Next Article in Journal
On the Degree of Plastic Strain during Laser Shock Peening of Ti-6Al-4V
Next Article in Special Issue
Preparation and Characterization of an Antibrowning Nanosized Ag-CeO2 Composite with Synergistic Antibacterial Ability
Previous Article in Journal
Enhanced Codelivery of Gefitinib and Azacitidine for Treatment of Metastatic-Resistant Lung Cancer Using Biodegradable Lipid Nanoparticles
Previous Article in Special Issue
Laser-Induced Transferred Antibacterial Nanoparticles for Mixed-Species Bacteria Biofilm Inactivation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of the Antibacterial, Fungicidal and Antiviral Properties of Selenium Nanoparticles

1
Prokhorov General Physics Institute of the Russian Academy of Sciences, Vavilove St. 38, 119991 Moscow, Russia
2
Institute of Biology and Biomedicine, Lobachevsky State University of Nizhny Novgorod, Gagarin av. 23, 603105 Nizhny Novgorod, Russia
3
State Key Laboratory of Radiation Medicine and Protection, School for Radiological and Interdisciplinary Sciences (RAD-X), Collaborative Innovation Center of Radiological Medicine of Jiangsu Higher Education Institutions, Suzhou Medical College, Soochow University, Suzhou 215123, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(15), 5363; https://doi.org/10.3390/ma16155363
Submission received: 25 June 2023 / Revised: 21 July 2023 / Accepted: 28 July 2023 / Published: 30 July 2023
(This article belongs to the Special Issue Advances in Nanoparticle Antibacterial Materials)

Abstract

:
The resistance of microorganisms to antimicrobial drugs is an important problem worldwide. To solve this problem, active searches for antimicrobial components, approaches and therapies are being carried out. Selenium nanoparticles have high potential for antimicrobial activity. The relevance of their application is indisputable, which can be noted due to the significant increase in publications on the topic over the past decade. This review of research publications aims to provide the reader with up-to-date information on the antimicrobial properties of selenium nanoparticles, including susceptible microorganisms, the mechanisms of action of nanoparticles on bacteria and the effect of nanoparticle properties on their antimicrobial activity. This review describes the most complete information on the antiviral, antibacterial and antifungal effects of selenium nanoparticles.

1. Introduction

Despite the high level of medical development, microbial infections remain a significant factor in morbidity and mortality worldwide [1]. Sepsis and septic shock alone account for approximately 30 million clinical cases each year worldwide, of which 6 million are fatal [2]. The development of bacterial resistance to antibiotics is not a new problem and dates back to the 1940s–1960s [3,4]. Resistant microorganisms have a significant impact on human life and economic activity. Antibiotic-resistant bacteria significantly increase the risk of complications and death from bacteremia [1,5].
In addition, antibiotic-resistant bacteria complicate the course of foodborne illness, which accounts for more than a million deaths and 2 billion hospitalizations worldwide over a period of 20 years [6]. Complications of bacterial infections can affect almost all human tissues, organs and systems: gastrointestinal tract (gastritis, stomach ulcer, severe forms of diarrhea), central nervous system (meningitis, encephalitis), kidneys, liver, spleen, musculoskeletal system (reactive arthritis), cardiovascular system (endocarditis) and reproductive system (premature birth, stillbirth) [7,8,9,10,11,12,13]. By 2050, microbial resistance is predicted to cause a decrease in the total population of the Earth by 100–440 million people [14].
Apart from humans, antibiotic-resistant bacteria infect animals and plants, reducing the efficiency of agriculture [15,16,17]. Expected financial losses from antibiotic-resistant bacteria in 2025–2050 could be $85 trillion in GDP and $23 trillion in global trade [18]. The mechanisms of antibiotic resistance include enzymatic modification and inactivation of the antibiotic: hydrolysis, phosphorylation, glycosylation, etc. [19], reduction of cell wall permeability due to lipopolysaccharides or lipid enrichment, and removal of antibiotics from cells using special molecular pumps (multidrug [MDR] efflux pumps) [20].
In addition to bacterial antibiotic resistance, the development of fungal resistance to antifungal drugs is worth noting [21]. More than 300 million severe fungal infections have been registered in the world, of which over a million are fatal [22]. Additionally, the demand for antiviral drugs is increasing, in connection with the resistance of viruses, as well as in connection with the SARS-COV-2 pandemic [23], which affected all countries from 2019 until now. One of the promising ways to solve the problem of microbial resistance is the use of nanotechnology [24,25,26,27]. Antimicrobial nanomaterials operate through disruption of the electron transport chain [28,29], membrane destruction [30], cell division arrest [31], etc., and have garnered considerable attention due to their broad, potent and persistent bactericidal activities.
Unfortunately, recent data indicate the possibility of bacterial resistance to metal NPs and metal oxides [32,33,34,35,36,37,38,39]. Bacterial defense mechanisms include increased flagellin expression for NP sorption as well as pigment release to inactivate metal ions and activation of antioxidant defenses to combat oxidative stress [40,41,42]. Currently, an active search is underway for NPs with antibacterial properties among the nonmetallic chemical elements [33,43,44,45].
Selenium nanoparticles are one such type of nanoparticle. The study of their antimicrobial properties is a young branch of science: the vast majority of research papers (>95%) have been published in the last 10 years. According to PubMed NCBI (https://pubmed.ncbi.nlm.nih.gov/ accessed on 15 May 2023), more than 220 papers have been published with the keywords “selenium nanoparticles antimicrobial activity”, approximately 150 papers with the keywords “selenium nanoparticles antibacterial activity”, approximately 30 papers with the keywords “selenium nanoparticles antifungal activity” and approximately 20 papers with the keywords “selenium nanoparticles antiviral activity” (see Figure 1). It is important to note that every year, an increasing number of researchers pay attention to this problem. Over the past few years, research publication activity worldwide on SeNP antimicrobial properties has been growing by approximately 25–30% per year. Over 100 articles were published in 2022.
The reality of our days is the resistance of microorganisms (bacteria, fungi and viruses) to modern antibiotic, antifungal and antiviral drugs. Today, the scientific community is faced with the task of searching for new potential molecular structures to solve the problem of therapy in bacterial, fungal and viral pathogenesis. Selenium nanoparticles act as such antimicrobial agents. These nanoparticles have a fairly wide range of applications in the biomedical industry. For example, they can affect the activity of neutrophils: this immunomodulatory property of selenium nanoparticles can potentially be used in the treatment of cancer and other diseases associated with inflammation [46]. Moreover, selenium nanoparticles have a cytoprotective effect on the cells of the cerebral cortex under conditions of ischemia/reoxygenation [47]. The mechanism of action of this effect is based on the fact that nanoparticles regulate the expression of protective proteins in the cells of the cerebral cortex and reduce the total level of Ca2+ ions. The combination of selenium nanoparticles with the flavonoid taxifolin makes it possible to increase the neuroprotective effect in the cells of the cerebral cortex under conditions of ischemia/reoxygenation [48]. Beyond the immunomodulatory and neuroprotective effects, selenium nanoparticles also demonstrate anticancer effects. Selenium-sorafenib nanocomplexes showed better anticancer effects on HepG2 hepatic carcinoma cells than pure sorafenib [49]
According to patent searches and recent publications, SeNPs may have the several potential commercial applications (Figure 2).
The first application is the development of nutritional supplements for humans and veterinary needs [50]. It is noteworthy that SeNPs provide a more dosed supply of Se to the body compared to alternative sources, such as selenium cysteine. In particular, SeNPs, when administered orally to mice, causes less pronounced toxicity (survival is 4.5–5 times higher) and liver failure compared with the same amount of selencysteine also administered orally [51]. SeNPs can also be used in veterinary medicine as immunomodulatory agents. A drug based on nonspecific IgG/SeNPs for the correction of immunization in veterinary use did not have acute toxicity to laboratory mice [52].
The second application is the development of a test system for virus detection, for example, a test strip for the detection of anti-SARS-CoV-2 IgM and IgG in human serum and blood [53,54].
The third application is the creation of antimicrobial coatings for medical devices and personal care products [55,56]. Polyurethane/SeNP and polyvinylchloride/SeNP composites (polymers with NP coatings) inhibited the growth of bacteria Staphylococcus aureus, Staphylococcus epidermidis, Pseudomonas aeruginosa, Staphylococcus aureus, Escherichia coli, Streptococcus pneumoniae, etc., and fungi Cryptococcus neoformans, S. cerevisiae, Rhodotorula rubra, C. albicans, etc. [57,58]. The development of dressings and bandaging materials based on SeNPs to accelerate infected wound healing is also possible [59,60,61]
The fourth application is cancer prevention and treatment [62]. Anticancer drugs based on SeNPs cause the generation of ROS, which cause ER stress and the development of mitochondrial dysfunction. These processes enhance apoptosis of tumor cells [51,63]. Typically, to improve anticancer properties, SeNPs are functionalized with polymers or other biological antigens: arabinogalactans against A549 and MCF-7 cells (apoptosis induction), HepG2, polysaccharide–protein complexes (MCF-7 apoptosis induction), Ru(II)thiol/SeNPs (suppress tumor growth and angiogenesis) and polyethylenimine with folic acid [64,65,66]. The accuracy of delivery is ensured by a higher level of expression of folic acid receptors on cancer cells of most forms of cancer compared with normal human cells (for example, HepG2) [67,68]. Hyaluronic acid (A549 cells), ATP (depletion of mitochondrial membrane potential and oxidative stress) and other agents were also described [69,70]. The addition of anticancer agents (sorafenib) to SeNPs significantly increases the activity of others against aggressive cancer lines (increased calcium entry, ER stress and apoptosis of HepG2, gliablastoma A-172) [49,71,72]. Oridoni/SeNP composites have anticancer activity and reduce the viability of RAW264.7, KYSE-150 and EC970 cancer cells by ~50% [73].
The fifth application is the production of cosmeceuticals and nano-cosmeceuticals for skin, hair, nail and lip care and protection from wrinkles, photoaging, hyperpigmentation, dandruff and hair damage [74]. Selenium-containing cosmetics are developed and produced by Riga Stradins University, Phyto-C company (Newark, New York, USA), Cytolnat® (Paris, France), etc.
The sixth application is the production of nanoparticle fertilizers for crop production and soil health. Trichoderma-harzianum-culture-biosynthesized SeNPs (60 nm) reduced Fusarium beadium and Alternaria albicans growth, and fumonisin and Alternaria toxin production, in wheat crops [75]. The addition of selenium or its compounds (usually selenites, selenates, Se-methlyselenocysteine-containing peptides, etc.) to the soil is a common practice in agriculture [76]. All of these compounds are easily soluble and can quickly be washed out of the soil; no more than 20% of the initial selenium is retained for the next farming cycle [77]. The use of SeNPs should provide a longer and dosed supply of Se to the soil. It has been shown that the addition of 1 μg/L SeNPs to the soil improves seed germination and accelerates the growth of the radish Raphanus sativus, eggplant Solanum melongena, cucumber Cucumis sativus, tomato S. lycopersicum, and chilli pepper Capsicum annuum [78,79]. Selenium fertilizers are already commercially available, at least from the Yara company (https://www.yara.co.uk/crop-nutrition/fertiliser/yara-booster-range/ accessed on 20 June 2023). The mineral composition contains SeNPs, Ca(OH)2NPs and oxidized steel nanoparticles and can be used to improve the condition of fruit plants during periods of drought [80]. Agricultural uses may also include protecting plants from insect larvae. There is evidence of larvicidal activity of SeNPs [81]. SeNPs can be also used in the development of heavy metal and hydrogen peroxide sensors for agriculture [82,83,84].
The seventh application is the development of drugs for the treatment of inflammatory diseases (for example, recovery from a stroke) or diabetes [85,86]. In particular, it was shown that Protein/SeNPs inhibited hydroxyl radical productions in vitro [87]. Cystein/SeNPs inhibited hyperglycemia-induced ROS production in HUVEC by 40–50% [88]. Data on preclinical trials of a drug for the treatment of DM2T based on liposomes containing SeNPs have been published [89] (Table 1).
Other popular non-metal nanoparticles are SiNPs. SiNPs have a number of interesting advantages [92,93].
The first advantage is cheap synthesis. SiNPs are synthesized by the microemulsion method in the presence of oils or by the Stobers method, which requires relatively available reagents [94,95]. SiNPs can also be synthesized by a mechanochemical method, while the raw material for the synthesis of SeNPs can be river sand [96].
The second advantage is a wide range of applications, including targeted drug delivery with very high specificity [97,98,99], bioimaging [100,101,102], the creation of sensors for the detection of glucose, narcotic substances, or nucleic acids [103,104,105], the up-conversion of light (improved photosynthesis processes in agricultural plants), the protection of cereals from drought [106], and the creation of materials with antibacterial properties. SiNPs have the ability to significantly modify the surface and composition to create SiNP-based nanocomposites with antimicrobial properties [107,108,109].
The third advantage is the practically absent toxicity against eukaryotic cells. Oral administration of 1000–2000 mg/kg SiNPs did not have toxic effects in in vivo experiments and did not cause damage to internal organs in rats [110,111].
The disadvantages of SiNPs include the complexity/high cost of surface modification or the addition of metal NPs and the difficulty of obtaining SiNPs with uniform characteristics [93]. As a rule, precious-metal NPs and/or modification by several agents at once are required [104,108,109]. Antimicrobial properties have not been described for SiNPs without surface modification or addition of metal NPs [92]. SeNPs have been described as having their own antimicrobial properties [112,113,114]. In addition, modified SeNPs with antimicrobial properties can be obtained by biosynthesis, which significantly reduces the cost of their production [115,116,117].
Recently, a series of reviews has been published on the antibacterial, antifungal, anticancer, antiviral and antiparasitic properties of SeNPs [118,119,120]. It should be noted that in these works, the main emphasis is on a detailed description of the antimicrobial mechanisms of SeNPs and the contribution of conjugates to the properties of SeNPs. However, the numerical dependences of the properties of SeNPs—n particular, their size and method of preparation—are described to a lesser extent. In addition, it remains unknown to what extent the “size-antimicrobial-property” dependences of SeNPs are preserved when moving from one type of microorganism to another.
In this literature review, we aimed to analyze the dependence of the antimicrobial effect of selenium nanoparticles on their size, features of synthesis and microorganism species such as viruses, bacteria and fungi. There is no doubt that the relevance of this topic is great.

