nanomedicine. Adv. Drug Deliv. Rev. 2016, 99, 70–84. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  398. Andreani, T.; Fangueiro, J.F.; Severino, P.; de Souza, A.L.R.; Martins-Gomes, C.; Fernandes, P.M.V.; Calpena, A.C.; Gremião, M.P.; Souto, E.B.; Silva, A.M. The Influence of polysaccharide coating on the physicochemical parameters and cytotoxicity of silica nanoparticles for hydrophilic biomolecules delivery. Nanomaterials 2019, 9, 1081. [Google Scholar] [CrossRef] [Green Version]
  399. Lemarchand, C.; Gref, R.; Couvreur, P. Polysaccharide-decorated nanoparticles. Eur. J. Pharm. Biopharm. 2004, 58, 327–341. [Google Scholar] [CrossRef]
  400. Salatin, S.; Khosroushahi, A.Y. Overviews on the cellular uptake mechanism of polysaccharide colloidal nanoparticles. J. Cell Mol. Med. 2017, 21, 1668–1686. [Google Scholar] [CrossRef] [Green Version]
  401. Liu, Z.; Jiao, Y.; Wang, Y.; Zhou, C.; Zhang, Z. Polysaccharides-based nanoparticles as drug delivery systems. Adv. Drug Deliv. Rev. 2008, 60, 1650–1662. [Google Scholar] [CrossRef]
  402. Pugh, N.; Ross, S.; ElSohly, H.; ElSohly, M.; Pasco, D.S. Isolation of three high molecular weight polysaccharide preparations with potent immunostimulatory activity from Spirulina platensis, Aphanizomenon flos-aquae and Chlorella pyrenoidosa. Planta Med. 2001, 67, 737–742. [Google Scholar] [CrossRef] [PubMed]
  403. Yang, F.; Tang, Q.; Zhong, X.; Bai, Y.; Chen, T.; Zhang, Y.; Li, Y.; Zheng, W. Surface decoration by Spirulina polysaccharide enhances the cellular uptake and anticancer efficacy of selenium nanoparticles. Int. J. Nanomed. 2012, 7, 835–844. [Google Scholar] [CrossRef] [Green Version]
  404. Li, J.; Shen, B.; Nie, S.; Duan, Z.; Chen, K. A combination of selenium and polysaccharides: Promising therapeutic potential. Carbohydr. Polym. 2019, 206, 163–173. [Google Scholar] [CrossRef] [PubMed]
  405. Arner, E.S. Selenoproteins-What unique properties can arise with selenocysteine in place of cysteine? Exp. Cell Res. 2010, 316, 1296–1303. [Google Scholar] [CrossRef] [PubMed]
  406. Papp, L.V.; Lu, J.; Holmgren, A.; Khanna, K.K. From selenium to selenoproteins: Synthesis, identity, and their role in human health. Antioxid. Redox Signal. 2007, 9, 775–806. [Google Scholar] [CrossRef] [Green Version]
  407. Zoidis, E.; Seremelis, I.; Kontopoulos, N.; Danezis, G.P. Selenium-dependent antioxidant enzymes: Actions and properties of selenoproteins. Antioxidants 2018, 7, 66. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  408. Avery, J.C.; Hoffmann, P.R. Selenium, selenoproteins, and immunity. Nutrients 2018, 10, 1203. [Google Scholar] [CrossRef] [Green Version]
  409. Labunskyy, V.; Lee, B.C.; Handy, D.; Loscalzo, J.; Hatfield, D.L.; Gladyshev, V.N. Both maximal expression of selenoproteins and selenoprotein deficiency can promote development of type 2 diabetes-like phenotype in mice. Antioxid. Redox Signal. 2011, 14, 2327–2336. [Google Scholar] [CrossRef]
  410. Shchedrina, V.A.; Zhang, Y.; Labunskyy, V.; Hatfield, D.L.; Gladyshev, V.N. Structure–function relations, physiological roles, and evolution of mammalian ER-resident selenoproteins. Antioxid. Redox Signal. 2010, 12, 839–849. [Google Scholar] [CrossRef]
  411. Arbogast, S.; Ferreiro, A. Selenoproteins and protection against oxidative stress: Selenoprotein N as a novel player at the crossroads of redox signaling and calcium homeostasis. Antioxid. Redox Signal. 2010, 12, 893–904. [Google Scholar] [CrossRef]
  412. Shimada, B.K.; Alfulaij, N.; Seale, L.A. The impact of selenium deficiency on cardiovascular function. Int. J. Mol. Sci. 2021, 22, 10713. [Google Scholar] [CrossRef] [PubMed]
  413. Rocca, C.; Boukhzar, L.; Granieri, M.C.; Alsharif, I.; Mazza, R.; Lefranc, B.; Tota, B.; Leprince, J.; Cerra, M.C.; Anouar, Y.; et al. A selenoprotein T-derived peptide protects the heart against ischaemia/reperfusion injury through inhibition of apoptosis and oxidative stress. Acta Physiol. 2018, 223, e13067. [Google Scholar] [CrossRef] [PubMed]
  414. Schomburg, L. Selenium, selenoproteins and the thyroid gland: Interactions in health and disease. Nat. Rev. Endocrinol. 2011, 8, 160–171. [Google Scholar] [CrossRef] [PubMed]
  415. Valea, A.; Georgescu, C.E. Selenoproteins in human body: Focus on thyroid pathophysiology. Hormones 2018, 17, 183–196. [Google Scholar] [CrossRef] [PubMed]
  416. Gorini, F.; Vassalle, C. Selenium and selenoproteins at the intersection of type 2 diabetes and thyroid pathophysiology. Antioxidants 2022, 11, 1188. [Google Scholar] [CrossRef]
  417. Kang, D.; Lee, J.; Wu, C.; Guo, X.; Lee, B.J.; Chun, J.-S.; Kim, J.-H. The role of selenium metabolism and selenoproteins in cartilage homeostasis and arthropathies. Exp. Mol. Med. 2020, 52, 1198–1208. [Google Scholar] [CrossRef]
  418. Gong, T.; Torres, D.; Berry, M.J.; Pitts, M.W. Hypothalamic redox balance and leptin signaling—Emerging role of selenoproteins. Free Radic. Biol. Med. 2018, 127, 172–181. [Google Scholar] [CrossRef]
  419. Qazi, I.H.; Angel, C.; Yang, H.; Zoidis, E.; Pan, B.; Wu, Z.; Ming, Z.; Zeng, C.-J.; Meng, Q.; Han, H.; et al. Role of selenium and selenoproteins in male reproductive function: A review of past and present evidences. Antioxidants 2019, 8, 268. [Google Scholar] [CrossRef] [Green Version]
  420. Lin, J.; Sahakian, D.C.; De Morais, S.M.F.; Xu, J.J.; Polzer, R.J.; Winter, S.M. The role of absorption, distribution, metabolism, excretion and toxicity in drug discovery. Curr. Top. Med. Chem. 2003, 3, 1125–1154. [Google Scholar] [CrossRef]
  421. Kazi, M.; Alhajri, A.; AlShehri, S.M.; Elzayat, E.M.; Al Meanazel, O.T.; Shakeel, F.; Noman, O.; Altamimi, M.A.; Alanazi, F.K. Enhancing oral bioavailability of apigenin using a bioactive self-nanoemulsifying drug delivery system (Bio-SNEDDS): In vitro, in vivo and stability evaluations. Pharmaceutics 2020, 12, 749. [Google Scholar] [CrossRef]
  422. Al-Kassas, R.; Bansal, M.; Shaw, J. Nanosizing techniques for improving bioavailability of drugs. J. Control. Release 2017, 260, 202–212. [Google Scholar] [CrossRef] [PubMed]
  423. Byun, J.H.; Han, D.-G.; Cho, H.-J.; Yoon, I.-S.; Jung, I.H. Recent advances in physiologically based pharmacokinetic and pharmacodynamic models for anticancer nanomedicines. Arch. Pharmacal Res. 2020, 43, 80–99. [Google Scholar] [CrossRef] [PubMed]
  424. Rafiei, P.; Haddadi, A. Docetaxel-loaded PLGA and PLGA-PEG nanoparticles for intravenous application: Pharmacokinetics and biodistribution profile. Int. J. Nanomed. 2017, 12, 935–947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  425. Kiio, T.M.; Park, S. Physical properties of nanoparticles do matter. J. Pharm. Investig. 2020, 51, 35–51. [Google Scholar] [CrossRef]
  426. Singh, R.; Lillard, J.W., Jr. Nanoparticle-based targeted drug delivery. Exp. Mol. Pathol. 2009, 86, 215–223. [Google Scholar] [CrossRef] [Green Version]
  427. Kadam, R.S.; Bourne, D.W.A.; Kompella, U.B. Nano-advantage in enhanced drug delivery with biodegradable nanoparticles: Contribution of reduced clearance. Drug Metab. Dispos. 2012, 40, 1380–1388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  428. Choi, H.S.; Liu, W.; Liu, F.; Nasr, K.; Misra, P.; Bawendi, M.G.; Frangioni, J.V. Design considerations for tumour-targeted nanoparticles. Nat. Nanotechnol. 2010, 5, 42–47. [Google Scholar] [CrossRef] [Green Version]
  429. Yu, M.; Zheng, J. Clearance pathways and tumor targeting of imaging nanoparticles. ACS Nano 2015, 9, 6655–6674. [Google Scholar] [CrossRef] [Green Version]
  430. Ulbrich, K.; Holá, K.; Šubr, V.; Bakandritsos, A.; Tuček, J.; Zbořil, R. Targeted drug delivery with polymers and magnetic nanoparticles: Covalent and noncovalent approaches, release control, and clinical studies. Chem. Rev. 2016, 116, 5338–5431. [Google Scholar] [CrossRef] [Green Version]
  431. Raj, S.; Khurana, S.; Choudhari, R.; Kesari, K.K.; Kamal, M.A.; Garg, N.; Ruokolainen, J.; Das, B.C.; Kumar, D. Specific targeting cancer cells with nanoparticles and drug delivery in cancer therapy. Semin. Cancer Biol. 2021, 69, 166–177. [Google Scholar] [CrossRef]
  432. Li, Y.; Kröger, M.; Liu, W.K. Shape effect in cellular uptake of PEGylated nanoparticles: Comparison between sphere, rod, cube and disk. Nanoscale 2015, 7, 16631–16646. [Google Scholar] [CrossRef] [PubMed]
  433. Howard, F.; Muthana, M. Designer nanocarriers for navigating the systemic delivery of oncolytic viruses. Nanomedicine 2020, 15, 93–110. [Google Scholar] [CrossRef] [PubMed]
  434. Obrezanova, O.; Martinsson, A.; Whitehead, T.; Mahmoud, S.; Bender, A.; Miljković, F.; Grabowski, P.; Irwin, B.; Oprisiu, I.; Conduit, G.; et al. Prediction of in vivo pharmacokinetic parameters and time–exposure curves in rats using machine learning from the chemical structure. Mol. Pharm. 2022, 19, 1488–1504. [Google Scholar] [CrossRef] [PubMed]
  435. Loeschner, K.; Hadrup, N.; Hansen, M.; Pereira, S.A.; Gammelgaard, B.; Møller, L.H.; Mortensen, A.; Lam, H.R.; Larsen, E.H. Absorption, distribution, metabolism and excretion of selenium following oral administration of elemental selenium nanoparticles or selenite in rats. Metallomics 2014, 6, 330–337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  436. Yang, H.; Jia, X. Safety evaluation of Se-methylselenocysteine as nutritional selenium supplement: Acute toxicity, genotoxicity and subchronic toxicity. Regul. Toxicol. Pharmacol. 2014, 70, 720–727. [Google Scholar] [CrossRef]
  437. Zheng, S.; Xing, H.; Zhang, Q.; Xue, H.; Zhu, F.; Xu, S. Pharmacokinetics of sodium selenite administered orally in blood and tissues of selenium-deficient ducklings. Biol. Trace Elem. Res. 2019, 190, 509–516. [Google Scholar] [CrossRef]
  438. Zhang, S.-Q.; Shen, S.; Zhang, Y. Comparison of bioavailability, pharmacokinetics, and biotransformation of selenium-enriched yeast and sodium selenite in rats using plasma selenium and selenomethionine. Biol. Trace Elem. Res. 2020, 196, 512–516. [Google Scholar] [CrossRef]
  439. Xing, H.; Zheng, S.; Zhang, Z.; Zhu, F.; Xue, H.; Xu, S. Pharmacokinetics of selenium in healthy piglets after different routes of administration: Application of pharmacokinetic data to the risk assessment of selenium. Biol. Trace Elem. Res. 2019, 191, 403–411. [Google Scholar] [CrossRef]
  440. Brodin, O.; Eksborg, S.; Wallenberg, M.; Asker-Hagelberg, C.; Larsen, E.H.; Mohlkert, D.; Lenneby-Helleday, C.; Jacobsson, H.; Linder, S.; Misra, S.; et al. Pharmacokinetics and toxicity of sodium selenite in the treatment of patients with carcinoma in a phase I clinical trial: The SECAR study. Nutrients 2015, 7, 4978–4994. [Google Scholar] [CrossRef] [Green Version]
  441. Manzanares, W.; Biestro, A.; Galusso, F.; Torre, M.H.; Mañáy, N.; Facchin, G.; Hardy, G. High-dose selenium for critically ill patients with systemic inflammation: Pharmacokinetics and pharmacodynamics of selenious acid: A pilot study. Nutrition 2010, 26, 634–640. [Google Scholar] [CrossRef]
  442. Evans, S.O.; Jacobson, G.M.; Goodman, H.J.B.; Bird, S.; Jameson, M.B. Comparative safety and pharmacokinetic evaluation of three oral selenium compounds in cancer patients. Biol. Trace Elem. Res. 2019, 189, 395–404. [Google Scholar] [CrossRef] [PubMed]
  443. Ferro, C.; Florindo, H.F.; Santos, H.A. Selenium nanoparticles for biomedical applications: From development and characterization to therapeutics. Adv. Health Mater. 2021, 10, e2100598. [Google Scholar] [CrossRef] [PubMed]
  444. Gao, F.; Yuan, Q.; Gao, L.; Cai, P.; Zhu, H.; Liu, R.; Wang, Y.; Wei, Y.; Huang, G.; Liang, J.; et al. Cytotoxicity and therapeutic effect of irinotecan combined with selenium nanoparticles. Biomaterials 2014, 35, 8854–8866. [Google Scholar] [CrossRef] [PubMed]
  445. Hong, A.; Rao, L.; Zhuang, M.; Luo, T.; Wang, Y.; Ma, Y. Chitosan-decorated selenium nanoparticles as protein carriers to improve the in vivo half-life of the peptide therapeutic BAY 55-9837 for type 2 diabetes mellitus. Int. J. Nanomed. 2014, 9, 4819–4828. [Google Scholar] [CrossRef] [Green Version]
  446. Deng, W.; Xie, Q.; Wang, H.; Ma, Z.; Wu, B.; Zhang, X. Selenium nanoparticles as versatile carriers for oral delivery of insulin: Insight into the synergic antidiabetic effect and mechanism. Nanomed. Nanotechnol. Biol. Med. 2017, 13, 1965–1974. [Google Scholar] [CrossRef]