2. Synthesis Methods of Selenium Nanoparticles

In this subsection, we discuss various methods for the synthesis of selenium nanoparticles. SeNPs can be synthesized using a wide range of methods: sonochemical, reflux, microwave, hydrothermal, gamma irradiation, pulsed laser ablation, physical evaporation and “green synthesis” (biological reduction) (Figure 3). The most common method for selenium nanoparticle synthesis is chemical reduction [114]. The most commonly used precursors are SeO2, Na2SeO3, NaHSeO3 and H2SeO3 [121,122,123,124]. Less-common precursors are H2Se, Na2SeO4, SeCl4 or cyclo-octeno-1,2,3-selenadiazole [125,126,127,128].
As stabilizers, substances such as polysaccharides, quercetin, gallic and ascorbic acids and polyvinyl alcohol are used [129]. In addition to the stabilizer, a reducing agent (or reducer) is added to the solution, such as potassium tetrahydroborate (KBH4), ascorbic acid, hydrazine chloride, hydrazine hydrate (N2H4∙3H2O) or dimethylsulfoxide (C2H6OS) [130]. Sometimes substances of biological origin, mainly plant extracts, are used as reducing agents. This method of synthesis is called biological reduction [131,132]. The sonochemical method is a type of chemical reduction method in which the formation of metal NPs is formed by mixing soluble salts that react with each other to form a precipitate. Ultrasound accelerates of sediment formation. For exposure, a Ti tip immersed in a salt solution is used. It is noteworthy that, during precipitation in the presence of organic compounds, it is possible to obtain conjugated NPs [133]. The hydrothermal method is based on the reduction of inorganic precursors of SeNPs (Na2SO3) in aqueous solutions in the presence of organic reducing agents (for example, L-ascorbate) at an elevated temperature (~90 °C). An elevated temperature is necessary to accelerate the Se reduction reaction [134]. The reflux method is a variant of the chemical reduction method that occurs when the reaction mixture is boiled for a long time, which makes it very similar to the hydrothermal method. To increase the boiling time, a special unit with a refrigerant is used to condense the solvent vapors and return them to the reaction mixture [135]. The microwave method for the synthesis of nanoparticles is based on the reduction reaction of selenium-containing precursors in a medium with a reducing agent [136]. Microwave radiation provides a multiple acceleration of the reduction reaction due to heating of the reaction. The time of microwave exposure determines the type of crystal lattice of the synthesized Se nanoparticles [128,137]. In the method of gamma irradiation on the reduction of SeO2 to Se0 [138], recovery occurs due to processes associated with water radiolysis [139]. In addition, it has been shown that gamma radiation can enhance the biogenic synthesis of SeNPs using fungi [140,141]. Principally, the microwave method and the use of gamma radiation can be attributed to the group of chemical reduction methods, but we singled them out separately, since they additionally need high-energy electromagnetic sources.
Laser ablation in a liquid belongs to the class of physical synthesis methods. It usually consists of obtaining nanoparticles using intense laser radiation from the surface of massive crystalline selenium targets immersed in a liquid. The laser ablation method has been described in detail [142]. During further laser irradiation of a colloid with nanoparticles, it is possible to obtain smaller NPs. This process is called fragmentation [143]. Physical evaporation is a method for synthesizing nanoparticles by treating a metal target with a laser in a vacuum, inert gas or atmospheric air. The method is essentially laser ablation carried out not in a solution, but in a gas and/or vacuum [144,145], so these methods were combined in this work.
In the case of the biological reduction method, biogenic synthesis and bioorganic synthesis are distinguished. In a significant number of works, the synthesis of SeNPs was carried out by biological reduction (“green synthesis”). This approach is also based on the reduction of selenium-containing precursors to Se0. The essential difference between the first and the second type is that biogenic synthesis uses cellular structures [115,131,146,147,148,149] and bioorganic synthesis uses noncellular extracts of plant and microbial origin to synthesize nanoparticles [117,132,150,151,152,153]. In this case, reducing agents are usually secondary metabolites, such as flavonoids, thiamine and capsaicin, as well as other agents containing amino groups, extracted from various plants: Spirulina platensis, Azadirachta indica, Trigonella foenum-graecum, Allium sativum, etc. [152,154,155,156,157]. The use of extracts during synthesis provides antibacterial, anticancer, or antioxidant properties of SeNPs [154,155,156,157]. SeNPs can be synthesized by cultivating microorganisms in a medium with an excess of SeNP precursors (sodium selenite (Na2SeO3) or SeO2) [158]. For the synthesis of SeNPs, biomass and/or cell-free supernatants of bacterial cultures (Lactobacillus brevis, Lactobacillus casei, Bacillus licheniformis, Pseudomonas alcaliphila, etc.) [158,159,160,161,162], fungi (Aspergillus oryzae, Penicillium citrinum, Mariannaea sp.) [140,141,163] or yeasts (Saccharomyces cerevisiae, Magnusiomyces ingen) [164,165] can be used. Figure 4a shows the percentages of the various methods for the synthesis of selenium nanoparticles found in the literature.

2.1. Influence of the Method of Synthesis of Selenium Nanoparticles on the Resulting Size and Shape of Nanoparticles

Does the method of synthesis of nanoparticles affect their size? The results of our analysis are presented in Figure 4b. Obviously, physical methods of synthesis, such as laser ablation or microwave irradiation, make it possible to achieve a narrow size distribution of nanoparticles (Figure 4b). Most often, these are spherical particles less than 200 nm. Chemical or biological synthesis methods produce a wide range of particle sizes from 5 to 500 nm (Figure 4b). Within the framework of one type of synthesis, a preparation of nanoparticles with a wide size distribution of nanoparticles is usually obtained. At the same time, the vast majority of publications have synthesized spherical selenium nanoparticles. In some cases, other shapes, such as elongated cylinders [166], polygons [167] and granules [146], are observed.

2.2. Influence of Selenium Nanoparticle Synthesis Method on the Minimum Inhibitory Concentration in Antibacterial Studies

It has been established that the selenium nanoparticle synthesis method affects the value of the minimum inhibitory concentration in antibacterial studies (Table 1). The results of the analysis are shown in Figure 4c. It was noted that when using physical methods for the synthesis of selenium nanoparticles, the minimum inhibited concentration for effective antibacterial action did not exceed 100 µg/mL. When using microwave generation of nanoparticles, the MIC is approximately 100–300 µg/mL, which is significantly worse compared with nanoparticles obtained by other methods.
It should be noted that there are few studies that use nanoparticles synthesized using physical methods. Most likely, in the future, we should expect clarification of the data presented. For the methods of chemical and biological synthesis, the distribution of the obtained values of the minimum inhibited concentration is quite wide. In some cases, MICs of less than 1 µg/mL have been reported. At the same time, the average efficiency of nanoparticles obtained by both chemical and biological synthesis does not differ significantly. In general, the average efficiency of nanoparticles obtained by chemical/biological methods and using laser ablation does not differ significantly.
One of the ways to enhance the antibacterial and antifungal properties of SeNPs is their functionalization with enzymes and polymers during synthesis or inclusion in a polymer matrix. Examples of such components can be: arabinogalactan, propolis, lysozyme, bacterial cellulose, gelatin, reduced graphene oxide, chitosan, collagen and ε-poly-L-lysine [121,125,168,169,170,171,172,173]. The addition of SeNPs to textile fiber significantly enhances their antimicrobial properties [122,174]. The use of plant extracts (Urtica dioica, Ricinus communis, Artemisia annua, Nepeta sp., etc.) during synthesis also enhances the antibacterial and antifungal effects of SeNPs [117,148,150,175]. The addition of SeNPs to titanium oxide nanotubes enhances the antibacterial properties of the latter [176,177,178].

3. Effective Concentration/Minimum Inhibitory Concentration of Selenium Nanoparticles Depending on Their Size

3.1. Dependence of the Effective Concentration of Selenium Nanoparticles on Their Size for the Study of Antiviral Activity

We consider it necessary to start presenting the results with a study of antiviral activity. There are very few such research publications [120]. It seems that this is due to the increased complexity of organizing scientific work with viruses. Figure 5 shows the effective concentration of selenium nanoparticles depending on their size in the study of antiviral activity. The dependence is described with the equation y = 0.035·x + 1.08. By the nature of this dependence, it can be concluded that with a decrease in the size of nanoparticles, their effective concentration for antiviral action also decreases. This means that as the size of nanoparticles decreases, on average, they become more effective against viruses.
It should be noted that selenium nanoparticle synthesis, which is used in most antiviral studies, is a chemical reduction method. In the majority of research papers, spherical nanoparticles with a diameter of 10 to 200 nanometers were used. Published papers mainly use the H1N1 influenza virus (H1N1 influenza infecting Madin Darby canine kidney cell line) [179,180,181,182]. In addition, there are research publications investigating Enterovirus (Enterovirus 71—EV71) [183,184], hepatitis virus (HAV) [81,185,186], Cox-B4 virus (enteroviruses) [81], herpes virus (HSV-2 Herpes simplex II), influenza virus H1N1 [187] and adenovirus (Adenovirus strain 2) [187].
In the published research papers, two problems are investigated. The first task is to study the general antiviral action of selenium nanoparticles. The second task is to increase antiviral activity with the help of selenium nanoparticles in case of resistance of the virus to antiviral drugs. In other words, in the first task, researchers use “pure” selenium nanoparticles, and in the second task, selenium nanoparticles are processed (=functionalized) with the help of antiviral drug molecules. That is, selenium nanoparticles act as substrates for targeted delivery.

3.2. Dependence of the Minimum Inhibitory Concentration of Selenium Nanoparticles on Their Size and Shape in the Study of Antibacterial Activity