  447. Fu, X.; Yang, Y.; Li, X.; Lai, H.; Huang, Y.; He, L.; Zheng, W.; Chen, T. RGD peptide-conjugated selenium nanoparticles: Antiangiogenesis by suppressing VEGF-VEGFR2-ERK/AKT pathway. Nanomed. Nanotechnol. Biol. Med. 2016, 12, 1627–1639. [Google Scholar] [CrossRef]
  448. Rojekar, S.; Pai, R.; Abadi, L.F.; Mahajan, K.; Prajapati, M.K.; Kulkarni, S.; Vavia, P. Dual loaded nanostructured lipid carrier of nano-selenium and Etravirine as a potential anti-HIV therapy. Int. J. Pharm. 2021, 607, 120986. [Google Scholar] [CrossRef]
  449. Xiao, J.; Yan, M.; Zhou, K.; Chen, H.; Xu, Z.; Gan, Y.; Hong, B.; Tian, G.; Qian, J.; Zhang, G.; et al. A nanoselenium-coating biomimetic cytomembrane nanoplatform for mitochondrial targeted chemotherapy- and chemodynamic therapy through manganese and doxorubicin codelivery. J. Nanobiotechnol. 2021, 19, 227. [Google Scholar] [CrossRef]
  450. Yin, J.; Hou, Y.; Yin, Y.; Song, X. Selenium-coated nanostructured lipid carriers used for oral delivery of berberine to accomplish a synergic hypoglycemic effect. Int. J. Nanomed. 2017, 12, 8671–8680. [Google Scholar] [CrossRef] [Green Version]
  451. Mostafavi, E.; Medina-Cruz, D.; Vernet-Crua, A.; Chen, J.; Cholula-Diaz, J.L.; Guisbiers, G.; Webster, T.J. Green nanomedicine: The path to the next generation of nanomaterials for diagnosing brain tumors and therapeutics? Expert Opin. Drug Deliv. 2021, 18, 715–736. [Google Scholar] [CrossRef]
  452. Nath, D.; Banerjee, P. Green nanotechnology—A new hope for medical biology. Environ. Toxicol. Pharmacol. 2013, 36, 997–1014. [Google Scholar] [CrossRef] [PubMed]
  453. Foulkes, R.; Man, E.; Thind, J.; Yeung, S.; Joy, A.; Hoskins, C. The regulation of nanomaterials and nanomedicines for clinical application: Current and future perspectives. Biomater. Sci. 2020, 8, 4653–4664. [Google Scholar] [CrossRef] [PubMed]
  454. Hua, S.; de Matos, M.B.C.; Metselaar, J.M.; Storm, G. Current Trends and Challenges in the Clinical Translation of Nanoparticulate Nanomedicines: Pathways for Translational Development and Commercialization. Front. Pharmacol. 2018, 9, 790. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  455. Bondarenko, O.; Mortimer, M.; Kahru, A.; Feliu, N.; Javed, I.; Kakinen, A.; Lin, S.; Xia, T.; Song, Y.; Davis, T.P.; et al. Nanotoxicology and nanomedicine: The Yin and Yang of nano-bio interactions for the new decade. Nano Today 2021, 39, 101184. [Google Scholar] [CrossRef]
  456. Madani, M.; Hosny, S.; Alshangiti, D.M.; Nady, N.; Alkhursani, S.A.; Alkhaldi, H.; Al-Gahtany, S.A.; Ghobashy, M.M.; Gaber, G.A. Green synthesis of nanoparticles for varied applications: Green renewable resources and energy-efficient synthetic routes. Nanotechnol. Rev. 2022, 11, 731–759. [Google Scholar] [CrossRef]
  457. Gottardo, S.; Mech, A.; Drbohlavová, J.; Małyska, A.; Bøwadt, S.; Sintes, J.R.; Rauscher, H. Towards safe and sustainable innovation in nanotechnology: State-of-play for smart nanomaterials. Nanoimpact 2021, 21, 100297. [Google Scholar] [CrossRef]
  458. Hano, C.; Abbasi, B.H. Plant-based green synthesis of nanoparticles: Production, characterization and applications. Biomolecules 2021, 12, 31. [Google Scholar] [CrossRef]
  459. Dahoumane, S.A.; Mechouet, M.; Alvarez, F.J.; Agathos, S.N.; Jeffryes, C. Microalgae: An outstanding tool in nanotechnology. Bionatura 2016, 1, 196–201. [Google Scholar] [CrossRef] [Green Version]
  460. Rahman, A.; Lin, J.; Jaramillo, F.E.; Bazylinski, D.A.; Jeffryes, C.; Dahoumane, S.A. In vivo biosynthesis of inorganic nanomaterials using eukaryotes—A review. Molecules 2020, 25, 3246. [Google Scholar] [CrossRef]
  461. Dahoumane, S.A.; Wujcik, E.K.; Jeffryes, C. Noble metal, oxide and chalcogenide-based nanomaterials from scalable phototrophic culture systems. Enzym. Microb. Technol. 2016, 95, 13–27. [Google Scholar] [CrossRef]
  462. Varma, R.S. Greener approach to nanomaterials and their sustainable applications. Curr. Opin. Chem. Eng. 2012, 1, 123–128. [Google Scholar] [CrossRef]
  463. Akhtar, M.S.; Panwar, J.; Yun, Y.-S. Biogenic synthesis of metallic nanoparticles by plant extracts. ACS Sustain. Chem. Eng. 2013, 1, 591–602. [Google Scholar] [CrossRef]
  464. Saratale, R.G.; Karuppusamy, I.; Saratale, G.D.; Pugazhendhi, A.; Kumar, G.; Park, Y.; Ghodake, G.S.; Bharagava, R.N.; Banu, J.R.; Shin, H.S. A comprehensive review on green nanomaterials using biological systems: Recent perception and their future applications. Colloids Surf. B Biointerfaces 2018, 170, 20–35. [Google Scholar] [CrossRef] [PubMed]
  465. Jahangirian, H.; Lemraski, E.G.; Webster, T.J.; Rafiee-Moghaddam, R.; Abdollahi, Y. A review of drug delivery systems based on nanotechnology and green chemistry: Green nanomedicine. Int. J. Nanomed. 2017, 12, 2957–2978. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  466. Kanwar, R.; Rathee, J.; Salunke, D.B.; Mehta, S.K. Green nanotechnology-driven drug delivery assemblies. ACS Omega 2019, 4, 8804–8815. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  467. Jahangirian, H.; Lemraski, E.G.; Rafiee-Moghaddam, R.; Webster, T.J. A review of using green chemistry methods for biomaterials in tissue engineering. Int. J. Nanomed. 2018, 13, 5953–5969. [Google Scholar] [CrossRef] [Green Version]
  468. Ovais, M.; Khalil, A.T.; Ayaz, M.; Ahmad, I.; Nethi, S.K.; Mukherjee, S. Biosynthesis of metal nanoparticles via microbial enzymes: A mechanistic approach. Int. J. Mol. Sci. 2018, 19, 4100. [Google Scholar] [CrossRef] [Green Version]
  469. Sakr, T.M.; Korany, M.; Katti, K.V. Selenium nanomaterials in biomedicine—An overview of new opportunities in nanomedicine of selenium. J. Drug Deliv. Sci. Technol. 2018, 46, 223–233. [Google Scholar] [CrossRef]
  470. Mahmoudvand, H.; Harandi, M.F.; Shakibaie, M.; Aflatoonian, M.R.; ZiaAli, N.; Makki, M.S.; Jahanbakhsh, S. Scolicidal effects of biogenic selenium nanoparticles against protoscolices of hydatid cysts. Int. J. Surg. 2014, 12, 399–403. [Google Scholar] [CrossRef] [Green Version]
  471. Pyrzynska, K.; Sentkowska, A. Biosynthesis of selenium nanoparticles using plant extracts. J. Nanostruct. Chem. 2021, 12, 467–480. [Google Scholar] [CrossRef]
  472. Sonkusre, P.; Nanduri, R. Improved extraction of intracellular biogenic selenium nanoparticles and their specificity for cancer chemoprevention. J. Nanomed. Nanotechnol. 2014, 5, 1. [Google Scholar] [CrossRef]
  473. Uzair, B.; Liaqat, A.; Iqbal, H.; Menaa, B.; Razzaq, A.; Thiripuranathar, G.; Rana, N.F.; Menaa, F. Green and cost-effective synthesis of metallic nanoparticles by algae: Safe methods for translational medicine. Bioengineering 2020, 7, 129. [Google Scholar] [CrossRef] [PubMed]
  474. Bhaumik, J.; Thakur, N.S.; Aili, P.K.; Ghanghoriya, A.; Mittal, A.K.; Banerjee, U.C. Bioinspired nanotheranostic agents: Synthesis, surface functionalization, and antioxidant potential. ACS Biomater. Sci. Eng. 2015, 1, 382–392. [Google Scholar] [CrossRef] [PubMed]
  475. Harsij, M.; Kanani, H.G.; Adineh, H. Effects of antioxidant supplementation (nano-selenium, vitamin C and E) on growth performance, blood biochemistry, immune status and body composition of rainbow trout (Oncorhynchus mykiss) under sub-lethal ammonia exposure. Aquaculture 2020, 521, 734942. [Google Scholar] [CrossRef]
  476. Rathore, S.S.; Murthy, H.S.; Mamun, M.A.-A.; Nasren, S.; Rakesh, K.; Kumar, B.T.N.; Abhiman, P.B.; Khandagale, A.S. Nano-selenium supplementation to ameliorate nutrition physiology, immune response, antioxidant system and disease resistance against Aeromonas hydrophila in monosex Nile tilapia (Oreochromis niloticus). Biol. Trace Elem. Res. 2021, 199, 3073–3088. [Google Scholar] [CrossRef]
  477. Pisoschi, A.M.; Pop, A.; Iordache, F.; Stanca, L.; Predoi, G.; Serban, A.I. Oxidative stress mitigation by antioxidants—An overview on their chemistry and influences on health status. Eur. J. Med. Chem. 2021, 209, 112891. [Google Scholar] [CrossRef]
  478. Gulcin, İ. Antioxidants and antioxidant methods: An updated overview. Arch. Toxicol. 2020, 94, 651–715. [Google Scholar] [CrossRef] [Green Version]
  479. Azeez, L.; Lateef, A.; Adebisi, S.A. Silver nanoparticles (AgNPs) biosynthesized using pod extract of Cola nitida enhances antioxidant activity and phytochemical composition of Amaranthus caudatus Linn. Appl. Nanosci. 2017, 7, 59–66. [Google Scholar] [CrossRef] [Green Version]
  480. Flieger, J.; Flieger, W.; Baj, J.; Maciejewski, R. Antioxidants: Classification, natural sources, activity/capacity measurements, and usefulness for the synthesis of nanoparticles. Materials 2021, 14, 4135. [Google Scholar] [CrossRef]
  481. Guillin, O.M.; Vindry, C.; Ohlmann, T.; Chavatte, L. Interplay between selenium, selenoproteins and HIV-1 replication in human CD4 T-lymphocytes. Int. J. Mol. Sci. 2022, 23, 1394. [Google Scholar] [CrossRef]
  482. Chen, W.; Cheng, H.; Xia, W. Construction of Polygonatum sibiricum polysaccharide functionalized selenium nanoparticles for the enhancement of stability and antioxidant activity. Antioxidants 2022, 11, 240. [Google Scholar] [CrossRef] [PubMed]
  483. Chauhan, S.S.; Ponnampalam, E.N.; Celi, P.; Hopkins, D.L.; Leury, B.J.; Dunshea, F.R. High dietary vitamin E and selenium improves feed intake and weight gain of finisher lambs and maintains redox homeostasis under hot conditions. Small Rumin. Res. 2016, 137, 17–23. [Google Scholar] [CrossRef]
  484. Adeyemi, J.A.; Ogunwole, G.A.; Bamidele, O.S.; Adedire, C.O. Effects of pre-treatment with waterborne selenium on redox homeostasis and humoral innate immune parameters in African catfish, Clarias gariepinus (Burchell, 1822), experimentally challenged with Serratia marcescens. Fish Physiol. Biochem. 2022, 48, 409–418. [Google Scholar] [CrossRef] [PubMed]
  485. Li, Y.; Fan, M.; Qiu, Q.; Wang, Y.; Shen, X.; Zhao, K. Nano-selenium and Macleaya cordata extracts improved immune function and reduced oxidative damage of Sows and IUGR piglets after heat stress of Sows in late gestation. Biol. Trace Elem. Res. 2022, 200, 5081–5090. [Google Scholar] [CrossRef]
  486. Ma, Y.; Shi, Y.; Wang, Y.; Wu, Q.; Cheng, B.; Li, Y.; Wang, Z.; Chai, X.; Ren, A.; Li, G. Role of endoplasmic reticulum stress in nano-selenium alleviating prehierarchical follicular atresia induced by mercury in laying hens. Biol. Trace Elem. Res. 2022, 200, 5205–5217. [Google Scholar] [CrossRef]
  487. Mellinas, C.; Jiménez, A.; Garrigós, M.D.C. Microwave-assisted green synthesis and antioxidant activity of selenium nanoparticles using Theobroma cacao L. bean shell extract. Molecules 2019, 24, 4048. [Google Scholar] [CrossRef] [Green Version]
  488. Wang, Y.-Y.; Qiu, W.-Y.; Sun, L.; Ding, Z.-C.; Yan, J.-K. Preparation, characterization, and antioxidant capacities of selenium nanoparticles stabilized using polysaccharide–protein complexes from Corbicula fluminea. Food Biosci. 2018, 26, 177–184. [Google Scholar] [CrossRef]
  489. Chen, W.; Yue, L.; Jiang, Q.; Xia, W. Effect of chitosan with different molecular weight on the stability, antioxidant and anticancer activities of well-dispersed selenium nanoparticles. IET Nanobiotechnol. 2019, 13, 30–35. [Google Scholar] [CrossRef]
  490. Kokila, K.; Elavarasan, N.; Sujatha, V. Diospyros montana leaf extract-mediated synthesis of selenium nanoparticles and their biological applications. New J. Chem. 2017, 41, 7481–7490. [Google Scholar] [CrossRef]
  491. El-Zayat, M.M.; Eraqi, M.M.; Alrefai, H.; El-Khateeb, A.Y.; Ibrahim, M.A.; Aljohani, H.M.; Aljohani, M.M.; Elshaer, M.M. The antimicrobial, antioxidant, and anticancer activity of greenly synthesized selenium and zinc composite nanoparticles using Ephedra aphylla extract. Biomolecules 2021, 11, 470. [Google Scholar] [CrossRef]
  492. Liu, X.; Mao, Y.; Huang, S.; Li, W.; Zhang, W.; An, J.; Jin, Y.; Guan, J.; Wu, L.; Zhou, P. Selenium nanoparticles derived from Proteus mirabilis YC801 alleviate oxidative stress and inflammatory response to promote nerve repair in rats with spinal cord injury. Regen. Biomater. 2022, 9, rbac042. [Google Scholar] [CrossRef] [PubMed]
  493. Akçay, F.A.; Avcı, A. Effects of process conditions and yeast extract on the synthesis of selenium nanoparticles by a novel indigenous isolate Bacillus sp. EKT1 and characterization of nanoparticles. Arch. Microbiol. 2020, 202, 2233–2243. [Google Scholar] [CrossRef]