A large number of studies are published on this topic every year. For convenience, we present Table 2, which contains the analyzed publications of the antibacterial action of selenium nanoparticles. It should be noted that in most published research papers, three types of bacteria are used as objects: Bacillus subtilis, Staphylococcus aureus, and Escherichia coli. A box-and-whisker plot of the values of the minimum inhibitory concentration for these three types of bacteria is shown in Figure 6a.
Table 2. Dependence of the antimicrobial properties of SeNPs on the size, composition and method of synthesis.
Table 2. Dependence of the antimicrobial properties of SeNPs on the size, composition and method of synthesis.
PrecursorCompositionMethod of the SynthesisParticle Size, nmMicroorganism StrainsEffectMICResultsReference
1Na2SeO3Lysozyme SeNPsChemical reduction35.6Escherichia coli,
Staphylococcus aureus
BS82 μg/mLSeNPs and lysozyme demonstrated synergetic bacteriostatic activity.[121]
2Na2SeO3Propolis SeNPsBioorganic chemical reduction159, 151.9, 11.2 and 169.3Salmonella typhimurium ATCC 14028,
Escherichia coli ATCC 25922,
Staphylococcus aureus- ATCC 25923
BC25 mg/L
27.5 mg/L
30 mg/L
BNCt/Pro/SeNPs were the most effective against all bacterial strains.[168]
3Na2SeO3SeNPsChemical reduction32.3Staphylococcus aureus (MSSA),
Staphylococcus aureus (MRSA),
Staphylococcus aureus (VRSA),
Enterococci (VRE)
BC; BS20 µg/mL,
80 µg/mL, 320 µg/mL, and >320 µg/mL
SeNPs showed a synergistic effect with linezolid (LZD) through protein degradation against MSSA and MRSA.[112]
4Na2SeO3SeNPsBiosynthesis of SeNPs by Providencia sp.120P. aeruginosa,
E. coli,
V. parahemolyticus,
S. aureus,
B. cereus,
B. subtilis
BC; BS500 mg/LBio-SeNPs showed strong antibacterial effects on the five of pathogens at 100 mg/L. It was shown that most of G-bacteria (P. aeruginosa, E. coli and V. parahemolyticus) were locally killed by 500 mg/L of the bio-SeNPs after 12 h, which was better than G+-bacteria (S. aureus and B. cereus, except for B. subtilis).[188]
5Se (solid)SeNPsPulsed laser ablation in liquids~80 and ~10E. coli (MDR-EC) ATCC BAA-2471,
P. aeruginosa (PA) ATCC 27853,
S. aureus (MRSA) ATCC 4330
Staphylococcus epidermidis ATCC 35984
BC; BS25 µg/mLSeNPs showed a dose-dependent antibacterial effect toward both standard and antibiotic-resistant phenotypes of Gram-negative and Gram-positive bacteria.[113]
6NaHSeO3SeNPs with polyester
fabrics
Chemical reduction40–60Salmonella typhi,
Bacillus cereus,
Escherichia coli,
Pseudomonas aeruginosa
BC1980 µg/mLThe treated fabric under study showed excellent killing potentiality against Gram-positive and Gram-negative bacteria.[174]
7NaHSeO3Leather material/
SeNPs
Chemical reduction36–77 and
41–149
Bacillus cereus,
Pseudomonas aeruginosa,
Salmonella typhi,
Escherichia coli
BC1980 µg/mLPotential application to the footwear industry to color the leather as well as prevent the spread of bacterial infection promoted by humidity, poor breathability and temperature.[122]
8Na2SeO3SeNPs/
orange peel waste extract
Bioorganic chemical reduction16–95Pseudomonas aeruginosa PAO1, MDR,
S. aureus ATCC 29213
BS25 µg/mLThe biosynthesized SeNPs had a promising antibiofilm activity, where the largest inhibition of biofilm was noticed in MDR K. pneumonia.[169]
9Na2SeO3bacterial cellulose/
gelatin/
SeNPs hydrogels
Chemical reduction75E. coli,
S. aureus
BC65.44 μgBC/Gel/SeNPs nanocomposite hydrogel: potential wound dressing for preventing wound infection and accelerating skin regeneration in clinic.[189]
10H2SeO3rGO-S/Se compositeChemical reduction12Staphylococcus aureus,
Enterococcus faecalis
BS200 µg/mLConcentration and time-dependent BS activity of the rGO (Reduced graphene oxide)-S/Se NP against S. aureus cells[170]
11Na2SeO3SeNPsBiosynthesis of SeNPs by cyanobacteria Anabaena sp. 25Staphylococcus aureus Escherichia coliBS50 µg/mLThese biogenic SeNPs demonstrated significant antibacterial and anti-biofilm activity against bacterial pathogens.[131]
12Na2SeO3Ag-SeNPsBiosynthesis of SeNPs by Aureobasidium pullulans50 and 70Staphylococcus aureus F1557
E. coli WT F1693
BC; BS-The Ag–Se coating reduced 81.2% and 59.7% of viable bacterial adhesion. The antibacterial mechanism of Ag–Se coatings works through effective contact-killing activity against S. aureus.[146]
13Na2SeO3SeNP-chitosan,
SeNPs-carboxymethyl cellulose
Chemical reduction55–500
50–300
Staphylococcus aureus,
methicillin-resistant Staphylococcus aureus,
Staphylococcus epidermidis
BS5 µg/mLThe SeNP-modified collagenous scaffolds at SeNP concentrations low as 5 µg/mL showed a strong antibacterial effect (up to 94% of bacterial growth inhibition) toward laboratory and clinical isolates of Gram-positive bacteria from the genus Staphylococcus.[171]
14Na2SeO3SeNPsBiosynthesis from Stenotrophomonas maltophilia SeI TE02181P. aeruginosa PAO1, INT, BR1 and BR2,
S. maltophilia VR10 and VR20,
Achromobacter xyloxidans strain C,
Burkholderia cenocepacia strain LMG 16656,
Staphylococcus aureus Mu50 strain,
S. aureus UR1,
Staphylococcus epidermidis ET024,
Staphylococcus hemolitycus UST1
BS4–128 μg/mLThe progressive loss in protein and carbohydrate content of the organic cap determines a decrease in nanoparticle stability. This leads to an alteration of size and electrical properties of SeNPs along with a gradual attenuation of their antibacterial efficacy.[190]
15Na2SeO4SeNPsBiosynthesis of SeNPs from Aspergillus quadrilineatus and Aspergillus ochraceus isolated from the twigs and leaves of Ricinus communis45–75Pseudomonas aeruginosa ATCC 15442,
Bacillus cereus ATCC 10876,
Staphylococcus aureus ATCC 6538,
Klebsiella pneumoniae ATCC 13883,
Bacillus subtilis TCC 6633,
Escherichia coli ATCC 11229
BS62.5–1000 µg/mLSeNPs showed potent antifungal and antibacterial potentials against different human and phyto-pathogens.[175]
16Na2SeO3“Green” SeNPsBiosynthesis using aqueous leaf extract of U. dioica10–87.4Staphylococcus aureus ATCC 25923,
Bacillus subtilis ATCC605,
Escherichia coli ATCC 25922,
Pseudomonas aeruginosa ATCC 27853
Fungi:
Candida albicans,
Cryptococcus neoformans
BS, FS125,
62.5 and
15.62 µg/mL
3.9 and 7.81 µg/mL
SeNPs exhibited promising antibacterial activity against Gram-negative and Gram-positive bacteria and antifungal activity.[150]
17SeO2SeNPs/tree gumChemical reduction105.6Bacillus subtilis,
Micrococcus luteus
BS12 μg/mLThe synthesized SeNPs inhibited the growth of the Gram-positive bacteria B. subtilis only.[123]
18Na2SeO3SeNPs/chitosanChemical reduction100Staphylococcus aureus,
Escherichia coli
BS158 μg/mLThe antibacterial activity of CS(H)-SeNPs markedly decreased owing to the
aggregation of NPs.
[166]
19Na2SeO3SeNPsPhytofabrication of SeNPs from aqueous Spirulina platensis79.4Salmonella abony NCTC 6017,
Klebsiella pneumonia ATCC 700603,
E. coli ATCC 8739
BS25–200 µg/mLSeNPs have shown potent antimicrobial activity against Gram-negative bacteria. No toxic effect was observed for SeNPs on normal kidney and liver cell lines.[191]
20Na2SeO3SeNPsBiosynthesis of SeNPs from Nepeta plant powder75P. aeruginosa: ATCC 27853 and
A. baumannii: ATCC BAA-747
BS4 μg/mL
8 μg/mL
The inhibition of bacterial growth demonstrated in the presence of lower concentrations of SeNPs than common antibiotics.[117]
21Na2SeO3Collagen/Chitosan/SeNPsChemical reduction100–200Staphylococcus aureus NCTC 8511, MRSA CCM 7110 and
Escherichia coli NCTC 13216
BS0.5–5 µg/mLSeNPs are able to enhance the scaffold’s antibacterial properties toward S. aureus and MRSA at concentrations between 0.5 µg/mL and 5 µg/mL.[172]
22Na2SeO3B-SeNPsBiosynthesis of SeNPs from Anabaena variabilis (cyanobacteria)10.8Bacillus subtilis,
Staphylococcus aureus,
Escherichia coli,
Klebsiella pneumoniae
BS20 µg/mLCyanobacteria mediated synthesis
can be considered as safe and nontoxic way to synthesize SeNPs.
[116]
23Na2SeO3SeNPsBiosynthesis of SeNPs-S by Bacillus sp. Q33159.2E. coli,
P. aeruginosa,
S. aureus,
L. monocytogenes
BS200 µg/mLSeNPs-S (product of whole cells) and SeNPs-E (product of the extracellular extract) exhibited obvious inhibitory effects on the four pathogenic bacteria.[115]
24Se (wafer)SeNPsChemical reduction42E. coli,
S. aureus
BS0.2 mg/mLThe synergistic antibacterial effect of SeNPs and microstructured
parylene-C.
[192]
25H2SeArabinogalactan/
SeNPs
Bioorganic synthesis with AG from Larix Sibirica isolated94bacterial phytopathogen Clavibacter michiganensis sepedonicus (Cms)BS6.25 μg/mLAntimicrobial activity of AG/SeNPs is due to their ability to inhibit the dehydrogenase activity of Cms cells, to disrupt the integrity of the cell membrane, resulting in a decrease of transmembrane potential and reduction of cellular respiration.[125]
26Na2SeO3Cefotaxime/Ag–SeNPsGamma irradiation34.5; 24.9E. coli,
P. aeruginosa,
K. pneumoniae,
S. aureus,
Enterococcus sp.
BS; BC2.5–5 μg/mL;
0.625–2.5 μg/mL (with CFM)
Ag NPs-CFM, SeNPs-CFM and Ag–SeNPs-CFM possessed antimicrobial activity against Staphylococcus aureus, Escherichia coli.[193]
27Na2SeO3sSeNPsChemical reduction70Porphyromonas gingivalisBS; BC4–16 μg/mLThe growth of P. gingivalis
was significantly inhibited by SeNPs.
[114]
28SeO2Se NP-ε-poly-L-lysinechemical reduction82S. aureus ATCC 29213,
S. aureus (MRSA) ATCC 43300,
E. faecalis ATCC 29212,
E. coli ATCC 25922,
A. baumannii 2208,
ATCC 19606,
P. aeruginosa strain PAO1-LAC ATCC 47085,
K. pneumoniae ATCC 13883, and
K. pneumoniae (MDR) FADDI-KP628
BS; BC6−26 μg/mLThe MICs of Se NP-ε-PL against the eight different types of bacteria tested are approximately 6–26 μg/mL.[173]
29Na2SeO3SeNPsBiosynthesis by Se-resistant Bacillus subtilis AS1277Aeromonas hydrophilia
Staphylococcus aureus,
Bacillus cereus,
Listeria monocytogenes,
Escherichia coli,
Aeromonas hydrophilia,
Klebsiella pneumonia
BS; BC3–5 μg/mLBio-SeNPs can mitigate the accumulation of heavy metals and reduce the bacterial load in a concentration-dependent manner.[194]
30Na2SeO3TiO2 nanotube with SeNPschemical reduction88.93E. coliBS-Selenium nanoparticles improved antibacterial properties of titanium dioxide nanotubes.[176]
31Na2SeO3SeNPsmicrowave technique in the presence of citric acid10.5–20P. aeruginosa,
E. coli,
B. subtilis,
S. aureus
BS100 mg/mLSeNPs had the highest activity against E. coli, with a zone of inhibition (ZOI) of 25.2 ± 1.5 mm compared to 16.0 ± 0.6 mm for the standard antibiotic.[136]
32Na2SeO3SeNPsbiosynthesis of SeNPs by endophytic fungal strain Penicillium crustosum EP-13–22Bacillus subtilis ATCC 6633, Staphylococcus aureus ATCC 6538,
Escherichia coli ATCC 8739,
Pseudomonas aeruginosa ATCC 9022
BS12.5 µg/mL
50 µg/mL
50 µg/mL
25 µg/mL
(in the presence of light)
50 µg/mL, 100 µg/mL (under dark conditions)
The effect of SeNPs was dose-dependent, and higher activities against bacteria were attained in the presence of light than were attained under dark conditions.[147]
33Na2SeO3Mk-SeNPschemical reduction with the presence of aqueous berry extract of Murraya koenigii (Mk-SeNPs)50–150Streptococcus mutans (HQ 693279.1 & ATCC 25175),
Enterococcus faecalis Shigella sonnei,
Pseudomonas aeruginosa (K 7769531 & HQ 693272
BS; BC40 μg/mL
50 μg/mL
Mk-SeNPs are considered to be a prospective antibacterial agent
with effective antioxidant capacity at 25 and 50 μg/mL, which is target-specific only for the bacterial cells and not for the erythrocytes and macrophages at the same concentration.
[195]
34Na2SeO3SeNPsThe abiotic
reduction of selenite with the use of Enterococcus spp. cell-free extract (biotic and abiotic stages)
200E. coliBS3.2 g/LThe obtained nanoparticles exhibited antimicrobial properties by directly inhibiting the viability of an E. coli bacterial strain. The results demonstrate not only the potential of abiotic production of SeNPs but also the potential for these particles as microbial inhibitors in medical or similar fields.[196]
35Na2SeO3SeNPschemical reduction with PVA as a stabilizer30–70S. aureus (ATCC 29213),
E. coli (ATCC 25922)
BS1 μg/mLThe growth of S. aureus was inhibited by the nanoparticles at concentrations as low as 1 μg/mL.[197]
36Na2SeO3Artemisia annua extract/SeNPsBiosynthesized using Artemisia
annua, and subsequently, the surface of the biogenic SeNPs was functionally modified with starch.
<200Staphylococcus aureus,
Bacillus cereus,
Salmonella enterica,
Escherichia coli
BS; BC5–100 μg/mLStAaSeNPs showed the highest antibacterial activity against tested strains S. enterica (23.26 ± 0.35 mm). Based on the findings, it can be inferred that surface
chemistry is the most influential factor in determining the antibacterial activity of SeNPs.
[148]
37Na2SeO3Hollow SeNPsBioorganic synthesis of SeNPs with the potato extract115B. subtilis (MTCC441),
E. coli (MTCC40)
BC10–20 μg/mLThe hSeNPs showed good antibacterial activity against tested bacteria.[198]
38SeO2SeNPschemical reduction with the presence of PVA43–205Staphylococcus aureus (MSSA) ATCC 29213,
Staphylococcus aureus (MRSA) ATCC 43300
BC16 µg/mLThe SeNPs were shown to have multimodal mechanisms of action that depended on their size, including depleting internal ATP, inducing ROS production, and disrupting membrane potential.[199]
39Na2SeO3BSA/
SeNPs
chemical reduction method in the presence of the BSA20–30Escherichia coli (ATCC no. 25922),
Escherichia coli (ATCC no. BAA-2471),
Staphylococcus aureus (ATCC no. 25923)
BC1 mg/mLSeNPs achieved a 10-fold reduction for S. aureus.[200]
40SeO2eADF4(κ16)/SeNPs
PVA/
SeNPs
chemical reduction method in the presence of the spider silk protein eADF4(κ16) and PVA (Polyvinyl alcohol)46Escherichia coliBC8 ± 1 µg/mL
405 ± 80 µg/mL
The eADF4(κ16)-coated SeNPs demonstrated a much higher bactericidal efficacy against the E. coli, with a minimum bactericidal concentration (MBC) approximately 50 times lower than that of PVA/SeNPs.[201]
41Na2SeO3mycogenic
SeNPs
SeNPs-CN
2 methods:
a biogenic process using Penicillium chrysogenum filtrate and
by utilizing gentamicin drug
(CN) following the application of gamma irradiation
33.84
22.37
Staphylococcus aureus,
Bacillus subtilis,
Pseudomonas aeruginosa,
Escherichia coli,
Klebsiella pneumoniae
Fungi:
Candida albicans
BC, FC0.490 μg/mL
0.245 μg/mL
The synthesized SeNPs-CN possesses an encouraging antimicrobial potential with respect to the biogenic SeNPs against all examined microbes. Remarkably, SeNPs-CN showed antimicrobial potential toward 23.0 mm ZOI for Escherichia coli and 20.0 mm ZOI against Staphylococcus aureus. It also inhibited the expansion and invasion of
C. albicans suggested the use of gentamycin as antifungal agent after the combination with the synthesized SeNPs.
[202]
42Se (pellets)SeNPsPulsed laser
ablation in liquids
115Escherichia coli,
Staphylococcus aureus
BC50 µg/mLThe pure selenium nanoparticles determined the minimal concentration required for ~50% inhibition of either E. coli or S. aureus after 24 h to be at least ~50 µg/mL.[203]
43H2SeO3Green Orange Peel extract/SeNPsChemical reduction in the presence of BSA—stabilizer18.3S. aureus (ATCC 25923),
S. epidermidis (ATCC 1228)
BS4.94 μg/LThe SeNP sample demonstrated excellent antibacterial activity with an average diameter of inhibition zones of 20.0 mm and an MIC of 4.94 μg/L.[124]
44Na2SeO3SeNPsBioorganic synthesis of SeNPs with the use of Penicillium corylophilum As-1 biomass filtrate, in presence of ascorbic acid as a reducing agent29.1–48.9Staphylococcus aureus ATCC 6538,
Bacillus subtilis ATCC 6633,
Escherichia coli ATCC 8739,
Pseudomonas aeruginosa ATCC 9027
BS9.37 μg/mL
18.75 μg/mL
37.5 μg/mL
37.5 μg/mL
The formed SeNPs showed a prominent antimicrobial activity at different concentrations against the pathogens Staphylococcus aureus, Bacillus subtilis, Pseudomonas aeruginosa and E. coli.[132]
45SeO2Penicillium expansum/SeNPsbiosynthesis with Penicillium expansum ATTC 362004–12.7Bacillus subtilis ATCC 6051, Staphylococcus aureus ATCC 23235,
Escherichia coli ATCC 8739,
Pseudomonas aeruginosa ATCC 9027
Fungi:
Candida albicans ATCC 90028,
Aspergillus fumigatus RCMB 02568,
A. niger RCMB 02724
BS, FS62.5 μg/mL
62.5 μg/mL
125 μg/mL
125 μg/mL
125 μg/mL
125 μg/mL
125 μg/mL
The inhibitory effect against Gram-positive bacteria was more pronounced than against Gram-negative bacteria and fungi.[149]
46Na2SeO3Chitosan/SeNPschemical reduction77Streptococcus mutansBS128 and 64 µg/mLThe comparison between the treated and untreated groups showed that combining therapy with SeNPs and PDT markedly decreased colony-forming units of one-day-old S. mutans biofilm.[204]
47SeO2SeNPssolvothermal method using Moringa oleifera leaf extract as a reducing agent82.86Listeria innocua ATCC 33090,
Bacillus cereus ATCC 10876,
Escherichia coli ATCC 43888,
Salmonella typhimurium ATCC 14028
BS100 μg/mLZones of inhibition were observed only in
S. typhimurium (12.5 ± 0.5 mm), E. coli (10.1 ± 0.7 mm) and B. cereus (9.8 ± 0.7 mm).
[167]
48Na2SeO3SeNPsgreen synthesis using ascorbic acid as a reducing agent and methanolic
extract of Calendula officinalis L. flowers as a stabilizer
40–60Serratia marcescens,
Enterobacter cloacae,
Alcaligenes faecalis
BS-The antibacterial activity of the extract, AsAc, and Na2SeO3 was enhanced by producing the SeNPs, which significantly inhibited the growth of S. marcescens,
E. cloacae, and A. faecalis bacterial strains.
[205]
49Na2SeO3SeNPsChemical reduction synthesis from extracts of three plants: Allium cepa (onion), Malpighia emarginata (acerola), and Gymnanthemum amygdalinum (boldo)245–321Streptococcus agalactiae,
Staphylococcus aureus,
S. aureus,
Pseudomonas aeruginosa,
Escherichia coli
BS6.125 to 98 μg/mLThe antimicrobial
activity and low hemolytic concentration indicate the possibility of use against Gram-positive bacteria, including
multidrug-resistant ones, opening a wide variety of options for their application
[151]
50SeO2Algae/
SeNPs
Microwave-assisted synthesis of SeNPs40V. harveyi (PTCC 1755)BS200 μg/mLThe presence of different functional groups of Sargassum angustifolium on the surface of the algae-coated SeNPs might be responsible for the more effective reaction of these nanoparticles with the cell walls and/or membrane of V. harveyi.[206]
51Na2SeO3SeNPschemical reduction71V. cholerae O1 ATCC 14035 strainBS50–200 μg/mLSeNPs are safe as an antibacterial and antibiofilm agent against V. cholerae O1 ATCC 14035 strain.[207]
52Na2SeO3SeNPs,
NCT/GA,
NCT/GA/Eug and
NCT/GA/Eug/SeNPs
chemical reduction9.7, 124.8, 132.6 and 134.2 nmEscherichia coli,
Staphylococcus aureus
BS; BC15.0 μg/mL
20 μg/mL
The entire fabricated nanocomposite exhibited potent antibacterial activity and cell destruction capability within 5–10 h of exposure.[208]
53Na2SeO3SeNPssynthesis and purification, in the presence of pepper extract; chemical reduction method, plus microwave79–90 nmEscherichia coli ATCC BAA-2471,
Staphylococcus aureus ATCC 4330
BS72,2 μg/mL
85,1 μg/mL
Selenium nanoparticles were biocompatible and showed bacteriostatic activity[152]
54Na2SeO3SeNPsbiosynthesized with a standard strain of C. albicans38Fungi:
Candida albicans,
Candida glabrata
FS1 and
0.5 µg/mL
SeNPs showed much better fungistatic activity compared to itraconazole, amphotericin B and anidulafungin.[209]
55Na2SeO3SeNPsBiosynthesis from lactic acid bacteria (LAB)56Fungi:
Candida and Fusarium species
FC80–130 µg/mLThe LAB-SeNPs MFC was in the range of 80–130 µg/mL, which ensured the complete killing of all tested fungi.[210]
56Na2SeO4SeNPsBiosynthesis with standard strains of A. Flavus and C. albicans37 and 38Fungi:
Candida and Aspergillus species
FS0.5, and 0.25 μg/mLThe utilization of SeNPs at concentrations of 1, 0.5 and 0.25 μg/mL or in, some strains, even lesss than 0.125 μg/mL, resulted in zero growth of fungal agents.[126]
57Na2SeO3SeNPs“green” method using the Halomonas elongata bacterium11Fungi:
Candida albicans
FS-The synthesized NPs in optimal situation stopped the growth of Candida albicans up to 72%.[211]
58Se (pellets)Chitosan/SeNPslaser ablation in liquids100Fungi:
C. albicans TW17 and 6486 strains
FC, FS3.5 μg/mLTaken separately, SeNPs and CS have shown fungicidal properties, but when combined (CS-SeNPs), achieved a potent inhibitory effect against the mature biofilm in a dose–response manner.[212]
59Na2SeO3SeNPsbiosynthesis with the leaf extract of Melia azedarach74Fungi:
Fusarium mangiferae
FC300 μg/mLBiogenic selenium NPs are widely expected to be efficient and cost-effective treatments
for fungal plant diseases.
[213]
60Na2SeO3SeNPsgreen synthesis using extracts
from A. glaucum leaves and C. officinalis flowers
8 and 133Fungi:
Fusarium oxysporum,
Colletotrichum gloeosporioides
FS0.25 mg/mLIt was observed that both SeNPs had antifungal activity against both plant pathogens at concentrations of 0.25 mg/mL and above. SeNPs-AGL demonstrated better antifungal activity
and smaller size (approximately 8.0 nm) than SeNPs-COF (134.0 nm).
[214]
61Na2SeO3SeNPsbiosynthesis by Lactobacillus aci-
dophilus ML1
46Fungi:
Fusarium culmorum,
Fusarium graminearum
FC100 mg/mLUnder greenhouse conditions, the wheat supplemented with BioSeNPs (100 mg/mL) experienced significantly incidence of crown and root rot diseases by 75% and considerably enhanced plant growth, grain quantity and quality by 5–40%.[215]
62SeCl4SeNPsbiosynthesis using endophytic fungus Fusarium oxysporum42Fungi:
Aspergillus niger
FS8 mg/mL; diluted to 4, 2, 1, 0.5, and 0.25 mg/mLSeNPs showed excellent antifungal and antisporulant activity against black fungus Aspergillus niger, which has become life-threatening to SARS-CoV-2 patients during the pandemic.[127]
63Na2SeO4SeNPsbiosynthesis with the use of Aspergillus strains64.8Fungi:
Aspergillus fumigatus,
Aspergillus flavus
FS0.5 µg/mLThe MIC of itraconazole and amphotericin B against A. fumigatus and A. flavus was 4 μg/mL, whereas the MIC values for treated samples with SeNPs have decreased to 0.5 μg/mL.[216]
64Na2SeO3SeNPsbiosynthesis using the extract of Melia azedarach leaves61Fungi:
Puccinia striformis
FS30 mg/LSeNPs at a concentration of 30 mg/L reduced the disease severity and enhanced the morphological, physiological, biochemical and antioxidant parameters.[217]
65Na2SeO3PPE/SeNPs and
NCT/PPE/SeNPs
Pomegranate peel extract (PPE) used for biosynthesis9.4
85
Fungi:
Penicillium digitatum
FC22.5
15 mg/mL
NCT/PPE/SeNPs nanocomposite was the most effective and significantly exceeded the fungicidal action of standard fungicide. The direct treatment of fungal mycelia with NCT/PPE/SeNPs nanocomposite led to remarkable lysis and deformations of P. digitatum hyphae within 12 h of treatment.[218]
66H2SeO3SeNPsBiosynthesis by Bacillus megaterium
ATCC 55000
41.2Fungi:
Rhizoctonia solani RCMB 031001
FS,
FC
0.0625 and
1 mM
SeNPs improve morphological and metabolic indicators and yield significantly compared with infected control.[219]
67Na2SeO3SeNPschemical reduction method with the use of the Trichoderma
atroviride cell culture lysate
93.2–98.5Fungi:
Pyricularia grisea,
Colletotrichum capsici,
Alternaria solani on chili and tomato leaves
FS50 μg/mL
100 μg/mL
100 μg/mL
The synthesized nanoparticles displayed excellent in vitro antifungal activity against Pyricularia grisea and inhibited the infection of Colletotrichum capsici and Alternaria solani on chili and tomato leaves.[220]
68Na2SeO3bovine serum albumin (BSA)/
SeNP, ascorbic acid/)/
SeNP, chitosan/SeNP, glucose/SeNP
chemical reduction method70–300Staphylococcus aureus (ATCC 6538),
Enterococcus faecalis (ATCC 29212),
Bacillus subtilis (ATCC 6633), and
Kocuria rhizophila (ATCC 9341), Escherichia coli (ATCC 8739),
Salmonella sp. (NCTC 6017),
Klebsiella pneumoniae (NCIMB 9111), Pseudomonas aeruginosa (ATCC 9027),
Fungi:
Candida albicans (ATCC 10231)
BS,
BC,
FC
100 μg/mL
100 μg/mL
200 μg/mL
200 μg/mL
400 μg/mL
400 μg/mL
200 μg/mL
400 μg/mL
25 μg/mL
Chitosan/SeNPs had greater antibacterial and antifungal activity than BSA/SeNPs and glucose/SeNPs. The MIC for Gram-positive bacteria was higher.[221]
69Na2SeO3TiO2-nanotubes/
SeNPs, AgNPs or Ag2SeNP
electrolysis of Na2SO3<10Staphylococcus epidermidisBS-Nanocomposite reduced bacterial growth and biofilm formation. In comparison with the non-modifed control, the TiO2-nanotubes/SeNPs surfaces showed a signifcantly higher coverage area with osteoblastic MG-63-cells.[177]
70Na2SeO3TiO2-nanotubes/
SeNPs
Chemical reduction in presence of TiO2-nanotubes<10Escherichia coli,
Staphylococcus aureus
BS-Samples reduced the density of E. coli by 94.6% and of S. aureus by 89.6% compared to titanium controls.[178]
71Na2SeO3polycarbonate films/SeNPsChemical reduction by glutation50–100Staphylococcus aureusBS Polycarbonate films/SeNPs inhibited bacterial growth to 8.9% and 27% when compared with an uncoated polycarbonate surface after 24 and 72 h, respectively.
BC—bactericidal effect, BS—bacteriostatic effect, FC—fungicidal effect, FS—fungistatic effect.
Se nanospheres are described in a significant portion of the studies [117,121,122,123,168]. However, other forms of SeNPs are described in a number of works: nanowires, nanorods and nanotubes. The form of SeNPs depends on the method and conditions of synthesis (pH, the presence and nature of the conjugate) [112,222]. Nanowires have bacteriostatic activity, but their MIC is comparable to or ~4–16 times lower than that of spherical SeNPs [112,222]. Comparable or higher MICs against bacteria and fungi were also found for Se nanorods compared with spherical SeNPs [112,222]. The literature also describes selenium-containing nanotubular structures with antibacterial activity [221,222]. However, it is difficult to compare the data of these works with the rest, since it is not possible to accurately estimate the final concentration of SeNPs in the obtained composites. The works devoted to the synthesis of “true” Se nanotubes are few and describe their potential application in technology, for example, in the creation of photosensors or solar cells; there are practically no works on antimicrobial applications of Se nanotubes [223].
Using regression analysis, it was found that the smaller the size of the selenium nanoparticles, the lower the necessary concentration of selenium nanoparticles for effective inhibition of bacterial growth (Figure 6b–d). Graphs are described by the following equations: y = 0.35x + 31.75 for the bacterium E. coli; y = 1.8x + 4.55 for the bacterium B. subtilis; and y = 0.19x + 34.65 for the bacterium S. aureus. For the bacterium B. subtilis, the regression coefficient was 1.8 (Figure 6c); for the bacterium E. coli, it was 0.35 (Figure 6b); and for the bacterium S. aureus, the lowest value was 0.19 (Figure 6d). In other words, a more efficient relationship between size and MIC was observed for B. subtilis than for E. coli and S. aureus.
A number of studies have shown that SeNPs have different antibacterial activities against Gram-positive and Gram-negative bacteria. A more pronounced antibacterial effect of SeNPs against Gram-negative bacteria compared to Gram-positive bacteria was shown. This property is of particular interest because, at present, among the bacteria that cause bacteremia, including sepsis, the proportion of Gram-negative bacteria is significantly increasing [224].