  494. Qi, Y.; Yi, P.; He, T.; Song, X.; Liu, Y.; Li, Q.; Zheng, J.; Song, R.; Liu, C.; Zhang, Z.; et al. Quercetin-loaded selenium nanoparticles inhibit amyloid-β aggregation and exhibit antioxidant activity. Colloids Surf. A Physicochem. Eng. Asp. 2020, 602, 125058. [Google Scholar] [CrossRef]
  495. Mittal, A.K.; Kumar, S.; Banerjee, U.C. Quercetin and gallic acid mediated synthesis of bimetallic (silver and selenium) nanoparticles and their antitumor and antimicrobial potential. J. Colloid Interface Sci. 2014, 431, 194–199. [Google Scholar] [CrossRef]
  496. Qiao, L.; Dou, X.; Yan, S.; Zhang, B.; Xu, C. Biogenic selenium nanoparticles synthesized by Lactobacillus casei ATCC 393 alleviate diquat-induced intestinal barrier dysfunction in C57BL/6 mice through their antioxidant activity. Food Funct. 2020, 11, 3020–3031. [Google Scholar] [CrossRef] [PubMed]
  497. Xu, C.; Guo, Y.; Qiao, L.; Ma, L.; Cheng, Y.; Roman, A. Biogenic synthesis of novel functionalized selenium nanoparticles by Lactobacillus casei ATCC 393 and its protective effects on intestinal barrier dysfunction caused by enterotoxigenic Escherichia coli K88. Front. Microbiol. 2018, 9, 1129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  498. Xu, C.; Qiao, L.; Guo, Y.; Ma, L.; Cheng, Y. Preparation, characteristics and antioxidant activity of polysaccharides and proteins-capped selenium nanoparticles synthesized by Lactobacillus casei ATCC 393. Carbohydr. Polym. 2018, 195, 576–585. [Google Scholar] [CrossRef]
  499. Xu, C.; Qiao, L.; Ma, L.; Yan, S.; Guo, Y.; Dou, X.; Zhang, B.; Roman, A. Biosynthesis of polysaccharides-capped selenium nanoparticles using Lactococcus lactis NZ9000 and their antioxidant and anti-inflammatory activities. Front. Microbiol. 2019, 10, 1632. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  500. Zhang, W.; Zhang, J.; Ding, D.; Zhang, L.; Muehlmann, L.A.; Deng, S.-E.; Wang, X.; Li, W.; Zhang, W. Synthesis and antioxidant properties of Lycium barbarum polysaccharides capped selenium nanoparticles using tea extract. Artif. Cells Nanomed. Biotechnol. 2018, 46, 1463–1470. [Google Scholar] [CrossRef]
  501. Menon, S.; Shrudhi Devi, K.S.; Agarwal, H.; Shanmugam, V.K. Efficacy of biogenic selenium nanoparticles from an extract of ginger towards evaluation on anti-microbial and anti-oxidant activities. Colloid Interface Sci. Commun. 2019, 29, 1–8. [Google Scholar] [CrossRef]
  502. Salem, S.S.; Fouda, M.M.G.; Fouda, A.; Awad, M.A.; Al-Olayan, E.M.; Allam, A.A.; Shaheen, T.I. Antibacterial, cytotoxicity and larvicidal activity of green synthesized selenium nanoparticles using Penicillium corylophilum. J. Clust. Sci. 2020, 32, 351–361. [Google Scholar] [CrossRef]
  503. Lin, W.; Zhang, J.; Xu, J.-F.; Pi, J. The Advancing of selenium nanoparticles against infectious diseases. Front. Pharmacol. 2021, 12, 682284. [Google Scholar] [CrossRef] [PubMed]
  504. Souza, L.M.D.S.; Dibo, M.; Sarmiento, J.J.P.; Seabra, A.B.; Medeiros, L.P.; Lourenço, I.M.; Kobayashi, R.K.T.; Nakazato, G. Biosynthesis of selenium nanoparticles using combinations of plant extracts and their antibacterial activity. Curr. Res. Green Sustain. Chem. 2022, 5, 100303. [Google Scholar] [CrossRef]
  505. Zonaro, E.; Lampis, S.; Turner, R.J.; Qazi, S.J.S.; Vallini, G. Biogenic selenium and tellurium nanoparticles synthesized by environmental microbial isolates efficaciously inhibit bacterial planktonic cultures and biofilms. Front. Microbiol. 2015, 6, 584. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  506. Cittrarasu, V.; Kaliannan, D.; Dharman, K.; Maluventhen, V.; Easwaran, M.; Liu, W.C.; Balasubramanian, B.; Arumugam, M. Green synthesis of selenium nanoparticles mediated from Ceropegia bulbosa Roxb extract and its cytotoxicity, antimicrobial, mosquitocidal and photocatalytic activities. Sci. Rep. 2021, 11, 1032. [Google Scholar] [CrossRef] [PubMed]
  507. Sonkusre, P.; Cameotra, S.S. Biogenic selenium nanoparticles inhibit Staphylococcus aureus adherence on different surfaces. Colloids Surf. B Biointerfaces 2015, 136, 1051–1057. [Google Scholar] [CrossRef]
  508. Chudobova, D.; Cihalova, K.; Dostalova, S.; Ruttkay-Nedecky, B.; Rodrigo, M.A.; Tmejova, K.; Kopel, P.; Nejdl, L.; Kudr, J.; Gumulec, J.; et al. Comparison of the effects of silver phosphate and selenium nanoparticles on Staphylococcus aureus growth reveals potential for selenium particles to prevent infection. FEMS Microbiol. Lett. 2014, 351, 195–201. [Google Scholar] [CrossRef] [Green Version]
  509. Alghuthaymi, M.A.; Diab, A.M.; Elzahy, A.F.; Mazrou, K.E.; Tayel, A.A.; Moussa, S.H. Green biosynthesized selenium nanoparticles by cinnamon extract and their antimicrobial activity and application as edible coatings with nano-chitosan. J. Food Qual. 2021, 2021, 6670709. [Google Scholar] [CrossRef]
  510. Menon, S.; Agarwal, H.; Rajeshkumar, S.; Rosy, P.J.; Shanmugam, V.K. Investigating the antimicrobial activities of the biosynthesized selenium nanoparticles and its statistical analysis. Bionanoscience 2020, 10, 122–135. [Google Scholar] [CrossRef]
  511. Kheradmand, E.; Rafii, F.; Yazdi, M.H.; Sepahi, A.A.; Shahverdi, A.R.; Oveisi, M.R. The antimicrobial effects of selenium nanoparticle-enriched probiotics and their fermented broth against Candida albicans. DARU J. Pharm. Sci. 2014, 22, 48. [Google Scholar] [CrossRef] [Green Version]
  512. Hu, D.; Yu, S.; Yu, D.; Liu, N.; Tang, Y.; Fan, Y.; Wang, C.; Wu, A. Biogenic Trichoderma harzianum-derived selenium nanoparticles with control functionalities originating from diverse recognition metabolites against phytopathogens and mycotoxins. Food Control 2019, 106, 106748. [Google Scholar] [CrossRef]
  513. Shakibaie, M.; Ezzatkhah, F.; Gabal, E.; Badparva, E.; Jahanbakhsh, S.; Mahmoudvand, H. Prophylactic effects of biogenic selenium nanoparticles on acute toxoplasmosis: An in vivo study. Ann. Med. Surg. 2020, 54, 85–88. [Google Scholar] [CrossRef] [PubMed]
  514. Beheshti, N.; Soflaei, S.; Shakibaie, M.; Yazdi, M.H.; Ghaffarifar, F.; Dalimi, A.; Shahverdi, A.R. Efficacy of biogenic selenium nanoparticles against Leishmania major: In vitro and in vivo studies. J. Trace Elem. Med. Biol. 2013, 27, 203–207. [Google Scholar] [CrossRef] [PubMed]
  515. Vahidi, H.; Barabadi, H.; Saravanan, M. Emerging selenium nanoparticles to combat cancer: A systematic review. J. Clust. Sci. 2019, 31, 301–309. [Google Scholar] [CrossRef]
  516. Xu, X.; Ho, W.; Zhang, X.; Bertrand, N.; Farokhzad, O. Cancer nanomedicine: From targeted delivery to combination therapy. Trends Mol. Med. 2015, 21, 223–232. [Google Scholar] [CrossRef] [Green Version]
  517. Nie, S. Understanding and overcoming major barriers in cancer nanomedicine. Nanomedicine 2010, 5, 523–528. [Google Scholar] [CrossRef] [Green Version]
  518. Barabadi, H.; Ovais, M.; Shinwari, Z.K.; Saravanan, M. Anti-cancer green bionanomaterials: Present status and future prospects. Green Chem. Lett. Rev. 2017, 10, 285–314. [Google Scholar] [CrossRef] [Green Version]
  519. Song, X.; Chen, Y.; Zhao, G.; Sun, H.; Che, H.; Leng, X. Effect of molecular weight of chitosan and its oligosaccharides on antitumor activities of chitosan-selenium nanoparticles. Carbohydr. Polym. 2020, 231, 115689. [Google Scholar] [CrossRef]
  520. Ullah, A.; Mu, J.; Wang, F.; Chan, M.W.H.; Yin, X.; Liao, Y.; Mirani, Z.A.; Sebt-E-Hassan, S.; Aslam, S.; Naveed, M.; et al. Biogenic selenium nanoparticles and their anticancer effects pertaining to probiotic bacteria—A Review. Antioxidants 2022, 11, 1916. [Google Scholar] [CrossRef]
  521. Sholkamy, E.; Ahmad, M.; Yaser, M.M.; Ali, A.; Mehanni, M. Anticancer activity of biostabilized selenium nanorods synthesized by Streptomyces bikiniensis strain Ess_amA-1. Int. J. Nanomed. 2015, 10, 3389–3401. [Google Scholar] [CrossRef] [Green Version]
  522. Anu, K.; Singaravelu, G.; Murugan, K.; Benelli, G. Green-synthesis of selenium nanoparticles using garlic cloves (Allium sativum): Biophysical characterization and cytotoxicity on Vero cells. J. Clust. Sci. 2016, 28, 551–563. [Google Scholar] [CrossRef]
  523. Cruz, L.Y.; Wang, D.; Liu, J. Biosynthesis of selenium nanoparticles, characterization and X-ray induced radiotherapy for the treatment of lung cancer with interstitial lung disease. J. Photochem. Photobiol. B Biol. 2019, 191, 123–127. [Google Scholar] [CrossRef] [PubMed]
  524. Sonkusre, P. Specificity of biogenic selenium nanoparticles for prostate cancer therapy with reduced risk of toxicity: An in vitro and in vivo study. Front. Oncol. 2020, 9, 1541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  525. Ramamurthy, C.; Sampath, K.S.; Arunkumar, P.; Kumar, M.S.; Sujatha, V.; Premkumar, K.; Thirunavukkarasu, C. Green synthesis and characterization of selenium nanoparticles and its augmented cytotoxicity with doxorubicin on cancer cells. Bioprocess Biosyst. Eng. 2013, 36, 1131–1139. [Google Scholar] [CrossRef]
  526. Srivastava, P.; Kowshik, M. Anti-neoplastic selenium nanoparticles from Idiomarina sp. PR58-8. Enzym. Microb. Technol. 2016, 95, 192–200. [Google Scholar] [CrossRef]
  527. A Wadhwani, S.; Gorain, M.; Banerjee, P.; Shedbalkar, U.U.; Singh, R.; Kundu, G.C.; A Chopade, B. Green synthesis of selenium nanoparticles using Acinetobacter sp. SW30: Optimization, characterization and its anticancer activity in breast cancer cells. Int. J. Nanomed. 2017, 12, 6841–6855. [Google Scholar] [CrossRef] [Green Version]
  528. Shakibaie, M.; Khorramizadeh, M.; Faramarzi, M.A.; Sabzevari, O.; Shahverdi, A.R. Biosynthesis and recovery of selenium nanoparticles and the effects on matrix metalloproteinase-2 expression. Biotechnol. Appl. Biochem. 2010, 56, 7–15. [Google Scholar] [CrossRef]
  529. Cui, D.; Liang, T.; Sun, L.; Meng, L.; Yang, C.; Wang, L.; Liang, T.; Li, Q. Green synthesis of selenium nanoparticles with extract of hawthorn fruit induced HepG2 cells apoptosis. Pharm. Biol. 2018, 56, 528–534. [Google Scholar] [CrossRef] [Green Version]
  530. Rajasekar, S.; Kuppusamy, S. Eco-friendly formulation of selenium nanoparticles and its functional characterization against breast cancer and normal cells. J. Clust. Sci. 2020, 32, 907–915. [Google Scholar] [CrossRef]
  531. Krishnan, V.; Loganathan, C.; Thayumanavan, P. Green synthesized selenium nanoparticle as carrier and potent delivering agent of s-allyl glutathione: Anticancer effect against hepatocarcinoma cell line (HepG2) through induction of cell cycle arrest and apoptosis. J. Drug Deliv. Sci. Technol. 2019, 53, 101207. [Google Scholar] [CrossRef]
  532. Chen, T.; Wong, Y.-S.; Zheng, W.; Bai, Y.; Huang, L. Selenium nanoparticles fabricated in Undaria pinnatifida polysaccharide solutions induce mitochondria-mediated apoptosis in A375 human melanoma cells. Colloids Surf. B Biointerfaces 2008, 67, 26–31. [Google Scholar] [CrossRef] [PubMed]
  533. Jia, X.; Liu, Q.; Zou, S.; Xu, X.; Zhang, L. Construction of selenium nanoparticles/β-glucan composites for enhancement of the antitumor activity. Carbohydr. Polym. 2015, 117, 434–442. [Google Scholar] [CrossRef] [PubMed]
  534. Nonsuwan, P.; Puthong, S.; Palaga, T.; Muangsin, N. Novel organic/inorganic hybrid flower-like structure of selenium nanoparticles stabilized by pullulan derivatives. Carbohydr. Polym. 2018, 184, 9–19. [Google Scholar] [CrossRef] [PubMed]
  535. Ren, J.; Liao, W.; Zhang, R.; Dong, C.; Yu, Z. Novel walnut peptide–selenium hybrids with enhanced anticancer synergism: Facile synthesis and mechanistic investigation of anticancer activity. Int. J. Nanomed. 2016, 11, 1305–1321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  536. Fan, T.J.; Han, L.H.; Cong, R.S.; Liang, J. Caspase family proteases and apoptosis. Acta Biochim. Biophys. Sin. 2005, 37, 719–727. [Google Scholar] [CrossRef] [Green Version]
  537. Spyridopoulou, K.; Aindelis, G.; Pappa, A.; Chlichlia, K. Anticancer activity of biogenic selenium nanoparticles: Apoptotic and immunogenic cell death markers in colon cancer cells. Cancers 2021, 13, 5335. [Google Scholar] [CrossRef] [PubMed]
  538. Domínguez-Álvarez, E.; Rácz, B.; Marć, M.A.; Nasim, M.J.; Szemerédi, N.; Viktorová, J.; Jacob, C.; Spengler, G. Selenium and tellurium in the development of novel small molecules and nanoparticles as cancer multidrug resistance reversal agents. Drug Resist. Updat. 2022, 63, 100844. [Google Scholar] [CrossRef]