3.3. Dependence of the Minimum Inhibitory Concentration of Selenium Nanoparticles on Their Size in the Study of Antifungal Activity

Data were analyzed not only on the antiviral and antibacterial but also on the antifungal effect of selenium nanoparticles. Many more research papers have been devoted to the antifungal effect of selenium nanoparticles than to the antiviral effect; however, this effect is less than the antibacterial effect. In most fungal studies, the following species are used: Candida [126,150,202,209,210,211,212] and Fusarium [210,213,214,215]. Apart from these, the following fungal species are used: Colletotrichum [214,220], Puccinia [217], Aspergillus [127,216], Cryptococcus [150], Penicillium [218], Rhizoctonia [219], Pyricularia [220] and Alternaria [220]. Data on articles taken for analysis are also presented in Table 2.
Figure 7 shows the dependence of the minimum inhibitory concentration on the size of the nanoparticles. Using regression analysis, we can conclude that the smaller the size of the nanoparticles, the lower the value of the minimum inhibitory concentration (dependence equation y = 1.39x − 40.11). The graph shows the methods by which selenium nanoparticle synthesis was obtained. Most of the analyzed results use nanoparticles obtained by biological synthesis. We have presented research papers that use physical synthesis methods (laser ablation, gamma irradiation) to obtain selenium nanoparticles to study their antifungal effect. The nanoparticles used in these articles are usually spherical, and their size varies from 20 to 130 nm.