  539. Senapati, S.; Mahanta, A.K.; Kumar, S.; Maiti, P. Controlled drug delivery vehicles for cancer treatment and their performance. Signal Transduct. Target. Ther. 2018, 3, 7. [Google Scholar] [CrossRef] [Green Version]
  540. Vekariya, K.K.; Kaur, J.; Tikoo, K. Alleviating anastrozole induced bone toxicity by selenium nanoparticles in SD rats. Toxicol. Appl. Pharmacol. 2013, 268, 212–220. [Google Scholar] [CrossRef]
  541. Shirsat, S.; Kadam, A.; Mane, R.S.; Jadhav, V.V.; Zate, M.K.; Naushad, M.; Kim, K.H. Protective role of biogenic selenium nanoparticles in immunological and oxidative stress generated by enrofloxacin in broiler chicken. Dalton Trans. 2016, 45, 8845–8853. [Google Scholar] [CrossRef]
  542. Dahdouh, F.; Bendjeffal, H.; Nouacer, Z.; Moumene, W.; Zeminour, M.E.-H.; Naous, M.; Djebar, H. Selenium nanoparticles attenuate gentamycin-induced nephrotoxicity and hematotoxicity in female Swiss albino mice. Bionanoscience 2019, 9, 356–364. [Google Scholar] [CrossRef]
  543. Ghosh, S. Cisplatin: The first metal based anticancer drug. Bioorg. Chem. 2019, 88, 102925. [Google Scholar] [CrossRef] [PubMed]
  544. Rezvanfar, M.A.; Shahverdi, A.R.; Ahmadi, A.; Baeeri, M.; Mohammadirad, A.; Abdollahi, M. Protection of cisplatin-induced spermatotoxicity, DNA damage and chromatin abnormality by selenium nano-particles. Toxicol. Appl. Pharmacol. 2013, 266, 356–365. [Google Scholar] [CrossRef] [PubMed]
  545. Li, Y.; Li, X.; Wong, Y.-S.; Chen, T.; Zhang, H.; Liu, C.; Zheng, W. The reversal of cisplatin-induced nephrotoxicity by selenium nanoparticles functionalized with 11-mercapto-1-undecanol by inhibition of ROS-mediated apoptosis. Biomaterials 2011, 32, 9068–9076. [Google Scholar] [CrossRef]
  546. Saif-Elnasr, M.; Abdel-Aziz, N.; El-Batal, A.I. Ameliorative effect of selenium nanoparticles and fish oil on cisplatin and gamma irradiation-induced nephrotoxicity in male albino rats. Drug Chem. Toxicol. 2019, 42, 94–103. [Google Scholar] [CrossRef]
  547. Li, Y.; Li, X.; Zheng, W.; Fan, C.; Zhang, Y.; Chen, T. Functionalized selenium nanoparticles with nephroprotective activity, the important roles of ROS-mediated signaling pathways. J. Mater. Chem. B 2013, 1, 6365–6372. [Google Scholar] [CrossRef] [PubMed]
  548. Kumar, G.S.; Kulkarni, A.; Khurana, A.; Kaur, J.; Tikoo, K. Selenium nanoparticles involve HSP-70 and SIRT1 in preventing the progression of type 1 diabetic nephropathy. Chem. Interact. 2014, 223, 125–133. [Google Scholar] [CrossRef] [PubMed]
  549. Sadeghian, S.; Kojouri, G.A.; Mohebbi, A. Nanoparticles of selenium as species with stronger physiological effects in sheep in comparison with sodium selenite. Biol. Trace Elem. Res. 2012, 146, 302–308. [Google Scholar] [CrossRef] [Green Version]
  550. Sadek, K.M.; Lebda, M.A.; Abouzed, T.K.; Nasr, S.M.; Shoukry, M. Neuro- and nephrotoxicity of subchronic cadmium chloride exposure and the potential chemoprotective effects of selenium nanoparticles. Metab. Brain Dis. 2017, 32, 1659–1673. [Google Scholar] [CrossRef]
  551. Zhang, C.; Huang, Y.; Talukder, M.; Ge, J.; Lv, M.-W.; Bi, S.-S.; Li, J.-L. Selenium sources differ in their potential to alleviate the cadmium-induced testicular dysfunction. Environ. Pollut. 2020, 267, 115610. [Google Scholar] [CrossRef]
  552. Qamar, N.; John, P.; Bhatti, A. Toxicological and anti-rheumatic potential of Trachyspermum ammi derived biogenic selenium nanoparticles in arthritic balb/c mice. Int. J. Nanomed. 2020, 15, 3497–3509. [Google Scholar] [CrossRef] [PubMed]
  553. Ren, S.-X.; Zhang, B.; Lin, Y.; Ma, D.-S.; Yan, H. Selenium nanoparticles dispersed in phytochemical exert anti-inflammatory activity by modulating catalase, GPx1, and COX-2 gene expression in a rheumatoid arthritis rat model. J. Pharmacol. Exp. Ther. 2019, 25, 991–1000. [Google Scholar] [CrossRef] [PubMed]
  554. Yuan, X.; Fu, Z.; Ji, P.; Guo, L.; O Al-Ghamdy, A.; Alkandiri, A.; A Habotta, O.; Moneim, A.E.A.; Kassab, R.B. Selenium nanoparticles pre-treatment reverse behavioral, oxidative damage, neuronal loss and neurochemical alterations in pentylenetetrazole-induced epileptic seizures in mice. Int. J. Nanomed. 2020, 15, 6339–6353. [Google Scholar] [CrossRef] [PubMed]
  555. A Alkhudhayri, A.; A Dkhil, M.; Al-Quraishy, S. Nanoselenium prevents eimeriosis-induced inflammation and regulates mucin gene expression in mice jejunum. Int. J. Nanomed. 2018, 13, 1993–2003. [Google Scholar] [CrossRef] [Green Version]
  556. El-Ghazaly, M.; Fadel, N.; Rashed, E.; El-Batal, A.; Kenawy, S. Anti-inflammatory effect of selenium nanoparticles on the inflammation induced in irradiated rats. Can. J. Physiol. Pharmacol. 2017, 95, 101–110. [Google Scholar] [CrossRef]
  557. Ebokaiwe, A.P.; Okori, S.; Nwankwo, J.O.; Ejike, C.E.C.C.; Osawe, S.O. Selenium nanoparticles and metformin ameliorate streptozotocin-instigated brain oxidative-inflammatory stress and neurobehavioral alterations in rats. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2021, 394, 591–602. [Google Scholar] [CrossRef]
  558. Wynn, T.A.; Chawla, A.; Pollard, J.W. Macrophage biology in development, homeostasis and disease. Nature 2013, 496, 445–455. [Google Scholar] [CrossRef] [Green Version]
  559. Zhu, C.; Zhang, S.; Song, C.; Zhang, Y.; Ling, Q.; Hoffmann, P.R.; Li, J.; Chen, T.; Zheng, W.; Huang, Z. Selenium nanoparticles decorated with Ulva lactuca polysaccharide potentially attenuate colitis by inhibiting NF-kB mediated hyper inflammation. J. Nanobiotechnol. 2017, 15, 20. [Google Scholar] [CrossRef] [Green Version]
  560. Lu, K.Y.; Lin, P.Y.; Chuang, E.Y.; Shih, C.M.; Cheng, T.M.; Lin, T.Y.; Sung, H.W.; Mi, F.L. H2O2-depleting and O2-generating selenium nanoparticles for fluorescence imaging and photodynamic treatment of proinflammatory-activated macrophages. ACS Appl. Mater. Interfaces 2017, 9, 5158–5172. [Google Scholar] [CrossRef]
  561. Wang, J.; Zhang, Y.; Yuan, Y.; Yue, T. Immunomodulatory of selenium nano-particles decorated by sulfated Ganoderma lucidum polysaccharides. Food Chem. Toxicol. 2014, 68, 183–189. [Google Scholar] [CrossRef]
  562. Simos, Y.V.; Spyrou, K.; Patila, M.; Karouta, N.; Stamatis, H.; Gournis, D.; Dounousi, E.; Peschos, D. Trends of nanotechnology in type 2 diabetes mellitus treatment. Asian J. Pharm. Sci. 2021, 16, 62–76. [Google Scholar] [CrossRef] [PubMed]
  563. Veiseh, O.; Tang, B.C.; Whitehead, K.A.; Anderson, D.G.; Langer, R. Managing diabetes with nanomedicine: Challenges and opportunities. Nat. Rev. Drug Discov. 2015, 14, 45–57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  564. Wei, J.; Zeng, C.; Gong, Q.-Y.; Yang, H.-B.; Li, X.-X.; Lei, G.-H.; Yang, T.-B. The association between dietary selenium intake and diabetes: A cross-sectional study among middle-aged and older adults. Nutr. J. 2015, 14, 18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  565. Sokmen, B.B.; Basaraner, H.; Yanardag, R. Combined effects of treatment with vitamin C, vitamin E and selenium on the skin of diabetic rats. Hum. Exp. Toxicol. 2013, 32, 379–384. [Google Scholar] [CrossRef]
  566. Rayman, M.P.; Stranges, S. Epidemiology of selenium and type 2 diabetes: Can we make sense of it? Free Radic. Biol. Med. 2013, 65, 1557–1564. [Google Scholar] [CrossRef]
  567. Zhou, J.; Huang, K.; Lei, X.G. Selenium and diabetes—Evidence from animal studies. Free Radic. Biol. Med. 2013, 65, 1548–1556. [Google Scholar] [CrossRef] [Green Version]
  568. Yuan, Y.; Xiao, Y.; Yu, Y.; Liu, Y.; Feng, W.; Qiu, G.; Wang, H.; Liu, B.; Wang, J.; Zhou, L.; et al. Associations of multiple plasma metals with incident type 2 diabetes in Chinese adults: The Dongfeng-Tongji Cohort. Environ. Pollut. 2018, 237, 917–925. [Google Scholar] [CrossRef]
  569. Galan-Chilet, I.; Grau-Perez, M.; De Marco, G.; Guallar, E.; Martin-Escudero, J.C.; Dominguez-Lucas, A.; Gonzalez-Manzano, I.; Lopez-Izquierdo, R.; Briongos-Figuero, L.S.; Redon, J.; et al. A gene-environment interaction analysis of plasma selenium with prevalent and incident diabetes: The Hortega study. Redox Biol. 2017, 12, 798–805. [Google Scholar] [CrossRef]
  570. Zhou, C.; Na, L.; Shan, R.; Cheng, Y.; Li, Y.; Wu, X.; Sun, C. Dietary vitamin C intake reduces the risk of type 2 diabetes in Chinese adults: HOMA-IR and T-AOC as potential mediators. PLoS ONE 2016, 11, e0163571. [Google Scholar] [CrossRef] [Green Version]
  571. Bahmani, F.; Kia, M.; Soleimani, A.; Mohammadi, A.A.; Asemi, Z. The effects of selenium supplementation on biomarkers of inflammation and oxidative stress in patients with diabetic nephropathy: A randomised, double-blind, placebo-controlled trial. Br. J. Nutr. 2016, 116, 1222–1228. [Google Scholar] [CrossRef]
  572. Oztürk, Z.; Gurpinar, T.; Vural, K.; Boyacıoglu, S.; Korkmaz, M.; Var, A. Effects of selenium on endothelial dysfunction and metabolic profile in low dose streptozotocin induced diabetic rats fed a high fat diet. Biotech. Histochem. 2015, 90, 506–515. [Google Scholar] [CrossRef] [PubMed]
  573. Zhao, S.-J.; Wang, D.-H.; Li, Y.-W.; Han, L.; Xiao, X.; Ma, M.; Wan, D.C.-C.; Hong, A.; Ma, Y. A novel selective VPAC2 agonist peptide-conjugated chitosan modified selenium nanoparticles with enhanced anti-type 2 diabetes synergy effects. Int. J. Nanomed. 2017, 12, 2143–2160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  574. Ahmed, H.H.; El-Maksoud, M.D.A.; Moneim, A.E.A.; Aglan, H.A. Pre-clinical study for the antidiabetic potential of selenium nanoparticles. Biol. Trace Elem. Res. 2017, 177, 267–280. [Google Scholar] [CrossRef] [PubMed]
  575. Abd El-Hakim, Y.M.; Abdel-Rahman Mohamed, A.; Khater, S.I.; Hamed Arisha, A.; Metwally, M.M.M.; Nassan, M.A.; Hassan, M.E. Chitosan-stabilized selenium nanoparticles and metformin synergistically rescue testicular oxidative damage and steroidogenesis-related genes dysregulation in high-fat diet/streptozotocin-induced diabetic rats. Antioxidants 2021, 10, 17. [Google Scholar] [CrossRef]
  576. Mohamed, A.A.-R.; Khater, S.I.; Arisha, A.H.; Metwally, M.M.; Mostafa-Hedeab, G.; El-Shetry, E.S. Chitosan-stabilized selenium nanoparticles alleviate cardio-hepatic damage in type 2 diabetes mellitus model via regulation of caspase, Bax/Bcl-2, and Fas/FasL-pathway. Gene 2020, 768, 145288. [Google Scholar] [CrossRef]
  577. Golmohammadi, R.; Najar-Peerayeh, S.; Moghadam, T.T.; Hosseini, S.M.J. Synergistic antibacterial activity and wound healing properties of selenium-chitosan-mupirocin nanohybrid system: An in vivo study on rat diabetic Staphylococcus aureus wound infection model. Sci. Rep. 2020, 10, 2854. [Google Scholar] [CrossRef] [Green Version]
  578. Shakeri-Zadeh, A.; Zareyi, H.; Sheervalilou, R.; Laurent, S.; Ghaznavi, H.; Samadian, H. Gold nanoparticle-mediated bubbles in cancer nanotechnology. J. Control. Release 2021, 330, 49–60. [Google Scholar] [CrossRef]
  579. Bromma, K.; Chithrani, D.B. Advances in gold nanoparticle-based combined cancer therapy. Nanomaterials 2020, 10, 1671. [Google Scholar] [CrossRef]
  580. Kerry, R.G.; Mahapatra, G.P.; Maurya, G.K.; Patra, S.; Mahari, S.; Das, G.; Patra, J.K.; Sahoo, S. Molecular prospect of type-2 diabetes: Nanotechnology based diagnostics and therapeutic intervention. Rev. Endocr. Metab. Disord. 2021, 22, 421–451. [Google Scholar] [CrossRef]
  581. Souto, E.B.; Souto, S.B.; Campos, J.R.; Severino, P.; Pashirova, T.N.; Zakharova, L.Y.; Silva, A.M.; Durazzo, A.; Lucarini, M.; Izzo, A.A.; et al. Nanoparticle delivery systems in the treatment of diabetes complications. Molecules 2019, 24, 4209. [Google Scholar] [CrossRef] [Green Version]
  582. Qiu, A.; Wang, Y.; Zhang, G.; Wang, H. Natural polysaccharide-based nanodrug delivery systems for treatment of diabetes. Polymers 2022, 14, 3217. [Google Scholar] [CrossRef] [PubMed]
  583. Jia, W.; Xu, L.; Xu, W.; Yang, M.; Zhang, Y. Application of nanotechnology in the diagnosis and treatment of acute pancreatitis. Nanoscale Adv. 2022, 4, 1949–1961. [Google Scholar] [CrossRef] [PubMed]
  584. Fortina, P.; Kricka, L.J.; Graves, D.J.; Park, J.; Hyslop, T.; Tam, F.; Halas, N.; Surrey, S.; Waldman, S.A. Applications of nanoparticles to diagnostics and therapeutics in colorectal cancer. Trends Biotechnol. 2007, 25, 145–152. [Google Scholar] [CrossRef] [PubMed]