4. Mechanisms of Selenium Nanoparticle Antimicrobial Action

We decided to present the mechanisms of the antibacterial action of selenium nanoparticles in the form of a list (see below) and in the form of a diagram (Figure 8a).
(1)
Degradation of proteins due to the bactericidal action of selenium nanoparticles [112].
(2)
Slow emission of selenium ions from the surface of nanoparticles can lead to their interaction with -SH, -NH or -COOH functional groups of proteins and enzymes and the subsequent loss of their tertiary and quaternary structure and functions [125].
(3)
SeNPs contribute to the inactivation of the natural mechanisms of membrane transport of ions and nutrients through the cell walls, which blocks the vital activity of the cell [225].
(4)
Hyperproduction of ROS, disturbance of membrane potential, and depletion of internal ATP [199].
(5)
Inhibition of the activity of the dehydrogenase enzyme, as well as destruction of the integrity of the cell membrane [125].
(6)
Inhibition of the ability of bacteria to attach to the surface and form bacterial films [146].
(7)
Photocatalytic action against bacteria [226].
Thus, SeNPs can potentially be candidates for antibacterial substitutes and additives against antibiotic-resistant bacteria. Antimicrobial NPs can damage bacterial cells through multiple pathways. This multimodal antimicrobial behavior makes nanoparticles attractive, as bacteria are expected to have difficulty developing resistance to multiple forms of attack [227]. Researchers have also found that the viability of eukaryotic cells is preserved under the effective antibacterial action of selenium NPs [199].
In the case of the bacteriostatic action of selenium nanoparticles, the activity of the dehydrogenase enzyme is inhibited, and the integrity of the cell membrane is destroyed. This effect was observed when using selenium nanoparticles stabilized with arabinogalactan polysaccharide [125]. Researchers suggest possible mechanisms of the antibacterial action of selenium nanoparticles, which are triggered by the contact of the nanoparticle with a living cell.
In particular, this is the hyperproduction of ROS on the surface of nanoparticles, followed by a cascade of the LPO (lipid peroxidase) reaction, damage to cell membranes and organelles, blocking of the transcriptional gene and activation of apoptosis genes, as well as impaired synthesis of a number of cellular proteins and enzymes. In addition, the adhesion of nanoparticles on the cell surface can be accompanied by depolarization of the cell membrane, destruction of its integrity and, subsequently, cell death [125]. In the case of the combined bacteriostatic and bactericidal action of selenium nanoparticles, the ability of bacteria to attach to the surface and form bacterial films is presumably inhibited. This conclusion was made in the study of selenium nanoparticles with a silver shell [146]. For convenience, some mechanisms are shown schematically in Figure 8a,d.
Antifungal mechanisms includes antibiofilm activity [221,228], ROS generation and oxidative stress (with the addition of antifungal drug ketoconazole) [229] and influence on expression of fungicidal drug resistance genes [230] (Figure 8b).
The antiviral action of SeNPs is realized through several mechanisms: disruption of the functioning of viral capside proteins (in particular, hemagglutinin and neuraminidase activities of influenza virus), blocking of the virus-induced activation of the AKT-p52-Caspase3-dependent proapoptotic pathway, inhibition of viral replication in the host cell and enhancement of the action of antiviral drugs [183,184,187] (Figure 8c).

5. Methods for Studying the Characteristics of Selenium Nanoparticles

To describe the physicochemical characteristics of selenium NPs, a number of methods are usually used in the analyzed literature. Basically, in all articles, data on morphology, size and elemental composition are given.
Various microscopic methods are used to characterize the morphology of NPs. The most common is transmission electron microscopy (TEM) [231], and rarely atomic force microscopy (AFM) [232], scanning tunnelling microscopy (STM) [233] or scanning electron microscopy (SEM).
To characterize the sizes of NPs, the dynamic light scattering (DLS) method is most often used [234]. The method makes it possible to measure the hydrodynamic radius of nanoparticles, that is, the size of the NPs themselves and their solvate shell. The less-commonly used CPS Disc Centrifuge method allows estimation of the size of nanoparticles; however, with sizes less than 7 nm, the procedure can take several hours. Often, the size distribution of NPs is calculated using photographs or reconstructions obtained using microscopy. Differential centrifugal sedimentation (DCS) [235], particle size mobility scanning (SMPS) [236] and ion occlusion scanning (SIOS) are relatively inexpensive methods for determining particle size based on the Coulter counting principle [237]. The rarely used nanoparticle tracking analysis (NTA) method provides information on the diffusion of nanoparticles and their size [238].
A large number of methods are used to characterize the elemental composition of NPs. It should be noted that in biological applications, there are no NPs consisting of selenium oxide, and this greatly simplifies the task, since for most nanoparticles from other elements, it is necessary to prove the absence or presence of oxides [239]. Selenium oxide is soluble in water. When the surface of a selenium nanoparticle is oxidized, selenium oxide goes into solution, and the surface again consists of selenium atoms. The process continues until the complete dissolution of the selenium NP. To characterize the chemical composition, energy dispersive spectroscopy (EDX) [240] is usually used; this is very convenient since this method is usually integrated into modern transmission electron microscopes. X-ray photoelectron spectroscopy (XPS) is also often used [241]. In addition, the crystal structure of nanoparticles is often studied using the X-ray diffraction (XRD) method [242]. Selenium nanoparticles with impurities and conjugates are usually characterized with absorption spectroscopy in the UV–visible region of the spectrum [243] and Fourier transform IR spectroscopy (FTIR) [244,245].
Sometimes, differential scanning calorimetry and the Brunauer–Emmett–Teller (BET) method [246] are used to characterize NPs; these methods are used to study the surface area and rheological properties of NPs. Modulation interference microscopy (MIM) is used to study the spatial distribution of nanoparticles inside a polymer matrix [247]. The stability of NP colloids in a solvent is studied by measuring the zeta potential [227].

6. Cytotoxicity to Eukaryotic Cells

In addition to effective antimicrobial activity, it is also important to determine the safety of selenium nanoparticles for eukaryotic cells. This is a fundamental point that will allow the use of selenium nanoparticles in clinical practice. Zeraatkar et al. proved that selenium nanoparticles do not exhibit cytotoxicity to mouse fibroblasts (3T3 cell line) up to a concentration of 64 μg/mL, while the minimum inhibitory concentration is 4–8 μg/mL for Pseudomonas aeruginosa and Acinetobacter baumannii bacteria [117]. In another work by Jason Hou et al., it was shown that at concentrations from 2 to 16 μg/mL, no cytotoxicity was observed for osteoblast precursor cells (MC3T3-E1 osteoblast precursor cell line), which is important, while at concentrations of 4 μg/mL and more, the growth of the bacterium Porphyromonas gingivalis was inhibited [114].
In a published article [191], using selenium nanoparticles synthesized by a biogenic method, it was found that the inhibitory concentrations (IC50 is the concentration sufficient to inhibit the viability of 50% of cells) of SeNPs were 233.08 and 849.21 µg/mL for normal kidney cells and liver cells, respectively. Thus, normal liver cells showed greater viability to selenium nanoparticles compared with kidney cells. Importantly, the minimum inhibitory concentration of such nanoparticles against bacteria of the genera Salmonella, Klebsiella and Escherichia is 25–200 µg/mL, which is much lower than the concentration of cytotoxicity for the studied normal cell lines.
Importantly, the potential application of selenium nanoparticles lies precisely in their selective cytotoxicity; that is, nanoparticles show effective cytotoxicity for cancer cells and are safe for normal cells. For example, an article [131] was published in which the effective concentration for the antiproliferative activity of HeLa cells is 5.5 µg/mL, while the MIC for E. coli and S. aureus bacteria is 50 µg/mL. Thus, we can speak of effective inhibition of both the growth of cancerous and bacterial cells.
Moreover, a study was conducted and published in our laboratory that showed that selenium nanoparticles can have a cytoprotective effect on neuroglial cells of the cerebral cortex during ischemia/reoxygenation [48]. At concentrations as low as 3 µg/mL, selenium nanoparticles inhibit the hyperproduction of ROS in cells. The experiments also used a complex of selenium NPs with taxifolin (a flavonoid that lacks the “high dosage effect” that occurs in selenium NPs), while at a concentration of 10 µg/mL, the Se-TAX complex inhibits ROS hyperproduction and does not have toxicity on brain cells [48].
SeNPs are less toxic in vivo than other organic and inorganic (selenite) sources of Se [248]. This fact makes SeNPs attractive for biomedical applications. However, a number of experiments have shown that concentrations of SeNPs above 2 mg/kg can cause Se toxicity in mammals. [249]. Manifestations of toxicity at concentrations below 5–30 mg/kg according to the WOS standard require the substance to be classified as Toxicity class I–II. In addition, SeNPs (5 μM) exhibit acute toxicity to marine unicellular algae, which makes products based on SeNPs potentially hazardous to the environment [250]. SeNPs have comparable or even greater toxicity against microorganisms than metal NPs, for example, CuNPs [251]. The average MIC/MBC CuNPs against different strains of Escherichia coli and Staphylococcus aureus are ~210/235 and ~140/160 μg/mL, respectively [252]. For SeNPs, these values can reach <10 μg/mL and <20 μg/mL for MIC and MBC, respectively [114]. The inhibitory concentration for fungi of CuNPs is 13–22 μg/mL [253], which is comparable to or higher than that of SeNPs [150,209,212]. The toxic concentration of CuNPs for animals is 200 μg/kg, which is significantly higher than that of SiNPs, but lower than that of SiNPs [110,111,249]. CuNPs accumulate in the liver, inhibit CYP450 enzymes, and cause activation of pro-inflammatory reactions through the signaling pathway of NF-κB, MAPK, and STAT5 [254]. Thus, SeNPs have more pronounced antimicrobial properties than CuNPs, but may be more toxic against eukaryotes. These facts require caution in further biomedical developments using SeNPs.

7. Biomedical Applications

SeNPs can be used not only as antibacterial agents, but also as therapeutic agents for various pathologies. A detailed review of the biomedical applications of SeNPs can be found in [130,255]. Briefly, potential medical applications of SeNPs include cancer therapy (lung carcinoma, breast, liver, bone or kidney cancer, melanoma, etc.) [156,160,256,257,258,259], protection against poisoning (arsenic, patulin) [260,261], consequences of oxidative stress [157,256], therapy of diabetes mellitus I (oral delivery of insulin, relief of the consequences of hyperglycemia) [262,263,264] and Alzheimer’s disease [265]. In addition to the antimicrobial action, SeNPs have antiparasitic activity, which has been shown against protozoa (Leishmania, Promastigote, Giardia duodenalis, Toxoplasma) [266,267,268,269], tapeworms (Cystic echinococcus, Echinococcus granulosus) [270,271] and roundworms (Trichinella) [272].

8. Conclusions

Today, the problem of antimicrobial resistance is acute. A potential approach that can solve this problem is the use of selenium nanoparticles for antimicrobial activity against viruses, bacteria and fungi. That is why, in recent years, the interest of the scientific community in this topic has grown significantly. Selenium nanoparticles are especially attractive because of their relatively simple and inexpensive synthesis, as well as their low cytotoxicity for eukaryotic cells. Unfortunately, the chronic cytotoxicity of SeNPs has been little studied; therefore, in the future, it is necessary to carefully and thoroughly investigate this issue. It is expected that the development of scientific research in this area will effectively solve the problem of antimicrobial resistance in the near future. We consider that, in practical terms, it will be especially interesting to develop nanocoatings and nanocomposites with antibacterial properties based on SeNPs. It is probably worth focusing especially on antiviral research due to its potentially high threat. Recently, the number of publications on antiviral properties of SeNPs has been growing.