  585. Yang, M.; Zhang, Y.; Ma, Y.; Yan, X.; Gong, L.; Zhang, M.; Zhang, B. Nanoparticle-based therapeutics of inflammatory bowel diseases: A narrative review of the current state and prospects. J. Bio-X Res. 2020, 3, 157–173. [Google Scholar] [CrossRef]
  586. Qasim, M.; Lim, D.-J.; Park, H.; Na, D. Nanotechnology for diagnosis and treatment of infectious diseases. J. Nanosci. Nanotechnol. 2014, 14, 7374–7387. [Google Scholar] [CrossRef]
  587. Zhong, W.; Zhang, X.; Zeng, Y.; Lin, D.; Wu, J. Recent applications and strategies in nanotechnology for lung diseases. Nano Res. 2021, 14, 2067–2089. [Google Scholar] [CrossRef]
  588. Tobin, E.; Brenner, S. Nanotechnology fundamentals applied to clinical infectious diseases and public health. Open Forum Infect. Dis. 2021, 8, ofab583. [Google Scholar] [CrossRef]
  589. Homayoonnia, S.; Lee, Y.; Andalib, D.; Rahman, S.; Shin, J.; Kim, K.; Kim, S. Micro/nanotechnology-inspired rapid diagnosis of respiratory infectious diseases. Biomed. Eng. Lett. 2021, 11, 335–365. [Google Scholar] [CrossRef]
  590. Pehlivan, S.B. Nanotechnology-based drug delivery systems for targeting, imaging and diagnosis of neurodegenerative diseases. Pharm. Res. 2013, 30, 2499–2511. [Google Scholar] [CrossRef]
  591. Re, F.; Gregori, M.; Masserini, M. Nanotechnology for neurodegenerative disorders. Maturitas 2012, 73, 45–51. [Google Scholar] [CrossRef]
  592. Tapeinos, C. Graphene-based nanotechnology in neurodegenerative disorders. Adv. NanoBiomed Res. 2021, 1, 2000059. [Google Scholar] [CrossRef]
  593. De Andrade, V.; Nikitin, V.; Wojcik, M.; Deriy, A.; Bean, S.; Shu, D.; Mooney, T.; Peterson, K.; Kc, P.; Li, K.; et al. Fast X-ray Nanotomography with sub-10 nm resolution as a powerful imaging tool for nanotechnology and energy storage applications. Adv. Mater. 2021, 33, 2008653. [Google Scholar] [CrossRef] [PubMed]
  594. Song, Y.; Huang, Y.-Y.; Liu, X.; Zhang, X.; Ferrari, M.; Qin, L. Point-of-care technologies for molecular diagnostics using a drop of blood. Trends Biotechnol. 2014, 32, 132–139. [Google Scholar] [CrossRef] [Green Version]
  595. Bayford, R.H.; Rademacher, T.; Roitt, I.; Wang, S.X. Emerging applications of nanotechnology for diagnosis and therapy of disease: A review. Physiol. Meas. 2017, 38, R183–R203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  596. Kim, D.; Kim, J.; Park, Y.I.; Lee, N.; Hyeon, T. Recent development of inorganic nanoparticles for biomedical imaging. ACS Central Sci. 2018, 4, 324–336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  597. Han, X.; Xu, K.; Taratula, O.; Farsad, K. Applications of nanoparticles in biomedical imaging. Nanoscale 2019, 11, 799–819. [Google Scholar] [CrossRef]
  598. Kolay, A.; Maity, D.; Ghosal, P.; Deepa, M. Selenium nanoparticle-decorated silicon nanowires with enhanced liquid-junction photoelectrochemical solar cell performance. J. Phys. Chem. C 2019, 123, 8614–8622. [Google Scholar] [CrossRef]
  599. Malekzad, H.; Zangabad, P.S.; Mirshekari, H.; Karimi, M.; Hamblin, M.R. Noble metal nanoparticles in biosensors: Recent studies and applications. Nanotechnol. Rev. 2017, 6, 301–329. [Google Scholar] [CrossRef]
  600. Yang, T.; Duncan, T.V. Challenges and potential solutions for nanosensors intended for use with foods. Nat. Nanotechnol. 2021, 16, 251–265. [Google Scholar] [CrossRef]
  601. Zhao, S.; Huang, J.; Lei, J.; Huo, D.; Huang, Q.; Tan, J.; Li, Y.; Hou, C.; Tian, F. A portable and automatic dual-readout detector integrated with 3D-printed microfluidic nanosensors for rapid carbamate pesticides detection. Sens. Actuators B Chem. 2021, 346, 130454. [Google Scholar] [CrossRef]
  602. Sargazi, S.; Fatima, I.; Kiani, M.H.; Mohammadzadeh, V.; Arshad, R.; Bilal, M.; Rahdar, A.; Díez-Pascual, A.M.; Behzadmehr, R. Fluorescent-based nanosensors for selective detection of a wide range of biological macromolecules: A comprehensive review. Int. J. Biol. Macromol. 2022, 206, 115–147. [Google Scholar] [CrossRef] [PubMed]
  603. Steffens, C.; Ballen, S.C.; Scapin, E.; da Silva, D.M.; Steffens, J.; Jacques, R.A. Advances of nanobiosensors and its application in atrazine detection in water: A review. Sens. Actuators Rep. 2022, 4, 100096. [Google Scholar] [CrossRef]
  604. Tessaro, L.; Aquino, A.; Rodrigues, P.D.A.; Joshi, N.; Ferrari, R.G.; Conte-Junior, C.A. Nucleic acid-based nanobiosensor (NAB) used for Salmonella detection in foods: A Systematic Review. Nanomaterials 2022, 12, 821. [Google Scholar] [CrossRef] [PubMed]
  605. Farouq MA, H.; Al Qaraghuli, M.M.; Kubiak-Ossowska, K.; Ferro, V.A.; Mulheran, P.A. Biomolecular interactions with nanoparticles: Applications for COVID-2019. Curr. Opin. Colloid Interface Sci. 2021, 54, 101461. [Google Scholar] [CrossRef]
  606. Yu, L.; Banerjee, I.A.; Gao, X.; Nuraje, N.; Matsui, H. Fabrication and application of enzyme-incorporated peptide nanotubes. Bioconjug. Chem. 2005, 16, 1484–1487. [Google Scholar] [CrossRef]
  607. Rouse, I.; Lobaskin, V. A hard-sphere model of protein corona formation on spherical and cylindrical nanoparticles. Biophys. J. 2021, 120, 4457–4471. [Google Scholar] [CrossRef]
  608. Malik, P.; Katyal, V.; Malik, V.; Asatkar, A.; Inwati, G.; Mukherjee, T.K. Nanobiosensors: Concepts and variations. ISRN Nanomater. 2013, 2013, 327435. [Google Scholar] [CrossRef]
  609. Naresh, V.; Lee, N. A Review on biosensors and recent development of nanostructured materials-enabled biosensors. Sensors 2021, 21, 1109. [Google Scholar] [CrossRef]
  610. Kaur, N.; Bhardwaj, P.; Singh, G.; Arya, S.K. Applicative insights on nascent role of biochar production, tailoring and immobilization in enzyme industry—A review. Process Biochem. 2021, 107, 153–163. [Google Scholar] [CrossRef]
  611. Lyu, X.; Gonzalez, R.; Horton, A.; Li, T. Immobilization of enzymes by polymeric materials. Catalysts 2021, 11, 1211. [Google Scholar] [CrossRef]
  612. Rafiee, F.; Rezaee, M. Different strategies for the lipase immobilization on the chitosan based supports and their applications. Int. J. Biol. Macromol. 2021, 179, 170–195. [Google Scholar] [CrossRef] [PubMed]
  613. Jun, L.Y.; Yon, L.S.; Mubarak, N.; Bing, C.H.; Pan, S.; Danquah, M.K.; Abdullah, E.; Khalid, M. An overview of immobilized enzyme technologies for dye and phenolic removal from wastewater. J. Environ. Chem. Eng. 2019, 7, 102961. [Google Scholar] [CrossRef]
  614. Gan, J.; Bagheri, A.R.; Aramesh, N.; Gul, I.; Franco, M.; Almulaiky, Y.Q.; Bilal, M. Covalent organic frameworks as emerging host platforms for enzyme immobilization and robust biocatalysis—A review. Int. J. Biol. Macromol. 2021, 167, 502–515. [Google Scholar] [CrossRef] [PubMed]
  615. Alvarado-Ramírez, L.; Rostro-Alanis, M.; Rodríguez-Rodríguez, J.; Sosa-Hernández, J.E.; Melchor-Martínez, E.M.; Iqbal, H.M.N.; Parra-Saldívar, R. Enzyme (single and multiple) and nanozyme biosensors: Recent developments and their novel applications in the water-food-health nexus. Biosensors 2021, 11, 410. [Google Scholar] [CrossRef]
  616. Yamaguchi, H.; Kiyota, Y.; Miyazaki, M. Techniques for preparation of cross-linked enzyme aggregates and their applications in bioconversions. Catalysts 2018, 8, 174. [Google Scholar] [CrossRef] [Green Version]
  617. Fopase, R.; Paramasivam, S.; Kale, P.; Paramasivan, B. Strategies, challenges and opportunities of enzyme immobilization on porous silicon for biosensing applications. J. Environ. Chem. Eng. 2020, 8, 104266. [Google Scholar] [CrossRef]
  618. Chapman, J.; Ismail, A.E.; Dinu, C.Z. Industrial Applications of Enzymes: Recent Advances, Techniques, and Outlooks. Catalysts 2018, 8, 238. [Google Scholar] [CrossRef] [Green Version]
  619. Karthik, V.; Kumar, P.S.; Vo, D.-V.N.; Selvakumar, P.; Gokulakrishnan, M.; Keerthana, P.; Audilakshmi, V.; Jeyanthi, J. Enzyme-loaded nanoparticles for the degradation of wastewater contaminants: A review. Environ. Chem. Lett. 2021, 19, 2331–2350. [Google Scholar] [CrossRef]
  620. Wong, J.K.H.; Tan, H.K.; Lau, S.Y.; Yap, P.-S.; Danquah, M.K. Potential and challenges of enzyme incorporated nanotechnology in dye wastewater treatment: A review. J. Environ. Chem. Eng. 2019, 7, 103261. [Google Scholar] [CrossRef]
  621. Meena, J.; Gupta, A.; Ahuja, R.; Singh, M.; Panda, A.K. Recent advances in nano-engineered approaches used for enzyme immobilization with enhanced activity. J. Mol. Liq. 2021, 338, 116602. [Google Scholar] [CrossRef]
  622. Pandit, C.; Roy, A.; Ghotekar, S.; Khusro, A.; Islam, M.N.; Bin Emran, T.; Lam, S.E.; Khandaker, M.U.; Bradley, D.A. Biological agents for synthesis of nanoparticles and their applications. J. King Saud Univ.-Sci. 2022, 34, 101869. [Google Scholar] [CrossRef]
  623. Dahoumane, S.A.; Mechouet, M.; Wijesekera, K.; Filipe, C.D.M.; Sicard, C.; Bazylinski, D.A.; Jeffryes, C. Algae-mediated biosynthesis of inorganic nanomaterials as a promising route in nanobiotechnology—A review. Green Chem. 2017, 19, 552–587. [Google Scholar] [CrossRef]
  624. Singh, K.R.; Nayak, V.; Singh, J.; Singh, A.K.; Singh, R.P. Potentialities of bioinspired metal and metal oxide nanoparticles in biomedical sciences. RSC Adv. 2021, 11, 24722–24746. [Google Scholar] [CrossRef] [PubMed]
  625. Andleeb, A.; Andleeb, A.; Asghar, S.; Zaman, G.; Tariq, M.; Mehmood, A.; Nadeem, M.; Hano, C.; Lorenzo, J.; Abbasi, B. A systematic review of biosynthesized metallic nanoparticles as a promising anti-cancer-strategy. Cancers 2021, 13, 2818. [Google Scholar] [CrossRef]
  626. Rezazadeh, N.H.; Buazar, F.; Matroodi, S. Synergistic effects of combinatorial chitosan and polyphenol biomolecules on enhanced antibacterial activity of biofunctionalaized silver nanoparticles. Sci. Rep. 2020, 10, 19615. [Google Scholar] [CrossRef]
  627. Rahman, A.; Kumar, S.; Bafana, A.; Dahoumane, S.A.; Jeffryes, C. Biosynthetic conversion of Ag+ to highly Stable Ag0 nanoparticles by wild type and cell wall deficient strains of Chlamydomonas reinhardtii. Molecules 2018, 24, 98. [Google Scholar] [CrossRef] [Green Version]
  628. Bulgarini, A.; Lampis, S.; Turner, R.J.; Vallini, G. Biomolecular composition of capping layer and stability of biogenic selenium nanoparticles synthesized by five bacterial species. Microb. Biotechnol. 2020, 14, 198–212. [Google Scholar] [CrossRef]
  629. Torres-Chavolla, E.; Ranasinghe, R.J.; Alocilja, E.C. Characterization and functionalization of biogenic gold nanoparticles for biosensing enhancement. IEEE Trans. Nanotechnol. 2010, 9, 533–538. [Google Scholar] [CrossRef]
  630. Pumera, M.; Sánchez, S.; Ichinose, I.; Tang, J. Electrochemical nanobiosensors. Sens. Actuators B Chem. 2007, 123, 1195–1205. [Google Scholar] [CrossRef]
  631. Lopez-Marzo, A.M.; Hoyos-de-la-Torre, R.; Baldrich, E. NaNO3/NaCl oxidant and polyethylene glycol (PEG) capped gold nanoparticles (AuNPs) as a novel green route for AuNPs detection in electrochemical biosensors. Anal. Chem. 2018, 90, 4010–4018. [Google Scholar] [CrossRef]
  632. Ma, E.; Chen, P.; Wilkins, H.M.; Wang, T.; Swerdlow, R.H.; Chen, Q. Pharmacologic ascorbate induces neuroblastoma cell death by hydrogen peroxide mediated DNA damage and reduction in cancer cell glycolysis. Free Radic. Biol. Med. 2017, 113, 36–47. [Google Scholar] [CrossRef] [PubMed]
  633. Eno, C.O.; Zhao, G.; Venkatanarayan, A.; Wang, B.; Flores, E.R.; Li, C. Noxa couples lysosomal membrane permeabilization and apoptosis during oxidative stress. Free Radic. Biol. Med. 2013, 65, 26–37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  634. Geiser, T.; Ishigaki, M.; van Leer, C.; Matthay, M.A.; Broaddus, V.C. H2O2 inhibits alveolar epithelial wound repair in vitro by induction of apoptosis. Am. J. Physiol. Lung Cell. Mol. Physiol. 2004, 287, L448–L453. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  635. Meier, J.; M Hofferber, E.; A Stapleton, J.; Iverson, N.M. Hydrogen peroxide sensors for biomedical applications. Chemosensors 2019, 7, 64. [Google Scholar] [CrossRef] [Green Version]
  636. Sawant, V.J.; Sawant, V.J. Biogenic capped selenium nano rods as naked eye and selective hydrogen peroxide spectrometric sensor. Sens. Bio-Sens. Res. 2020, 27, 100314. [Google Scholar] [CrossRef]
  637. Wang, T.; Yang, L.; Zhang, B.; Liu, J. Extracellular biosynthesis and transformation of selenium nanoparticles and application in H2O2 biosensor. Colloids Surf. B Biointerfaces 2010, 80, 94–102. [Google Scholar] [CrossRef]