Author Contributions

Conceptualization, S.V.G. and R.L.; formal analysis, V.V.K.; investigation, V.V.K.; resources, S.V.G.; data curation, R.L.; writing—original draft preparation, V.V.K. and D.A.S.; writing—review and editing, V.V., R.L. and S.V.G.; visualization, V.V.K. and D.A.S.; supervision, S.V.G.; project administration, S.V.G.; funding acquisition, S.V.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a grant of the Ministry of Science and Higher Education of the Russian Federation (075-15-2022-315) for the organization and development of a world-class research center “Photonics”.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors, without undue reservation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Minasyan, H. Sepsis: Mechanisms of bacterial injury to the patient. Scand. J. Trauma Resusc. Emerg. Med. 2019, 27, 19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Lee, C.-R.; Lee, J.H.; Park, K.S.; Kim, Y.B.; Jeong, B.C.; Lee, S.H. Global Dissemination of Carbapenemase-Producing Klebsiella pneumoniae: Epidemiology, Genetic Context, Treatment Options, and Detection Methods. Front. Microbiol. 2016, 7, 895. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Rozgonyi, F.; Valenta, B.; Brátovics, A.; Csire, B. The sensitivity of “polyresistant” microorganisms to new antibiotics. Changes in the resistance to antibiotics of the more important pathogenic bacteria isolated from clinical specimens during 1962–1965. Orvosi Hetil. 1967, 108, 337–342. [Google Scholar]
  4. Abraham, E.P.; Chain, E. An Enzyme from Bacteria able to Destroy Penicillin. Nature 1940, 146, 837. [Google Scholar] [CrossRef]
  5. Kozlov, A.V.; Gusyakova, O.A.; Lyamin, A.V.; Kezko, J.L.; Khaliulin, A.V.; Ereshchenko, A.A. Polyresistent microflora in the structure of microorganisms divided from blood of patients of the general hospital. Klin. Lab. Diagn. 2018, 63, 574–578. [Google Scholar]
  6. Kirk, M.D.; Pires, S.M.; Black, R.E.; Caipo, M.; Crump, J.A.; Devleesschauwer, B.; Döpfer, D.; Fazil, A.; Fischer-Walker, C.L.; Hald, T.; et al. World Health Organization Estimates of the Global and Regional Disease Burden of 22 Foodborne Bacterial, Protozoal, and Viral Diseases, 2010: A Data Synthesis. PLoS Med. 2015, 12, e1001921. [Google Scholar] [CrossRef] [Green Version]
  7. Rowe, S.Y.; Rocourt, J.R.; Shiferaw, B.; Kassenborg, H.D.; Segler, S.D.; Marcus, R.; Daily, P.J.; Hardnett, F.P.; Slutsker, L. Breast-feeding decreases the risk of sporadic salmonellosis among infants in FoodNet sites. Clin. Infect. Dis. 2004, 38 (Suppl. 3), S262–S270. [Google Scholar] [CrossRef]
  8. Batz, M.B.; Henke, E.; Kowalcyk, B. Long-term consequences of foodborne infections. Infect. Dis. Clin. N. Am. 2013, 27, 599–616. [Google Scholar] [CrossRef]
  9. Kalyoussef, S.; Feja, K.N. Foodborne Illnesses. Adv. Pediatr. 2014, 61, 287–312. [Google Scholar] [CrossRef] [PubMed]
  10. Quaglia, N.C.; Dambrosio, A. Helicobacter pylori: A foodborne pathogen? World J. Gastroenterol. 2018, 24, 3472–3487. [Google Scholar] [CrossRef]
  11. Posfay-Barbe, K.M.; Wald, E.R. Listeriosis. Semin. Fetal Neonatal Med. 2009, 14, 228–233. [Google Scholar] [CrossRef] [PubMed]
  12. Kelesidis, T.; Salhotra, A.; Fleisher, J.; Uslan, D.Z. Listeria endocarditis in a patient with psoriatic arthritis on infliximab: Are biologic agents as treatment for inflammatory arthritis increasing the incidence of Listeria infections? J. Infect. 2010, 60, 386–396. [Google Scholar] [CrossRef]
  13. Chlebicz, A.; Śliżewska, K. Campylobacteriosis, Salmonellosis, Yersiniosis, and Listeriosis as Zoonotic Foodborne Diseases: A Review. Int. J. Environ. Res. Public Health 2018, 15, 863. [Google Scholar] [CrossRef] [Green Version]
  14. Ahmad, M.; Khan, A.U. Global economic impact of antibiotic resistance: A review. J. Glob. Antimicrob. Resist. 2019, 19, 313–316. [Google Scholar] [CrossRef]
  15. Gebreyes, W.A.; Thakur, S. Multidrug-Resistant Salmonella enterica Serovar Muenchen from Pigs and Humans and Potential Interserovar Transfer of Antimicrobial Resistance. Antimicrob. Agents Chemother. 2005, 49, 503–511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Endimiani, A.; Hujer, K.M.; Hujer, A.M.; Bertschy, I.; Rossano, A.; Koch, C.; Gerber, V.; Francey, T.; Bonomo, R.A.; Perreten, V. Acinetobacter baumannii isolates from pets and horses in Switzerland: Molecular characterization and clinical data. J. Antimicrob. Chemother. 2011, 66, 2248–2254. [Google Scholar] [CrossRef]
  17. Fodor, A.; Varga, I.; Hevesi, M.; Mathe-Fodor, A.; Racsko, J.; Hogan, A.J. Novel Anti-Microbial Peptides of Xenorhabdus Origin against Multidrug Resistant Plant Pathogens. In A Search for Antibacterial Agents; InTech: Singapore, 2012. [Google Scholar]
  18. Ahmed, S.A.; Barış, E.; Go, D.S.; Lofgren, H.; Osorio-Rodarte, I.; Thierfelder, K. Assessing the global poverty effects of antimicrobial resistance. World Dev. 2018, 111, 148–160. [Google Scholar] [CrossRef]
  19. Aziz, R.K.; Bhullar, K.; Waglechner, N.; Pawlowski, A.; Koteva, K.; Banks, E.D.; Johnston, M.D.; Barton, H.A.; Wright, G.D. Antibiotic Resistance Is Prevalent in an Isolated Cave Microbiome. PLoS ONE 2012, 7, e34953. [Google Scholar] [CrossRef]
  20. Blair, J.M.A.; Richmond, G.E.; Piddock, L.J.V. Multidrug efflux pumps in Gram-negative bacteria and their role in antibiotic resistance. Future Microbiol. 2014, 9, 1165–1177. [Google Scholar] [CrossRef]
  21. Cowen, L.E.; Sanglard, D.; Howard, S.J.; Rogers, P.D.; Perlin, D.S. Mechanisms of Antifungal Drug Resistance. Cold Spring Harb. Perspect. Med. 2015, 5, a019752. [Google Scholar] [CrossRef]
  22. Brown, G.D.; Denning, D.W.; Gow, N.A.R.; Levitz, S.M.; Netea, M.G.; White, T.C. Hidden Killers: Human Fungal Infections. Sci. Transl. Med. 2012, 4, 165rv13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Tian, L.; Pang, Z.; Li, M.; Lou, F.; An, X.; Zhu, S.; Song, L.; Tong, Y.; Fan, H.; Fan, J. Molnupiravir and Its Antiviral Activity against COVID-19. Front. Immunol. 2022, 13, 855496. [Google Scholar] [CrossRef]
  24. Gudkov, S.V.; Burmistrov, D.E.; Serov, D.A.; Rebezov, M.B.; Semenova, A.A.; Lisitsyn, A.B. Do Iron Oxide Nanoparticles Have Significant Antibacterial Properties? Antibiotics 2021, 10, 884. [Google Scholar] [CrossRef]
  25. Gudkov, S.V.; Burmistrov, D.E.; Serov, D.A.; Rebezov, M.B.; Semenova, A.A.; Lisitsyn, A.B. A Mini Review of Antibacterial Properties of ZnO Nanoparticles. Front. Phys. 2021, 9, 641481. [Google Scholar] [CrossRef]
  26. Gudkov, S.V.; Burmistrov, D.E.; Smirnova, V.V.; Semenova, A.A.; Lisitsyn, A.B. A Mini Review of Antibacterial Properties of Al2O3 Nanoparticles. Nanomaterials 2022, 12, 2635. [Google Scholar] [CrossRef] [PubMed]
  27. Giedraitienė, A.; Ruzauskas, M.; Šiugždinienė, R.; Tučkutė, S.; Milcius, D. Antimicrobial Properties of CuO Particles Deposited on a Medical Mask. Materials 2022, 15, 7896. [Google Scholar] [CrossRef]
  28. Li, R.; Mansukhani, N.D.; Guiney, L.M.; Ji, Z.; Zhao, Y.; Chang, C.H.; French, C.T.; Miller, J.F.; Hersam, M.C.; Nel, A.E.; et al. Identification and Optimization of Carbon Radicals on Hydrated Graphene Oxide for Ubiquitous Antibacterial Coatings. ACS Nano 2016, 10, 10966–10980. [Google Scholar] [CrossRef] [Green Version]
  29. Zheng, H.; Ma, R.; Gao, M.; Tian, X.; Li, Y.-Q.; Zeng, L.; Li, R. Antibacterial applications of graphene oxides: Structure-activity relationships, molecular initiating events and biosafety. Sci. Bull. 2018, 63, 133–142. [Google Scholar] [CrossRef]
  30. Zheng, H.; Ji, Z.; Roy, K.R.; Gao, M.; Pan, Y.; Cai, X.; Wang, L.; Li, W.; Chang, C.H.; Kaweeteerawat, C.; et al. Engineered Graphene Oxide Nanocomposite Capable of Preventing the Evolution of Antimicrobial Resistance. ACS Nano 2019, 13, 11488–11499. [Google Scholar] [CrossRef]
  31. Pan, Y.; Zheng, H.; Li, G.; Li, Y.; Jiang, J.; Chen, J.; Xie, Q.; Wu, D.; Ma, R.; Liu, X.; et al. Antibiotic-Like Activity of Atomic Layer Boron Nitride for Combating Resistant Bacteria. ACS Nano 2022, 16, 7674–7688. [Google Scholar] [CrossRef]
  32. Xie, M.; Gao, M.; Yun, Y.; Malmsten, M.; Rotello, V.M.; Zboril, R.; Akhavan, O.; Kraskouski, A.; Amalraj, J.; Cai, X.; et al. Antibacterial Nanomaterials: Mechanisms, Impacts on Antimicrobial Resistance and Design Principles. Angew. Chem. Int. Ed. 2023, 62, e202217345. [Google Scholar] [CrossRef]
  33. Yougbaré, S.; Mutalik, C.; Okoro, G.; Lin, I.H.; Krisnawati, D.I.; Jazidie, A.; Nuh, M.; Chang, C.-C.; Kuo, T.-R. Emerging Trends in Nanomaterials for Antibacterial Applications. Int. J. Nanomed. 2021, 16, 5831–5867. [Google Scholar] [CrossRef]
  34. Yougbare, S.; Chang, T.-K.; Tan, S.-H.; Kuo, J.-C.; Hsu, P.-H.; Su, C.-Y.; Kuo, T.-R. Antimicrobial Gold Nanoclusters: Recent Developments and Future Perspectives. Int. J. Mol. Sci. 2019, 20, 2924. [Google Scholar] [CrossRef] [Green Version]
  35. Lin, M.-H.; Wang, Y.-H.; Kuo, C.-H.; Ou, S.-F.; Huang, P.-Z.; Song, T.-Y.; Chen, Y.-C.; Chen, S.-T.; Wu, C.-H.; Hsueh, Y.-H.; et al. Hybrid ZnO/chitosan antimicrobial coatings with enhanced mechanical and bioactive properties for titanium implants. Carbohydr. Polym. 2021, 257, 117639. [Google Scholar] [CrossRef]
  36. Butler, K.S.; Peeler, D.J.; Casey, B.J.; Dair, B.J.; Elespuru, R.K. Silver nanoparticles: Correlating nanoparticle size and cellular uptake with genotoxicity. Mutagenesis 2015, 30, 577–591. [Google Scholar] [CrossRef]
  37. Shuguang, W.; Lawson, R.; Ray, P.C.; Hongtao, Y. Toxic effects of gold nanoparticles on Salmonella typhimurium bacteria. Toxicol. Ind. Health 2011, 27, 547–554. [Google Scholar] [CrossRef] [Green Version]
  38. Panáček, A.; Kvítek, L.; Smékalová, M.; Večeřová, R.; Kolář, M.; Röderová, M.; Dyčka, F.; Šebela, M.; Prucek, R.; Tomanec, O.; et al. Bacterial resistance to silver nanoparticles and how to overcome it. Nat. Nanotechnol. 2017, 13, 65–71. [Google Scholar] [CrossRef]
  39. Li, X.Z.; Nikaido, H.; Williams, K.E. Silver-resistant mutants of Escherichia coli display active efflux of Ag+ and are deficient in porins. J. Bacteriol. 1997, 179, 6127–6132. [Google Scholar] [CrossRef] [Green Version]
  40. Niño-Martínez, N.; Salas Orozco, M.F.; Martínez-Castañón, G.-A.; Torres Méndez, F.; Ruiz, F. Molecular Mechanisms of Bacterial Resistance to Metal and Metal Oxide Nanoparticles. Int. J. Mol. Sci. 2019, 20, 2808. [Google Scholar] [CrossRef] [Green Version]
  41. Amaro, F.; Morón, Á.; Díaz, S.; Martín-González, A.; Gutiérrez, J.C. Metallic Nanoparticles-Friends or Foes in the Battle against Antibiotic-Resistant Bacteria? Microorganisms 2021, 9, 364. [Google Scholar] [CrossRef]
  42. Helmlinger, J.; Sengstock, C.; Groß-Heitfeld, C.; Mayer, C.; Schildhauer, T.; Köller, M.; Epple, M. Silver nanoparticles with different size and shape: Equal cytotoxicity, but different antibacterial effects. RSC Adv. 2016, 6, 18490–18501. [Google Scholar] [CrossRef] [Green Version]
  43. Napierska, D.; Thomassen, L.C.J.; Lison, D.; Martens, J.A.; Hoet, P.H. The nanosilica hazard: Another variable entity. Part. Fibre Toxicol. 2010, 7, 39. [Google Scholar] [CrossRef] [Green Version]
  44. Guo, C.; Wang, J.; Jing, L.; Ma, R.; Liu, X.; Gao, L.; Cao, L.; Duan, J.; Zhou, X.; Li, Y.; et al. Mitochondrial dysfunction, perturbations of mitochondrial dynamics and biogenesis involved in endothelial injury induced by silica nanoparticles. Environ. Pollut. 2018, 236, 926–936. [Google Scholar] [CrossRef]
  45. Farooq, A.; Whitehead, D.; Azzawi, M. Attenuation of endothelial-dependent vasodilator responses, induced by dye-encapsulated silica nanoparticles, in aortic vessels. Nanomedicine 2014, 9, 413–425. [Google Scholar] [CrossRef]
  46. Maltseva, V.N.; Gudkov, S.; Turovsky, E. Modulation of the Functional State of Mouse Neutrophils by Selenium Nanoparticles In Vivo. Int. J. Mol. Sci. 2022, 23, 13651. [Google Scholar] [CrossRef]
  47. Varlamova, E.; Plotnikov, E.; Gudkov, S.; Turovsky, E. Size-Dependent Cytoprotective Effects of Selenium Nanoparticles during Oxygen-Glucose Deprivation in Brain Cortical Cells. Int. J. Mol. Sci. 2022, 23, 7464. [Google Scholar] [CrossRef]
  48. Varlamova, E.G.; Khabatova, V.V.; Gudkov, S.V.; Plotnikov, E.Y.; Turovsky, E.A. Cytoprotective Properties of a New Nanocomplex of Selenium with Taxifolin in the Cells of the Cerebral Cortex Exposed to Ischemia/Reoxygenation. Pharmaceutics 2022, 14, 2477. [Google Scholar] [CrossRef]
  49. Varlamova, E.; Goltyaev, M.; Simakin, A.; Gudkov, S.; Turovsky, E. Comparative Analysis of the Cytotoxic Effect of a Complex of Selenium Nanoparticles Doped with Sorafenib, “Naked” Selenium Nanoparticles, and Sorafenib on Human Hepatocyte Carcinoma HepG2 Cells. Int. J. Mol. Sci. 2022, 23, 6641. [Google Scholar] [CrossRef]
  50. Malyugina, S.; Skalickova, S.; Skladanka, J.; Slama, P.; Horky, P. Biogenic Selenium Nanoparticles in Animal Nutrition: A Review. Agriculture 2021, 11, 1244. [Google Scholar] [CrossRef]
  51. Zhang, J.; Wang, X.; Xu, T. 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]
  52. Kozlov, S.V.; Staroverov, S.A.; Skvortsova, N.I.; Soldatov, D.A.; Chekunov, M.A.; Silina, E.V.; Kozlov, E.S.; Artemev, D.A. Method of Obtaining a Veterinary Drug Based on Non-Specific Immunoglobulins and Colloidal Particles of Selenium for the Correction of the Immune System. RU Patent 2798268C1, 20 June 2023. [Google Scholar]
  53. Chen, C.; Hu, H.; Li, X.; Zheng, Z.; Wang, Z.; Wang, X.; Zheng, P.; Cui, F.; Li, G.; Wang, Y.; et al. Rapid Detection of Anti-SARS-CoV-2 Antibody Using a Selenium Nanoparticle-Based Lateral Flow Immunoassay. IEEE Trans. NanoBiosci. 2022, 21, 37–43. [Google Scholar] [CrossRef]
  54. 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 A Chip 2020, 20, 4255–4261. [Google Scholar] [CrossRef]
  55. Wang, Q.; Webster, T.J. Nanostructured selenium for preventing biofilm formation on polycarbonate medical devices. J. Biomed. Mater. Res. Part A 2012, 100A, 3205–3210. [Google Scholar] [CrossRef]
  56. Wa, N.J. Method for Producing Hydrous Tissue Paper Having Antibacterial and Antifungal Functions. JP Patent 2011501977A, 12 April 2007. [Google Scholar]
  57. Webster, T.J.; Tran, P.A. Antipathogenic Surfaces Having Selenium Nanoclusters. U.S. Patent 9259005B2, 16 February 2016. [Google Scholar]
  58. Yeee, A.F.; Liang, L.; Ing, N.; Gibbs, M.; Dickson, M.N. Bactericidal Surface Patterns. U.S. Patent 10875235B2, 29 December 2020. [Google Scholar]
  59. Fang, M.; Zhang, H.; Wang, Y.; Zhang, H.; Zhang, D.; Xu, P. Biomimetic selenium nanosystems for infectious wound healing. Eng. Regen. 2023, 4, 152–160. [Google Scholar] [CrossRef]
  60. Abbaszadeh, A.; Tehmasebi-Foolad, A.; Rajabzadeh, A.; Beigi-Brojeni, N.; Zarei, L. Effects of Chitosan/Nano Selenium Biofilm on Infected Wound Healing in Rats; An Experimental Study. Bull. Emerg. Trauma 2019, 7, 284–291. [Google Scholar] [CrossRef]
  61. Huang, W.; Hu, B.; Yuan, Y.; Fang, H.; Jiang, J.; Li, Q.; Zhuo, Y.; Yang, X.; Wei, J.; Wang, X. Visible Light-Responsive Selenium Nanoparticles Combined with Sonodynamic Therapy to Promote Wound Healing. ACS Biomater. Sci. Eng. 2023, 9, 1341–1351. [Google Scholar] [CrossRef]
  62. Maiyo, F.; Singh, M. Selenium nanoparticles: Potential in cancer gene and drug delivery. Nanomedicine 2017, 12, 1075–1089. [Google Scholar] [CrossRef]
  63. Dipak, N.; Loveleen, K.; Harvinder, S.S.; Dharambeer, M.S.; Sonali, G.; Deepa, T. Application of Selenium Nanoparticles in Localized Drug Targeting for Cancer Therapy. Anti-Cancer Agents Med. Chem. 2022, 22, 2715–2725. [Google Scholar] [CrossRef]
  64. 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]
  65. 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]
  66. 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]