  638. Prasad, K.S.; Vaghasiya, J.V.; Soni, S.; Patel, J.; Patel, R.; Kumari, M.; Jasmani, F.; Selvaraj, K. Microbial selenium nanoparticles (SeNPs) and their application as a sensitive hydrogen peroxide biosensor. Appl. Biochem. Biotechnol. 2015, 177, 1386–1393. [Google Scholar] [CrossRef]
  639. Dwivedi, S.; Al-Khedhairy, A.; Ahamed, M.; Musarrat, J. Biomimetic synthesis of selenium nanospheres by bacterial strain JS-11 and its role as a biosensor for nanotoxicity assessment: A novel Se-bioassay. PLoS ONE 2013, 8, e57404. [Google Scholar] [CrossRef]
  640. Cao, H.; Xiao, J.; Liu, H. Enhanced oxidase-like activity of selenium nanoparticles stabilized by chitosan and application in a facile colorimetric assay for mercury (II). Biochem. Eng. J. 2019, 152, 107384. [Google Scholar] [CrossRef]
  641. Zhao, Y.; Sun, Q.; Zhang, X.; Baeyens, J.; Su, H. Self-assembled selenium nanoparticles and their application in the rapid diagnostic detection of small cell lung cancer biomarkers. Soft Matter 2018, 14, 481–489. [Google Scholar] [CrossRef]
  642. Sun, D.; Liu, Y.; Yu, Q.; Qin, X.; Yang, L.; Zhou, Y.; Chen, L.; Liu, J. Inhibition of tumor growth and vasculature and fluorescence imaging using functionalized ruthenium-thiol protected selenium nanoparticles. Biomaterials 2014, 35, 1572–1583. [Google Scholar] [CrossRef] [PubMed]
  643. Huang, J.; Huang, W.; Zhang, Z.; Lin, X.; Lin, H.; Peng, L.; Chen, T. Highly uniform synthesis of selenium nanoparticles with EGFR targeting and tumor microenvironment-responsive ability for simultaneous diagnosis and therapy of nasopharyngeal carcinoma. ACS Appl. Mater. Interfaces 2019, 11, 11177–11193. [Google Scholar] [CrossRef] [PubMed]
  644. Yu, Q.; Liu, Y.; Cao, C.; Le, F.; Qin, X.; Sun, D.; Liu, J. The use of pH-sensitive functional selenium nanoparticles shows enhanced in vivo VEGF-siRNA silencing and fluorescence imaging. Nanoscale 2014, 6, 9279–9292. [Google Scholar] [CrossRef] [PubMed]
  645. Khalid, A.; Tran, P.A.; Norello, R.; Simpson, D.A.; O’Connor, A.J.; Tomljenovic-Hanic, S. Intrinsic fluorescence of selenium nanoparticles for cellular imaging applications. Nanoscale 2016, 8, 3376–3385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  646. Jian, W.; Hui, D.; Lau, D. Nanoengineering in biomedicine: Current development and future perspectives. Nanotechnol. Rev. 2020, 9, 700–715. [Google Scholar] [CrossRef]
  647. Ansari, A.A.; Parchur, A.K.; Thorat, N.D.; Chen, G. New advances in pre-clinical diagnostic imaging perspectives of functionalized upconversion nanoparticle-based nanomedicine. Coord. Chem. Rev. 2021, 440, 213971. [Google Scholar] [CrossRef]
  648. Li, Y.-J.; Wu, J.-Y.; Liu, J.; Xu, W.; Qiu, X.; Huang, S.; Hu, X.-B.; Xiang, D.-X. Artificial exosomes for translational nanomedicine. J. Nanobiotechnol. 2021, 19, 242. [Google Scholar] [CrossRef]
  649. Mignani, S.; Shi, X.; Guidolin, K.; Zheng, G.; Karpus, A.; Majoral, J.-P. Clinical diagonal translation of nanoparticles: Case studies in dendrimer nanomedicine. J. Control. Release 2021, 337, 356–370. [Google Scholar] [CrossRef]
  650. Zhu, G.H.; Gray, A.B.; Patra, H.K. Nanomedicine: Controlling nanoparticle clearance for translational success. Trends Pharmacol. Sci. 2022, 43, 709–711. [Google Scholar] [CrossRef]
  651. Murday, J.S.; Siegel, R.W.; Stein, J.; Wright, J.F. Translational nanomedicine: Status assessment and opportunities. Nanomed. Nanotechnol. Biol. Med. 2009, 5, 251–273. [Google Scholar] [CrossRef]
  652. Riehemann, K.; Schneider, S.W.; Luger, T.A.; Godin, B.; Ferrari, M.; Fuchs, H. Nanomedicine-Challenge and perspectives. Angew. Chem. Int. Ed. 2009, 48, 872–897. [Google Scholar] [CrossRef]
  653. Van Der Meel, R.; Sulheim, E.; Shi, Y.; Kiessling, F.; Mulder, W.J.M.; Lammers, T. Smart cancer nanomedicine. Nat. Nanotechnol. 2019, 14, 1007–1017. [Google Scholar] [CrossRef] [PubMed]
  654. Cabral, H.; Nishiyama, N.; Kataoka, K. Supramolecular nanodevices: From design validation to theranostic nanomedicine. Acc. Chem. Res. 2011, 44, 999–1008. [Google Scholar] [CrossRef] [PubMed]
  655. Hare, J.I.; Lammers, T.; Ashford, M.B.; Puri, S.; Storm, G.; Barry, S.T. Challenges and strategies in anti-cancer nanomedicine development: An industry perspective. Adv. Drug Deliv. Rev. 2017, 108, 25–38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  656. Mirshafiee, V.; Jiang, W.; Sun, B.; Wang, X.; Xia, T. Facilitating translational nanomedicine via predictive safety assessment. Mol. Ther. 2017, 25, 1522–1530. [Google Scholar] [CrossRef] [Green Version]
  657. Li, T.; Xu, H. Selenium-containing nanomaterials for cancer treatment. Cell Rep. Phys. Sci. 2020, 1, 100111. [Google Scholar] [CrossRef]
  658. Mukherjee, A.; Madamsetty, V.S.; Paul, M.K.; Mukherjee, S. Recent advancements of nanomedicine towards antiangiogenic therapy in cancer. Int. J. Mol. Sci. 2020, 21, 455. [Google Scholar] [CrossRef] [Green Version]
  659. Mahmoudi, M.; Bertrand, N.; Zope, H.; Farokhzad, O.C. Emerging understanding of the protein corona at the nano-bio interfaces. Nano Today 2016, 11, 817–832. [Google Scholar] [CrossRef] [Green Version]
  660. Gebauer, J.S.; Malissek, M.; Simon, S.; Knauer, S.K.; Maskos, M.; Stauber, R.H.; Peukert, W.; Treuel, L. Impact of the nanoparticle–protein corona on colloidal stability and protein structure. Langmuir 2012, 28, 9673–9679. [Google Scholar] [CrossRef]
  661. D’Amico, R.S.; Khatri, D.; Reichman, N.; Patel, N.V.; Wong, T.; Fralin, S.R.; Li, M.; Ellis, J.A.; Ortiz, R.; Langer, D.J.; et al. Super selective intra-arterial cerebral infusion of modern chemotherapeutics after blood-brain barrier disruption: Where are we now, and where we are going. J. Neurooncol. 2020, 147, 261–278. [Google Scholar] [CrossRef]
  662. Omidi, Y.; Kianinejad, N.; Kwon, Y.; Omidian, H. Drug delivery and targeting to brain tumors: Considerations for crossing the blood-brain barrier. Expert Rev. Clin. Pharmacol. 2021, 14, 357–381. [Google Scholar] [CrossRef] [PubMed]
  663. Terstappen, G.C.; Meyer, A.H.; Bell, R.D.; Zhang, W. Strategies for delivering therapeutics across the blood–brain barrier. Nat. Rev. Drug Discov. 2021, 20, 362–383. [Google Scholar] [CrossRef] [PubMed]
  664. Han, L.; Jiang, C. Evolution of blood–brain barrier in brain diseases and related systemic nanoscale brain-targeting drug delivery strategies. Acta Pharm. Sin. B 2021, 11, 2306–2325. [Google Scholar] [CrossRef] [PubMed]
  665. Chen, M.; Huang, Y.; Zhu, X.; Hu, X.; Chen, T. Efficient overcoming of blood–brain barrier by functionalized selenium nanoparticles to treat glioma. Adv. Ther. 2018, 1, 1800074. [Google Scholar] [CrossRef]
  666. Yin, T.; Yang, L.; Liu, Y.; Zhou, X.; Sun, J.; Liu, J. Sialic acid (SA)-modified selenium nanoparticles coated with a high blood–brain barrier permeability peptide-B6 peptide for potential use in Alzheimer’s disease. Acta Biomater. 2015, 25, 172–183. [Google Scholar] [CrossRef]
  667. Karimzadeh, K.; Sharifi, E.; Bakhshi, N.; Ramzanpoor, M. Biogenic silver nanoparticles using Oxalis corniculata characterization and their clinical implications. J. Drug Deliv. Sci. Technol. 2019, 54, 101263. [Google Scholar] [CrossRef]
  668. Khandel, P.; Yadaw, R.K.; Soni, D.K.; Kanwar, L.; Shahi, S.K. Biogenesis of metal nanoparticles and their pharmacological applications: Present status and application prospects. J. Nanostruct. Chem. 2018, 8, 217–254. [Google Scholar] [CrossRef] [Green Version]
  669. Hosny, M.; Fawzy, M. Instantaneous phytosynthesis of gold nanoparticles via Persicaria salicifolia leaf extract, and their medical applications. Adv. Powder Technol. 2021, 32, 2891–2904. [Google Scholar] [CrossRef]
  670. Marouzi, S.; Sabouri, Z.; Darroudi, M. Greener synthesis and medical applications of metal oxide nanoparticles. Ceram. Int. 2021, 47, 19632–19650. [Google Scholar] [CrossRef]
  671. Vijayakumar, S.; Divya, M.; Vaseeharan, B.; Ranjan, S.; Kalaiselvi, V.; Dasgupta, N.; Chen, J.; Durán-Lara, E.F. Biogenic preparation and characterization of ZnO nanoparticles from natural polysaccharide Azadirachta indica L. (neem gum) and its clinical implications. J. Cluster Sci. 2021, 32, 983–993. [Google Scholar] [CrossRef]
  672. Fierascu, I.; Fierascu, I.C.; Brazdis, R.I.; Baroi, A.M.; Fistos, T.; Fierascu, R.C. Phytosynthesized metallic nanoparticles-between nanomedicine and toxicology. A brief review of 2019’s findings. Materials 2020, 13, 574. [Google Scholar] [CrossRef] [PubMed]
  673. Ovais, M.; Khalil, A.T.; Raza, A.; Islam, N.U.; Ayaz, M.; Saravanan, M.; Ali, M.; Ahmad, I.; Shahid, M.; Shinwari, Z.K. Multifunctional theranostic applications of biocompatible green-synthesized colloidal nanoparticles. Appl. Microbiol. Biotechnol. 2018, 102, 4393–4408. [Google Scholar] [CrossRef] [PubMed]
  674. Micke, O.; Schomburg, L.; Buentzel, J.; Kisters, K.; Muecke, R. Selenium in oncology: From chemistry to clinics. Molecules 2009, 14, 3975–3988. [Google Scholar] [CrossRef]
  675. Gabizon, A.A.; De Rosales, R.T.M.; La-Beck, N.M. Translational considerations in nanomedicine: The oncology perspective. Adv. Drug Deliv. Rev. 2020, 158, 140–157. [Google Scholar] [CrossRef]
  676. Noah, N.M.; Ndangili, P.M. Current trends of nnobiosensors for point-of-care diagnostics. J. Anal. Method Chem. 2019, 2019, 2179718. [Google Scholar] [CrossRef] [PubMed]
  677. Bellah, M.; Christensen, S.M.; Iqbal, S. Nanostructures for medical diagnostics. J. Nanomater. 2012, 2012, 486301. [Google Scholar] [CrossRef]
  678. Bregoli, L.; Movia, D.; Gavigan-Imedio, J.D.; Lysaght, J.; Reynolds, J.; Prina-Mello, A. Nanomedicine applied to translational oncology: A future perspective on cancer treatment. Nanomed. Nanotechnol. Biol. Med. 2016, 12, 81–103. [Google Scholar] [CrossRef] [Green Version]
  679. Danaei, M.; Dehghankhold, M.; Ataei, S.; Hasanzadeh Davarani, F.; Javanmard, R.; Dokhani, A.; Khorasani, S.; Mozafari, M.R. Impact of particle size and polydispersity index on the clinical applications of lipidic nanocarrier systems. Pharmaceutics 2018, 10, 57. [Google Scholar] [CrossRef] [Green Version]
  680. Tinkle, S.; McNeil, S.E.; Mühlebach, S.; Bawa, R.; Borchard, G.; Barenholz, Y.; Tamarkin, L.; Desai, N. Nanomedicines: Addressing the scientific and regulatory gap. Ann. N. Y. Acad. Sci. 2014, 1313, 35–56. [Google Scholar] [CrossRef]
  681. Mundekkad, D.; Cho, W.C. Nanoparticles in clinical translation for cancer therapy. Int. J. Mol. Sci. 2022, 23, 1685. [Google Scholar] [CrossRef]
Figure 1. Main advantages of nanoparticles in biomedicine. Explored nanoplatforms in the biomedical field might be organic, inorganic, or hybrid. Inorganic NPs might be of different compositions, such as core-shell quantum dots and passivated inorganic NPs by an organic/polymeric layer, and of various shapes, such as spheres and nanotubes. Liposomes, polymeric NPs, and biopolymer-based NPs are some examples of organic NPs.
Figure 1. Main advantages of nanoparticles in biomedicine. Explored nanoplatforms in the biomedical field might be organic, inorganic, or hybrid. Inorganic NPs might be of different compositions, such as core-shell quantum dots and passivated inorganic NPs by an organic/polymeric layer, and of various shapes, such as spheres and nanotubes. Liposomes, polymeric NPs, and biopolymer-based NPs are some examples of organic NPs.
Nanomaterials 13 00424 g001
Figure 2. Selenium toxicity in humans and animals: cytotoxicity, genotoxicity, and epidemiology.
Figure 2. Selenium toxicity in humans and animals: cytotoxicity, genotoxicity, and epidemiology.
Nanomaterials 13 00424 g002
Figure 3. Main biomedical properties and applications of SeNPs.
Figure 3. Main biomedical properties and applications of SeNPs.
Nanomaterials 13 00424 g003
Figure 4. Structure of SeNPs@GM1/TMP and their protective activity against t-BOOH induced neuron cell death. Reproduced with permission from Ref. [264]. 2019, the Royal Society of Chemistry.
Figure 4. Structure of SeNPs@GM1/TMP and their protective activity against t-BOOH induced neuron cell death. Reproduced with permission from Ref. [264]. 2019, the Royal Society of Chemistry.
Nanomaterials 13 00424 g004
Figure 5. The mechanism of ROS generation mediated by SeNPs. Adapted with permission from Ref. [174]. 2021, The Royal Society of Chemistry.
Figure 5. The mechanism of ROS generation mediated by SeNPs. Adapted with permission from Ref. [174]. 2021, The Royal Society of Chemistry.
Nanomaterials 13 00424 g005
Figure 6. Biomedical application of SeNPs against viral infections.
Figure 6. Biomedical application of SeNPs against viral infections.
Nanomaterials 13 00424 g006
Figure 7. Schematic representation of the different manners through which NPs enter the human body and are internalized inside the cells.
Figure 7. Schematic representation of the different manners through which NPs enter the human body and are internalized inside the cells.
Nanomaterials 13 00424 g007
Figure 8. Schematic representation of multifunctional NPs that may include an imaging component, a targeting element, and a therapeutic constituent.
Figure 8. Schematic representation of multifunctional NPs that may include an imaging component, a targeting element, and a therapeutic constituent.
Nanomaterials 13 00424 g008
Figure 9. Confocal laser scanning microscopy images of antibiofilm activity of biogenic SeNPs, produced using the aqueous extract of Murraya koenigii against Gram-positive (E. faecalis and S. mutans) and Gram-negative (S. sonnei and P. aeruginosa) bacteria. Reproduced with permission from Ref. [109]. 2019, Elsevier.
Figure 9. Confocal laser scanning microscopy images of antibiofilm activity of biogenic SeNPs, produced using the aqueous extract of Murraya koenigii against Gram-positive (E. faecalis and S. mutans) and Gram-negative (S. sonnei and P. aeruginosa) bacteria. Reproduced with permission from Ref. [109]. 2019, Elsevier.
Nanomaterials 13 00424 g009
Figure 10. Anticancer effect of biogenic, spherical SeNPs, synthesized by a living culture of B. lichenisformis, against PC3 cells via a necroptosis pathway. Adapted with permission from Ref. [295] under the terms of the Creative Commons CC BY license. 2017, Springer Nature.
Figure 10. Anticancer effect of biogenic, spherical SeNPs, synthesized by a living culture of B. lichenisformis, against PC3 cells via a necroptosis pathway. Adapted with permission from Ref. [295] under the terms of the Creative Commons CC BY license. 2017, Springer Nature.
Nanomaterials 13 00424 g010
Figure 11. Measurements of (A) ROS generation and (C) MMP induced by various concentrations (10–40 µg·mL−1) of SeNPs using flow cytometry; (B,D) microscopy illustration of ROS and MMP production, respectively. Reproduced with permission from Ref. [285]. 2020, Elsevier.
Figure 11. Measurements of (A) ROS generation and (C) MMP induced by various concentrations (10–40 µg·mL−1) of SeNPs using flow cytometry; (B,D) microscopy illustration of ROS and MMP production, respectively. Reproduced with permission from Ref. [285]. 2020, Elsevier.
Nanomaterials 13 00424 g011
Figure 12. Stakeholders (academic researchers, clinicians and patients, industry, regulatory bodies) and the main challenges faced in the translation of nanomedicine.
Figure 12. Stakeholders (academic researchers, clinicians and patients, industry, regulatory bodies) and the main challenges faced in the translation of nanomedicine.
Nanomaterials 13 00424 g012
Table 1. Classification of selenium compounds based on structural features. Adapted from Ref. [79] under Creative Commons Attribution license. 2012, MDPI AG.
Table 1. Classification of selenium compounds based on structural features. Adapted from Ref. [79] under Creative Commons Attribution license. 2012, MDPI AG.
Selenium CompoundType
Selenoamino acidsSelenomethionine (SeMet)
Methyl selenocysteine (MeSeCys)
Selenocysteine (SeCys)
Selenocystamine
Se-heterocyclic compounds1,3-Selenazolin-4-one derivatives
2-Phenyl-1,2-benzisoselenazol-3(2H)-one (ebselen)
2,5-Bis(5-hydroxymethyl-2-selenienyl)-3-hydroxymethyl-N-methylpyrrole (D-501036)
1,2-[Bis(1,2-benzisoselenazolone-3-(2H)-ketone)] ethane (BBSKE)
2-Cyclohexylselenazolidine-4-(R)-carboxylic acid (ChSCA)
2-Buthylselenazolidine-4-(R)-carboxylic acid (BSCA)
SelenocyanatesIsatin analogs
Diphenylmethylselenocyanate
1-4-Phenylenebis(methylene) selenocyanate (p-XSC)
Temozolomide (TMZ)-Se
5-Phenylcarbamoylpentyl selenocyanide (SelSA-2)
SelenidesMethylimidoselenocarbamates
5-Phenylselenyl-methyl-2′-deoxyuridine (PhSe-T)
5-Methylselenyl-methyl-2′-deoxyuridine (MeSe-T)
β-Selenium amine derivatives
Se,Se’-1,4-phenylenebis(1,2-ethanediyl) bisisoselenourea (PBISe)
Di-selenidesBis(4-aminophenyl) diselenide
Bis(5-phenylcarbamoylpentyl) diselenide (SelSA-1)
Diselenodipropio nic acid (DSePA)
2-Selenium-bridged β-cyclodextrin (2-SeCD)
Se(IV) compoundsSodium selenite (Na2SeO3)
Selenous acid (H2SeO3)
Methyl selenic acid (MeSeA)
Selenium dioxide (SeO2)
Table 2. Main types of selenoproteins and their functions.
Table 2. Main types of selenoproteins and their functions.
SelenoproteinsAbbreviationLocalizationFunctionRef.
Glutathione peroxidaseGPX Protection against oxidative stress. Catalytic reduction of H2O2[75,246]
Cytosolic GPX1GPX1Cytoplasm, ubiquitousAntioxidative defense
Extracellular GPXGPX3Plasma and thyroid follicleAnti-inflammatory activity
Phospholipid GPXGPX4Mitochondrial membraneReduction of the phospholipid hydroperoxides. Membrane antioxidant
Thioredoxin reductase TrxR Oxidoreductase activity with NADPH as the cofactor [244]
Cytosolic TrxRTrxR1Mainly cytosolic, ubiquitousMain antioxidant “weapon” at the cellular level. Inhibition of apoptosis and redox state of transcription factors
Mitochondrial TrxRTrxR2Mitochondrial, ubiquitousRegulation of cell proliferation
Mitochondrial TrxRTrxR3Mainly mitochondrial, ubiquitousRegulation of apoptosis and signaling pathway
Iodothyronine deiodinaseDIO Catalytic conversion of T4 and T3[247]
Type I DIODIO1Liver, lung, eyes, kidney, thyroid,
pituitary, CNS *
Conversion of T4 into T3, T4 into rT3, and of T3 into T2
Type II DIODIO2 Local (intracellular) production of T3 from T4 and T2 from rT3
Type III DIODIO3Placenta, fetus, liver, gravid uterus, fetal
and neonatal brain, skin
Production of T2 from T4 and rT3 from T4
Selenoprotein PSePPThyroid and bloodSelenium transport and storage, endothelial antioxidant[248]
* NADPH: nicotinamide adenine dinucleotide phosphate (reduced form of the redox coenzyme nicotinamide adenine dinucleotide phosphate); CNS: Central nervous system; T2: diiodothyronine; T3: triiodothyronine; rT3: reverse T3; T4: thyroxine.
Table 3. Antioxidant activity of biogenic SeNPs: Biosynthetic route, NP features (size and shape), methods used and main outcomes.
Table 3. Antioxidant activity of biogenic SeNPs: Biosynthetic route, NP features (size and shape), methods used and main outcomes.
Biological System Used for SynthesisShape and Size of SeNPsAntioxidant Measurement TechniqueAntioxidant ActivityIC50 (μg/mL)EC50 (μg/mL)Ref.
Bacillus sp.
MSh-1.
Spherical; 80–220 nmDPPH and reducing power assaysRSA of 23.1 ± 3.4%
Dose-dependent reducing power within a 0–200 μg·mL−1 concentration range
41.5 ± 0.9N/A[260]
Saccharomyces cerevisiaeSpherical; 50 nmDPPHRSA increase of 21.7–48.5% in a dose-dependent mannerN/AN/A[150]
Bacillus sp.
EKT1
Spherical; 31–335 nmDPPHRSA of up to 56.5 ± 5.0% at 400 μg·mL−1322.8N/A[493]
Quercetin and gallic acidBimetallic Ag-Se NPs capped by flavonoids and phenolics; 30–35 nmABTS, DPPH and MTT assays59–62% T-AOC30–66N/A[495]
Aqueous chitosan microspheres Spherical; 36–95 nmH2O2 levels; measurement of GSH, TBARS (MDA equivalent), GSH-Px, SOD and CATIncrease in both intracorporeal Se retention and levels of GSH-Px, SOD and CAT; reduced levels of TBARSN/AN/A[129]
Emblica officinalis extractSpherical; 15–40 nmDPPH and ABTS radical scavenging assaysDose-dependent RSA, linear relationship with NP concentration127.28 ± 3.73DPPH: 15.67 ± 1.41 mg·mL−1
ABTS: 18.84 ± 1.02 mg·mL−1
[155]
Lactobacillus casei ATCC 39350–80 nmCellular methods:
T-AOC
MDA
T-SOD
GSH-Px levels
TrxR
Increased T-AOC, T-SOD, TrxR and GSH-Px
Reduced MDA levels in serum and jejunum
N/AN/A[496]
Pantoea
agglomerans
Spherical; 30–300 nmProduction of ROS using HUVEC
fluorescence determined using a microplate reader at 485-nm excitation and 583-nm emission
Decrease in fluorescence resulting from the oxidation of the intracellular probe dichlorofluorescein (DCF)N/AN/A[262]
L. casei ATCC 393Spherical; capped with proteins and polysaccharides; 50–80 nmH2O2 levelsIncreased GPX activity, reduced MDAN/AN/A[497]
H2O2-induced oxidative damage model of human colon mucosal epithelial cellsAlleviated increase in ROS, reduced ATP and MMP
Improved levels of Nrf2, HO-1, and NQO-1 proteins
N/AN/A[498]
L. lactis NZ9000Spherical; 38–152 nmH2O2 levels measurement;
MDA
T-SOD
GPX
Alleviated IPEC-J2 cell oxidative
damage caused by H2O2
Inhibition of intracellular ROS production
N/AN/A[499]
Cordyceps sinensis EPS conjugationAmorphous and monoclinic; 80–125 nm.ABTS and superoxide anion radical (O2 ) scavenging assaysSmaller SeNPs present high O2 scavenging ability.
Se/P ratios (1:3, 1:1 and 4:3) had a higher ABTS•+ scavenging ability, and could reach 88.89%, 85.53% and 69.88%, respectively, at 0.2 mg·mL−1.
N/AN/A[148]
Ephedra aphylla extractSpherical and tetragonal; 13.95–26.26 nmDPPH assayLower activity of SeNPs a than the plant extract0.213 and 0.296 mg·mL−1N/A[491]
Green tea extract and Lycium barbarum polysaccharidesSpherical and triangular; 83–160 nmDPPH and ABTS assaysStrong, concentration-dependent DPPH-scavenging activity
at 5–25 µM
High antioxidant activity with low EC50
Dose-dependent inhibition of ABTS free radicals
N/A22 µM[500]
Theobroma cacao L. bean shell extractSpherical; 1–3 nmABTS and FRAP assaysABTS: 28.6 ± 0.1 mg TE/g
FRAP: 12.4 ± 0.2 mg TE/g
N/AN/A[487]
ChitosanSpherical; 102–104 nmDPPH, ABTS and superoxide anion radical (O2 ) scavenging assaysDPPH: 83.06% at 0.5 mM
ABTS: 74.33, 80.23 and 81.99% at 2 mM
Superoxide: 25.20, 27.54, 31.44% at 1 mM
DPPH: 0.296, 0.306, 0.325, 0.370 mM
ABTS: 1.314, 1.249, 1.143 and 1.101 mM
N/A[489]
Streptomyces minutiscleroticusSpherical; 100–250 nmDPPH;
Reducing power assay;
T-AOC: Phosphomolybdenum method
All measurements increase in a dose-dependent manner
T-AOC was more or less equal to the standard ascorbic acid
N/AN/A[316]
Withania somniferaSpherical; 45–90 nmDPPHRSA increase in a dose-dependent manner in the range of 20–100 mg·mL−1 *14.81 μg·mg−1N/A[160]
Diospyros montana leaf extractSpherical; 4–16 nmDPPH and FRAP assaysDPPH: color change from purple to pale yellow. RSA of 61.12% at 200 μg·mL−1
FRAP: color change from yellow to shades of green and blue
0.225 0.435 [490]
Corbicula flumineaSpherical; 40–70 nmDPPH, TEAC and FRAP of plasma assays.DPPH RSA: 70, 77, 83, 79, and 53% at 1.5 mg·mL−1 *. Increase in a dose-dependent manner
TEAC: highest RSA at 226 μmol Trolox/g sample
FRAP: highest RSA at 150 μmol Fe2+/g sample
1.5 mg·mL−1 *N/A[488]
Murraya koenigiiSpherical; 50–150 nmDPPH and
Superoxide anion (O2 ) scavenging assay
Concentration-dependent RSA increase25 and 50N/A[109]
Ginger plant (Zingiber officinale) extractSpherical; 100–150 nmDPPHDose-dependent RSA increase (disappearance of the purple color)
SeNPs are free radical inhibitors or scavengers acting possibly as primary antioxidants
125N/A[501]
* We think these concentrations are in µg·mL−1. Abbreviations: ABTS: (2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid); APX: Ascorbate peroxidase; CAT: Catalase; DAO: Diamine oxidase assay; DCF: Dichlorofluorescein; DPPH: 2,2-diphenyl-1-picrylhydrazyl; ABTS: 2,20-azino-bis (3-ethylbenzothiazoline-6-sulphonic acid); EC50: Effective concentration required to inhibit 50% of free radicals; EPS: Exopolysaccharides; FRAP: Ferric reducing antioxidant power; GSH: Glutathione; GSH-Px: Glutathione peroxidase; HUVEC: Human umbilical vein endothelial cells; IC50: Half maximal inhibitory concentration; IPEC-J2: Intestinal porcine enterocytes; MDA: Malondialdehyde assay; MMP: Mitochondrial membrane potential; MTT: 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide; POX: Peroxidase; ROS: Reactive oxygen species; RSA: Radical scavenging activity; SOD: Superoxide dismutase; T-AOC: Total antioxidant capacity; TBARS: Thiobarbituric acid reactive substances; TEA: Trolox equivalent antioxidant; TEAC: Trolox equivalent antioxidant capacity; TrxR: Thioredoxin reductase; T-SOD: Total superoxide dismutase.
Table 4. Anticancer activity of several biogenic SeNPs against different cancer cell lines.
Table 4. Anticancer activity of several biogenic SeNPs against different cancer cell lines.
Biological System/Green MethodShape and Size (nm)Concentration/DosageAssay/PathwayCell Line *Key OutcomesRef.
Streptomyces bikiniensisNanorods; 17 nm10, 25, 50, and 100 μg·mL−1MTT dye reduction assayHep-G2 (M)
MCF-7 (M)
ID50%: 75.96 and 61.86 μg·mL−1, respectively.
Loss of cell-to-cell contact, cell shrinkage, and formation of apoptotic bodies.
Higher reduction of cell viability in Hep-G2 (42.3–86.9%) than in MCF-7 (37.5–69.1%).
[521]
Cassia auriculata leaf extract10–20 nm0.5 to 150 μg·mL −1MTT assayHL-60 (M)
Vero cell line (NM)
Antileukemia activity in a dose-dependent manner with a CC50: 7.01 μg·mL−1 (HL-60) and 109.13 μg·mL−1 (Vero cells)[162]
Garlic (Allium sativum) clove extractSpherical; 40–100 nm15, 30, 60 and
90 μg·mL−1
MTT assayVero cell line (NM)CC50 of 31.8 ± 0.6 μg·mL−1[522]
Chitosan (CTS-SeNPs) and Pleurotus ostreatus fermented fenugreek (SeNPs-AEFFP)Spherical; CTS-SeNPs: 45 nm
SeNPs-AEFFP: 11.8 nm
CTS-SeNPs: 1.187–38 μg·mL−1
SeNPs-AEFFP: 0.594–19 μg·mL−1
NPs exposed to γ-ray doses of 60 and 15 kGy, respectively, against EaC and CaCo-2.
Trypan blue (0.5%) assayEaC (M)
CaCo-2 (M)
For EaC:
CTS-SeNPs: IC50 = 23.12%
SeNPs-AEFFP: IC50 = 7.21%
For CaCo-2:
CTS-SeNPs: IC50 = 25.32%
SeNPs-AEFFP: IC50 = 8.57%
SeNPs exhibit a concentration-dependent repression against EaC and CaCo-2 and a selective cytotoxic effect.
[112]
E. coliSpherical, elliptical and nanorods; 60 nm20, 60, and 100 μg·mL−1MTT and addition of DMSO. Caspase 3-involved apoptosis pathway.A549 (M)
IMR-90 (NM)
Cell viability of ∼70%, ∼45% and ∼25%, high ROS generation and elevated caspase-3 activity.[523]
B. licheniformis JS2Spherical; 110 nm1, 2, 4, 6, 50, or
200 μg Se/mL for 24 h
Colorimetric XTT assay, activation of caspases 3 and 7, DMSO treatment and hemolysis assaysLNCaP-FGC (M)Overexpression of TNF and IRF1, reducing the expression of androgen receptors.
SeNPs decrease the cell viability regardless of apoptosis and necrosis. SeNPs induce cell death through neither apoptosis nor necrosis.
[524]
S. minutiscleroticus M10A62Spherical; 10–250 nm50–100 μg·mL−1MTTHep-G2 (M)
HeLa (NM)
50 μg·mL−1 concentration of SeNPs was required for 99.5% HepG2 growth inhibition and 100 μg·mL−1 for 100% growth inhibition of HeLa.[316]
B. licheniformis JS2110 nmMinimum of 2 μg Se/mLReal-time qPCR analysis; confocal microscopy; treatment with cytochalasin DPC-3 (M)ROS mediated necroptosis of PC-3 cells independent of RIP3 and MLKL and regulated by a RIP1 kinase. Increased expression of necroptosis associated to TNF and IRF1[295]
Fenugreek seed extractAmorphous; 50–150 nm25, 50, 75, and 100 μg·mL−1 for 24 hMTT assayMCF-7 (M)SeNPs augment the cytotoxicity of doxorubicin and induce MCF 7 cell death through apoptosis.[525]
Idiomarina sp. PR58-8Spherical; 150–350 nm5–100 μg·mL−1 for 24 hMTT assay,
ROS assay, apoptotic index assay, Western blot assay.
HaCaT (NM)
HeLa cells (M)
Caspase-dependent apoptosis in HeLa cell lines: decrease in expression of pro-caspase 3. SeNPs exhibited dose-dependent cytotoxicity with only 3% viability at 100 μg·mL−1.[526]
Acinetobacter sp. SW30Nanospheres and crystalline nanorods of 78 nm.
Polygonal-shaped SeNPs of 79 nm.
0–100 μg·mL−1MTT assay4T1 (M)
MCF-7 (M)
NIH/3T3 (NM)
HEK293 (NM)
Antiproliferative activity.[527]
Bacillus
sp. MSh-1
Spherical; 80–220 nm10, 20, 50 and 100 μg·mL−1MTT assay and gelatin zymographyHT-1080 (NM)SeNP dose of 100 μg·mL−1 decreases the viability of the cell line to 50%, whereas a lower dose (10 μg·mL−1) shows a low level of cytotoxicity with a viability of more than 80%.[528]
Asteriscus graveolens extractSpherical; 20 nm25–125 mg·mL–1 ‡ for 24 hMTT assay, flow cytometry analysis, measurement of ROS (conversion of DCFH-DA to DCFH); measurement of MMP and lipid peroxidationHepG2 (M)Cell viability (IC50): 51.8% at 3.98 μg·mL−1. SeNPs inhibit the growth of HepG2 cells mainly by induction of apoptosis. They also significantly and rapidly increase the ROS level.[285]
Moringa oleifera extractSpherical; 23–35 nm
Polygonal; 25–45 nm
N/AMTT assayCaCo-2 (M)
HepG2 (M)
MCF-7 (M)
IC50: 50.3% at 392.57 μg·mL−1[156]
Penicillium corylophilumSpherical; 29.1–48.9 nm1000, 500, 250, 125, 62.5 and 31.25 ppm incubated in 5%
CO2 at 37 °C for 24 h
MTT assayWI-38 (NM)
CaCo-2 (M)
IC50: 171.8 ppm (Wi 38) and 104.3 ppm (CaCo-2)[502]
Hawthorn fruit extract113 nm0, 5, 10 and
20 μg·mL−1 for 24 h
MTT assay; flow cytometric analysis; ROS detection; MMP measurement; Western blot assayHepG2 (M)IC50: 19.22 ± 5.3 μg·mL−1[529]
L. casei 393Spherical; 50–80 nm4, 8, and 16 μg·mL−1 for 12 hRT-PCR; mRNA expression levels of Bax, caspase 3,
p53 and bcl-2.
HepG2 (M)
NCM460 (NM)
Endocytosis of SeNPs induces cell death by reducing the viability of HepG2, increasing mRNA levels of caspase 3, Bax, and p53, and
reducing mRNA expression of bcl-2.
[497]
Carica papaya latexSpherical; 70 nm5, 10, 15, 20, 25, 30, 35, 40, 45, and 50 μg·mL−1 for 48 hMTT assayHBL100 (NM)
MDA-MB-231 (M)
IC50 (HBL100): 50 μg·mL−1
IC50 (MDA-MB-231): 34 μg·mL−1
[530]
Spermacoce hispida aqueous leaf extract (Sh-SeNPs)
+
S-allyl glutathione conjugation (SAG-Sh-SeNPs)
Sh-SeNPs: aggregation

SAG-Sh-SeNPs: spherical; 50 nm
HepG2: 1.88, 3.75,
7.50, 15.00 and 30.00 μg·mL−1 for 24 h
Vero cells: 3.7–60.0 μg·mL−1 for 48 h
MTT assay; determination of intracellular ROS production and MMP; cell cycle analysis by flow cytometry; DNA fragmentation assay; apoptosis determination by
acridine orange/ethidium bromide staining
HepG2 (M)
Vero cells (NM)
IC50 (Sh-SeNPs): 30.0 μg·mL−1
IC50 (SAG-Sh-SeNPs): 18.7 μg·mL−1
SAG-Sh-SeNPs induce cell cycle arrest at sub-G1 phase and further lead to apoptosis. The NPs increase ROS levels, disrupt MMP, initiate DNA fragmentation and decrease the endogenous levels of antioxidants, such as GSH, superoxide dismutase, catalase and GSH peroxidase.
[531]
Undaria pinnatifida polysaccharidesSpherical; 44–92 nm (average of 59 nm)N/AMTT assay; flow cytometry analysis; Annexin-V-FLUOS staining assay; measurement of ROS levels; MMP evaluationA375 (M)
CNE2 (M)
Hep G2 (M)
MCF-7 (M)
IC50 values ranging from 3.0 to 14.1 μM. Apoptosis with the involvement of oxidative stress and mitochondrial dysfunction.[532]
Ceropegia bulbosa tuber’s aqueous extractsSpherical; 55.9 nm0, 5, 10.0, 15.0, 20.0, 25.0, 30.0, 35.0, 40.0, 45.0, 50.0 μg·mL−1MTT assayMDA-MB-231 (M)
HBL-100 (M)
IC50 (MDA-MB-231): 34 μg·mL−1 for 48 h.
IC50 (HBL-100): more than 50 μg·mL−1 for 48 h.
[506]
Diospyros montana leaf extract4–16 nm50, 150, 250, 350 μg·mL−1MTT assayMCF-7 (M)IC50: 80.83 μg·mL−1, SeNPs enhance the cytotoxicity.[490]
Chitosan decorationSpherical; 50 nm10, 50,
100 μM
WST-1 assayHepG2 (M)SeNPs decrease the cell viability to 76.63, 63.31 and 56.34% and inhibit the growth of HepG2 cells in a time- and dose-dependent manner.[489]
Ephedra aphylla extractSpherical and tetragonal; 13.95–26.26 nmN/AMTT assayHepG-2 (M)
MCF-7 (M)
HCT-116 (M)
HeLa (M)
PC3 (M)
HEp2 (M)
IC50 (HePG-2): 7.56 ± 0.60 μg·mL−1
IC50 (MCF-7): 15.65 ± 1.40 μg·mL−1
IC50 (HCT-116): 10.02 ± 0.90 μg·mL−1
IC50 (HeLa): 9.23 ± 0.80 μg·mL−1
IC50 (PC3): 18.63 ± 1.50 μg·mL−1
IC50 (HEp2): 12.10 ± 1.20 μg·mL−1
[491]
Lentinan (LNT, denatured β-glucan)Spherical; 28 nmN/AMTT assayHeLa (M)IC50 of three complexes Se/s-LNT-1, Se/s-LNT-2, Se/s-LNT-3 were estimated to be 85, 37, and 19 μM. SeNPs with small, uniform size greatly enhanced the antitumor activity and bioavailability.[533]
Quercetin and gallic acidBimetallic Ag-Se NPs of 30–35 nm and capped by flavonoids and phenolics.50, 100, 250 and 500 μg·mL−1MTT assayDL (M)The viability of DL cells was 20% at 50 μg·mL−1 of Ag-SeNPs, while at 100 μg·mL−1, it was reduced to 15%. The Ag-SeNPs showed strong anticancer activity at a lower concentration.[495]
Cationic pullulan (CP)Spherical and microflowers; 50 nmN/AMTT assay;
Annexin-V-FITC and propidium iodide (PI) staining
L929 (M)
KB (M)
IC50: 0.060 μM. The early- and late-stage apoptotic rates of KB cells treated with doxorubicin only reached 0.52% and 4.64%, respectively, and the highest induction of 55.8% was arrested in the necrosis rate.[534]
Walnut peptidesSpherical; 89.22 nm200 μL·mL−1MTT assay; POM, flow cytometry; MMP assay; nuclear morphology analysis by Hoechst 33258; measurement of ROS production; DNA fragmentation assay; caspase activity assay; Western blot assayHL-7702 (L02) (M)
MCF-7 (M)
SGC-7901 (M)
A549 (M)
PC3 (M)
HeLa (M)
MCF-7 cells were the most sensitive to SeNPs. The apoptosis-inducing activity was proved by the accumulation of S-phase cell arrest, nuclear condensation, and DNA breakage.
The intrinsic signal pathway was through the activation of FADD and caspases 3, 8, and 9, in combination with the MMP depletion and ROS generation.
[535]
* Cell lines: 4T1: murine mammary carcinoma cell line; A375: human melanoma cell line; A549: human lung carcinoma cells; CaCo-2: human colorectal adenocarcinoma; CNE-2: human nasopharyngeal carcinoma cell line; DL: Dalton’s lymphoma (mice); EAC: Ehrlich ascites carcinoma (mice); HaCaT: normal human epidermal keratinocyte cell line; HBL100: human epithelial breast cell line (non-malignant); HCT-116: human colorectal carcinoma cell line; HDF: primary human dermal fibroblasts (non-malignant); HEK293: human embryonic kidney cell line (non-malignant); HeLa: human cervical carcinoma cell lines; HEp2: human epithelial carcinoma cell line; Hep-G2: human hepatic carcinoma cell line; HL-60: human leukemia cells; HL-7702 (L02): human liver cell line (non-malignant); HT-1080: human fibrosarcoma cell line; IMR-90: normal lung fibroblast cells (non-malignant); KB: human cervical carcinoma cell line; L292: fibroblast cell line (mice, non-malignant); LNCaP-FGC: human prostate epithelial carcinoma cell line; MCF-7: human breast adenocarcinoma cells; MDA-MB-231: human breast adenocarcinoma cell line; NCM460: human colon mucosal epithelial cell line (non-malignant); NIH/3T3: murine embryonic fibroblasts (non-malignant); PC3: human prostate adenocarcinoma cell line; PC-3: human prostate adenocarcinoma cell line; SGC-7901: human gastric cancer cell line, Vero cell lines (monkey, kidney, non-malignant); WI-38: human embryonic lung cell line (non-malignant); M: malignant; NM: non-malignant. Abbreviations: CC50: Cytotoxicity concentration; DCFH: 2′,7′-Dichlorodihydrofluorescein; DCFH-DA: 2′,7′–dichlorodihydrofluorescin diacetate; DMSO: Dimethylsulfoxide; DNA: Deoxyribonucleic acid; FADD: Fas associated via death domain; FITC: Fluorescein isothiocyanate; IC50: Half maximal inhibitory concentration; IRF: Interferon regulatory factor; MLKL: Mixed lineage kinase domain such as pseudokinase; MMP: Matrix metalloproteinases; mRNA: Messenger ribonucleic acid; MTT: 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide; POM: Polarized optical microscopy; RIP: Receptor interacting serine/threonine kinase; ROS: Reactive oxygen species; RT-(q)PCR: Real-time (quantitative) polymerase chain reaction; TNF: Tumor necrosis factor; WST-1: 4-[3-(4-Iodophenyl)-2-(4-nitro-phenyl)-2H-5-tetrazolio]-1,3-benzene sulfonate; XTT: sodium 3′-[1-(phenylaminocarbonyl)-3,4-tetrazolium]-bis (4-methoxy6-nitro) benzene sulfonic acid hydrate. We think the concentration is in µg·mL−1.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zambonino, M.C.; Quizhpe, E.M.; Mouheb, L.; Rahman, A.; Agathos, S.N.; Dahoumane, S.A. Biogenic Selenium Nanoparticles in Biomedical Sciences: Properties, Current Trends, Novel Opportunities and Emerging Challenges in Theranostic Nanomedicine. Nanomaterials 2023, 13, 424. https://doi.org/10.3390/nano13030424

AMA Style

Zambonino MC, Quizhpe EM, Mouheb L, Rahman A, Agathos SN, Dahoumane SA. Biogenic Selenium Nanoparticles in Biomedical Sciences: Properties, Current Trends, Novel Opportunities and Emerging Challenges in Theranostic Nanomedicine. Nanomaterials. 2023; 13(3):424. https://doi.org/10.3390/nano13030424

Chicago/Turabian Style

Zambonino, Marjorie C., Ernesto Mateo Quizhpe, Lynda Mouheb, Ashiqur Rahman, Spiros N. Agathos, and Si Amar Dahoumane. 2023. "Biogenic Selenium Nanoparticles in Biomedical Sciences: Properties, Current Trends, Novel Opportunities and Emerging Challenges in Theranostic Nanomedicine" Nanomaterials 13, no. 3: 424. https://doi.org/10.3390/nano13030424

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