  67. Dong, F.; Zhang, L.; Li, R.; Feng, F.; Wang, W.; Li, D.; Xiang, Q.; Yan, P. Folic Acid-Chitosan-Nano-Selenium Tumor Targeted Drug Delivery System and Preparation Method Thereof. CN Patent 111214460A, 02 June 2020. [Google Scholar]
  68. Xia, Y.; Zhao, M.; Chen, Y.; Hua, L.; Xu, T.; Wang, C.; Li, Y.; Zhu, B. Folate-targeted selenium nanoparticles deliver therapeutic siRNA to improve hepatocellular carcinoma therapy. RSC Adv. 2018, 8, 25932–25940. [Google Scholar] [CrossRef]
  69. Zou, J.; Su, S.; Chen, Z.; Liang, F.; Zeng, Y.; Cen, W.; Zhang, X.; Xia, Y.; Huang, D. Hyaluronic acid-modified selenium nanoparticles for enhancing the therapeutic efficacy of paclitaxel in lung cancer therapy. Artif. Cells Nanomed. Biotechnol. 2019, 47, 3456–3464. [Google Scholar] [CrossRef] [Green Version]
  70. Zhang, Y.; Li, X.; Huang, Z.; Zheng, W.; Fan, C.; Chen, T. Enhancement of cell permeabilization apoptosis-inducing activity of selenium nanoparticles by ATP surface decoration. Nanomed. Nanotechnol. Biol. Med. 2013, 9, 74–84. [Google Scholar] [CrossRef]
  71. Goltyaev, M.V.; Varlamova, E.G. The Role of Selenium Nanoparticles in the Treatment of Liver Pathologies of Various Natures. Int. J. Mol. Sci. 2023, 24, 10547. [Google Scholar] [CrossRef]
  72. Varlamova, E.G.; Khabatova, V.V.; Gudkov, S.V.; Turovsky, E.A. Ca2+-Dependent Effects of the Selenium-Sorafenib Nanocomplex on Glioblastoma Cells and Astrocytes of the Cerebral Cortex: Anticancer Agent and Cytoprotector. Int. J. Mol. Sci. 2023, 24, 2411. [Google Scholar] [CrossRef]
  73. Jiang, P.I.; Cai, J.; Hua, J. Oridonin Functionalized Nanoparticles and Method of Preparation Thereof Selenium. U.S. Patent 962,423,7B2, 18 April 2017. [Google Scholar]
  74. Walsh, S.K.; Kamali, N.; McGrath, J.; Hogan, J.J.; Hanrahan, J.P. Multidimensional Application of Selenium Nanoparticles. 2023. Available online: https://glantreo.com/multidimensional-application-of-selenium-nanoparticles/ (accessed on 12 May 2023).
  75. Wu, A.; Hu, D.; Na, L.I.U.; Yu, S.; Yu, D.; Tang, Y.; Wang, Y. Trichoderma-Derived Selenium Nanoparticles Foliar Fertilizer for Reducing Crop Fungal Diseases and Toxin Contamination. U.S. Patent 10807920B2, 20 October 2020. [Google Scholar]
  76. Usmani, Z.; Kumar, A.; Tripti; Ahirwal, J.; Prasad, M.N.V. Chapter 20—Scope for Applying Transgenic Plant Technology for Remediation and Fortification of Selenium. In Transgenic Plant Technology for Remediation of Toxic Metals and Metalloids; Prasad, M.N.V., Ed.; Academic Press: Cambridge, MA, USA, 2019; pp. 429–461. [Google Scholar]
  77. Wang, Q.; Yu, Y.; Li, J.; Wan, Y.; Huang, Q.; Guo, Y.; Li, H. Effects of Different Forms of Selenium Fertilizers on Se Accumulation, Distribution, and Residual Effect in Winter Wheat–Summer Maize Rotation System. J. Agric. Food Chem. 2017, 65, 1116–1123. [Google Scholar] [CrossRef]
  78. Gudkov, S.V.; Shafeev, G.A.; Glinushkin, A.P.; Shkirin, A.V.; Barmina, E.V.; Rakov, I.I.; Simakin, A.V.; Kislov, A.V.; Astashev, M.E.; Vodeneev, V.A.; et al. Production and Use of Selenium Nanoparticles as Fertilizers. ACS Omega 2020, 5, 17767–17774. [Google Scholar] [CrossRef]
  79. Shafeev, G.; Barmina, E.; Valiullin, L.; Simakin, A.; Ovsyankina, A.; Demin, D.; Kosolapov, V.; Korshunov, A.; Denisov, R. Soil fertilizer based on selenium nanoparticles. IOP Conf. Ser. Earth Environ. Sci. 2019, 390, 012041. [Google Scholar] [CrossRef]
  80. Hak, K.; Jong, L.; Hyo, K.; Gwang, L.; Jun, P.; Chan, L.; St, Y. Method for Cultivating High Quality and Functional Vegetable Fruit. KR Patent 101120635B1, 16 March 2012. [Google Scholar]
  81. Fouda, A.; Al-Otaibi, W.A.; Saber, T.; AlMotwaa, S.M.; Alshallash, K.S.; Elhady, M.; Badr, N.F.; Abdel-Rahman, M.A. Antimicrobial, Antiviral, and In-Vitro Cytotoxicity and Mosquitocidal Activities of Portulaca oleracea-Based Green Synthesis of Selenium Nanoparticles. J. Funct. Biomater. 2022, 13, 157. [Google Scholar] [CrossRef]
  82. Ahmed, F.; Dwivedi, S.; Shaalan, N.M.; Kumar, S.; Arshi, N.; Alshoaibi, A.; Husain, F.M. Development of Selenium Nanoparticle Based Agriculture Sensor for Heavy Metal Toxicity Detection. Agriculture 2020, 10, 610. [Google Scholar] [CrossRef]
  83. Dumore, N.S.; Mukhopadhyay, M. Sensitivity enhanced SeNPs-FTO electrochemical sensor for hydrogen peroxide detection. J. Electroanal. Chem. 2020, 878, 114544. [Google Scholar] [CrossRef]
  84. Mostafavi, E.; Medina-Cruz, D.; Truong, L.B.; Kaushik, A.; Iravani, S. Selenium-based nanomaterials for biosensing applications. Mater. Adv. 2022, 3, 7742–7756. [Google Scholar] [CrossRef] [PubMed]
  85. Amani, H.; Habibey, R.; Shokri, F.; Hajmiresmail, S.J.; Akhavan, O.; Mashaghi, A.; Pazoki-Toroudi, H. Selenium nanoparticles for targeted stroke therapy through modulation of inflammatory and metabolic signaling. Sci. Rep. 2019, 9, 6044. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. El-Ghazaly, M.A.; Fadel, N.; Rashed, E.; El-Batal, A.; Kenawy, S.A. Anti-inflammatory effect of selenium nanoparticles on the inflammation induced in irradiated rats. Can. J. Physiol. Pharmacol. 2017, 95, 101–110. [Google Scholar] [CrossRef] [PubMed]
  87. Gao, X.; Zhang, J.; Zhang, L. Hollow Sphere Selenium Nanoparticles: Their In-Vitro Anti Hydroxyl Radical Effect. Adv. Mater. 2002, 14, 290–293. [Google Scholar] [CrossRef]
  88. Torres, S.K.; Campos, V.L.; León, C.G.; Rodríguez-Llamazares, S.M.; Rojas, S.M.; González, M.; Smith, C.; Mondaca, M.A. Biosynthesis of selenium nanoparticles by Pantoea agglomerans and their antioxidant activity. J. Nanoparticle Res. 2012, 14, 1236. [Google Scholar] [CrossRef]
  89. Ahmed, H.H.; Abd El-Maksoud, M.D.; Abdel Moneim, A.E.; Aglan, H.A. Pre-Clinical Study for the Antidiabetic Potential of Selenium Nanoparticles. Biol. Trace Elem. Res. 2016, 177, 267–280. [Google Scholar] [CrossRef]
  90. Gao, X.; Sun, Y. Selenium Nanoparticles with Improved Biological Effects. U.S. Patent 844,502,6B2, 21 May 2013. [Google Scholar]
  91. Peng, Y.; Peng, L.; Liu, T. Protein-Bound Nano-Selenium and Preparation Method and Application Thereof. CN Patent 108484715B, 11 March 2022. [Google Scholar]
  92. Selvarajan, V.; Obuobi, S.; Ee, P.L.R. Silica Nanoparticles—A Versatile Tool for the Treatment of Bacterial Infections. Front. Chem. 2020, 8, 602. [Google Scholar] [CrossRef]
  93. Jafari, S.; Derakhshankhah, H.; Alaei, L.; Fattahi, A.; Varnamkhasti, B.S.; Saboury, A.A. Mesoporous silica nanoparticles for therapeutic/diagnostic applications. Biomed. Pharmacother. 2019, 109, 1100–1111. [Google Scholar] [CrossRef]
  94. Stöber, W.; Fink, A.; Bohn, E. Controlled growth of monodisperse silica spheres in the micron size range. J. Colloid Interface Sci. 1968, 26, 62–69. [Google Scholar] [CrossRef]
  95. Zulfiqar, U.; Subhani, T.; Wilayat Husain, S. Synthesis of silica nanoparticles from sodium silicate under alkaline conditions. J. Sol-Gel Sci. Technol. 2016, 77, 753–758. [Google Scholar] [CrossRef]
  96. Kuddus, A.; Islam, R.; Tabassum, S.; Ismail, A.B. Synthesis of Si NPs from River Sand Using the Mechanochemical Process and its Applications in Metal Oxide Heterojunction Solar Cells. Silicon 2019, 12, 1723–1733. [Google Scholar] [CrossRef]
  97. Zhou, Y.; Quan, G.; Wu, Q.; Zhang, X.; Niu, B.; Wu, B.; Huang, Y.; Pan, X.; Wu, C. Mesoporous silica nanoparticles for drug and gene delivery. Acta Pharm. Sin. B 2018, 8, 165–177. [Google Scholar] [CrossRef]
  98. Zhang, Y.; Wang, J.; Bai, X.; Jiang, T.; Zhang, Q.; Wang, S. Mesoporous Silica Nanoparticles for Increasing the Oral Bioavailability and Permeation of Poorly Water Soluble Drugs. Mol. Pharm. 2012, 9, 505–513. [Google Scholar] [CrossRef]
  99. Makarovsky, I.; Boguslavsky, Y.; Alesker, M.; Lellouche, J.; Banin, E.; Lellouche, J.-P. Novel Triclosan-Bound Hybrid-Silica Nanoparticles and their Enhanced Antimicrobial Properties. Adv. Funct. Mater. 2011, 21, 4295–4304. [Google Scholar] [CrossRef]
  100. Wu, S.-H.; Lin, Y.-S.; Hung, Y.; Chou, Y.-H.; Hsu, Y.-H.; Chang, C.; Mou, C.-Y. Multifunctional Mesoporous Silica Nanoparticles for Intracellular Labeling and Animal Magnetic Resonance Imaging Studies. ChemBioChem 2008, 9, 53–57. [Google Scholar] [CrossRef]
  101. Pandey, S.; Mewada, A.; Thakur, M.; Pillai, S.; Dharmatti, R.; Phadke, C.; Sharon, M. Synthesis of mesoporous silica oxide/C-dot complex (meso-SiO2/C-dots) using pyrolysed rice husk and its application in bioimaging. RSC Adv. 2014, 4, 1174–1179. [Google Scholar] [CrossRef]
  102. Li, Z.; Zhang, Y.; Wu, X.; Wu, X.; Maudgal, R.; Zhang, H.; Han, G. In Vivo Repeatedly Charging Near-Infrared-Emitting Mesoporous SiO2/ZnGa2O4:Cr3+ Persistent Luminescence Nanocomposites. Adv. Sci. 2015, 2, 1500001. [Google Scholar] [CrossRef] [Green Version]
  103. Chen, Z.; Tan, Y.; Xu, K.; Zhang, L.; Qiu, B.; Guo, L.; Lin, Z.; Chen, G. Stimulus-response mesoporous silica nanoparticle-based chemiluminescence biosensor for cocaine determination. Biosens. Bioelectron. 2016, 75, 8–14. [Google Scholar] [CrossRef]
  104. Bai, Y.; Yang, H.; Yang, W.; Li, Y.; Sun, C. Gold nanoparticles-mesoporous silica composite used as an enzyme immobilization matrix for amperometric glucose biosensor construction. Sens. Actuators B Chem. 2007, 124, 179–186. [Google Scholar] [CrossRef]
  105. Moon, J.H.; McDaniel, W.; Hancock, L.F. Facile fabrication of poly(p-phenylene ethynylene)/colloidal silica composite for nucleic acid detection. J. Colloid Interface Sci. 2006, 300, 117–122. [Google Scholar] [CrossRef]
  106. Boora, R.; Sheoran, P.; Rani, N.; Kumari, S.; Thakur, R.; Grewal, S. Biosynthesized Silica Nanoparticles (Si NPs) Helps in Mitigating Drought Stress in Wheat Through Physiological Changes and Upregulation of Stress Genes. Silicon 2023, 1–13. [Google Scholar] [CrossRef]
  107. Li, L.-L.; Wang, H. Enzyme-Coated Mesoporous Silica Nanoparticles as Efficient Antibacterial Agents In Vivo. Adv. Healthc. Mater. 2013, 2, 1351–1360. [Google Scholar] [CrossRef]
  108. González, B.; Colilla, M.; Díez, J.; Pedraza, D.; Guembe, M.; Izquierdo-Barba, I.; Vallet-Regí, M. Mesoporous silica nanoparticles decorated with polycationic dendrimers for infection treatment. Acta Biomater. 2018, 68, 261–271. [Google Scholar] [CrossRef]
  109. Díaz-García, D.; Ardiles, P.R.; Díaz-Sánchez, M.; Mena-Palomo, I.; del Hierro, I.; Prashar, S.; Rodríguez-Diéguez, A.; Páez, P.L.; Gómez-Ruiz, S. Copper-functionalized nanostructured silica-based systems: Study of the antimicrobial applications and ROS generation against gram positive and gram negative bacteria. J. Inorg. Biochem. 2020, 203, 110912. [Google Scholar] [CrossRef]
  110. Kim, M.; Park, J.-H.; Jeong, H.; Hong, J.; Choi, W.S.; Lee, B.-H.; Park, C.Y. An Evaluation of the in vivo Safety of Nonporous Silica Nanoparticles: Ocular Topical Administration versus Oral Administration. Sci. Rep. 2017, 7, 8238. [Google Scholar] [CrossRef] [Green Version]
  111. An, S.S.A.; Ryu, H.J.; Seong, N.-W.; So, B.J.; Seo, H.-S.; Kim, J.-H.; Hong, J.-S.; Park, M.-K.; Kim, M.-S.; Kim, Y.-R.; et al. Evaluation of silica nanoparticle toxicity after topical exposure for 90 days. Int. J. Nanomed. 2014, 9, 127–136. [Google Scholar] [CrossRef] [Green Version]
  112. Han, H.-W.; Patel, K.D.; Kwak, J.-H.; Jun, S.-K.; Jang, T.-S.; Lee, S.-H.; Knowles, J.C.; Kim, H.-W.; Lee, H.-H.; Lee, J.-H. Selenium Nanoparticles as Candidates for Antibacterial Substitutes and Supplements against Multidrug-Resistant Bacteria. Biomolecules 2021, 11, 1028. [Google Scholar] [CrossRef]
  113. Geoffrion, L.D.; Hesabizadeh, T.; Medina-Cruz, D.; Kusper, M.; Taylor, P.; Vernet-Crua, A.; Chen, J.; Ajo, A.; Webster, T.J.; Guisbiers, G. Naked Selenium Nanoparticles for Antibacterial and Anticancer Treatments. ACS Omega 2020, 5, 2660–2669. [Google Scholar] [CrossRef]
  114. Hou, J.; Tamura, Y.; Lu, H.-Y.; Takahashi, Y.; Kasugai, S.; Nakata, H.; Kuroda, S. An In Vitro Evaluation of Selenium Nanoparticles on Osteoblastic Differentiation and Antimicrobial Properties against Porphyromonas gingivalis. Nanomaterials 2022, 12, 1850. [Google Scholar] [CrossRef] [PubMed]
  115. Zhang, L.; Li, Z.; Zhang, L.; Lei, Z.; Jin, L.; Cao, J.; Quan, C. High-Efficiency Reducing Strain for Producing Selenium Nanoparticles Isolated from Marine Sediment. Int. J. Mol. Sci. 2022, 23, 11953. [Google Scholar] [CrossRef] [PubMed]
  116. Afzal, B.; Yasin, D.; Naaz; Sami, N.; Zaki, A.; Kumar, R.; Srivastava, P.; Fatma, T. Biomedical potential of Anabaena variabilis NCCU-44 based Selenium nanoparticles and their comparison with commercial SeNPs. Sci. Rep. 2021, 11, 13507. [Google Scholar] [CrossRef] [PubMed]
  117. Zeraatkar, S.; Tahan, M.; Sadeghian, H.; Nazari, R.; Behmadi, M.; Hosseini Bafghi, M. Effect of biosynthesized selenium nanoparticles using Nepeta extract against multidrug-resistant Pseudomonas aeruginosa and Acinetobacter baumannii. J. Basic Microbiol. 2023, 63, 210–222. [Google Scholar] [CrossRef] [PubMed]
  118. 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]
  119. Martínez-Esquivias, F.; Guzmán-Flores, J.M.; Pérez-Larios, A.; González Silva, N.; Becerra-Ruiz, J.S. A Review of the Antimicrobial Activity of Selenium Nanoparticles. J. Nanosci. Nanotechnol. 2021, 21, 5383–5398. [Google Scholar] [CrossRef]
  120. Kopel, J.; Fralick, J.; Reid, T.W. The Potential Antiviral Effects of Selenium Nanoparticles and Coated Surfaces. Antibiotics 2022, 11, 1683. [Google Scholar] [CrossRef]
  121. Vahdati, M.; Tohidi Moghadam, T. Synthesis and Characterization of Selenium Nanoparticles-Lysozyme Nanohybrid System with Synergistic Antibacterial Properties. Sci. Rep. 2020, 10, 510. [Google Scholar] [CrossRef] [Green Version]
  122. Abou Elmaaty, T.; Sayed-Ahmed, K.; Mohamed Ali, R.; El-Khodary, K.; Abdeldayem, S.A. Simultaneous Sonochemical Coloration and Antibacterial Functionalization of Leather with Selenium Nanoparticles (SeNPs). Polymers 2022, 14, 74. [Google Scholar] [CrossRef]
  123. Kora, A.J. Tree gum stabilised selenium nanoparticles: Characterisation and antioxidant activity. IET Nanobiotechnol. 2018, 12, 658–662. [Google Scholar] [CrossRef]
  124. Dang-Bao, T.; Ho, T.G.-T.; Do, B.L.; Phung Anh, N.; Phan, T.D.T.; Tran, T.B.Y.; Duong, N.L.; Hong Phuong, P.; Nguyen, T. Green Orange Peel-Mediated Bioinspired Synthesis of Nanoselenium and Its Antibacterial Activity against Methicillin-Resistant Staphylococcus aureus. ACS Omega 2022, 7, 36037–36046. [Google Scholar] [CrossRef]
  125. Lesnichaya, M.; Perfileva, A.; Nozhkina, O.; Gazizova, A.; Graskova, I. Synthesis, toxicity evaluation and determination of possible mechanisms of antimicrobial effect of arabinogalactane-capped selenium nanoparticles. J. Trace Elem. Med. Biol. 2022, 69, 126904. [Google Scholar] [CrossRef]
  126. Hosseini Bafghi, M.; Darroudi, M.; Zargar, M.; Zarrinfar, H.; Nazari, R. Biosynthesis of selenium nanoparticles by Aspergillus flavus and Candida albicans for antifungal applications. Micro Nano Lett. 2021, 16, 12096. [Google Scholar] [CrossRef]
  127. Islam, S.N.; Naqvi, S.M.A.; Raza, A.; Jaiswal, A.; Singh, A.K.; Dixit, M.; Barnwal, A.; Gambhir, S.; Ahmad, A. Mycosynthesis of highly fluorescent selenium nanoparticles from Fusarium oxysporum, their antifungal activity against black fungus Aspergillus niger, and in-vivo biodistribution studies. 3 Biotech 2022, 12, 309. [Google Scholar] [CrossRef]
  128. Jadhav, A.A.; Khanna, P.K. Impact of microwave irradiation on cyclo-octeno-1,2,3-selenadiazole: Formation of selenium nanoparticles and their polymorphs. RSC Adv. 2015, 5, 44756–44763. [Google Scholar] [CrossRef]
  129. Varlamova, E.G.; Turovsky, E.A.; Blinova, E.V. Therapeutic Potential and Main Methods of Obtaining Selenium Nanoparticles. Int. J. Mol. Sci. 2021, 22, 10808. [Google Scholar] [CrossRef] [PubMed]
  130. Khanna, P.; Bisht, N.; Singh, P. Selenium Nanoparticles: A Review on Synthesis and Biomedical Applications. Mater. Adv. 2022, 3, 1415–1431. [Google Scholar] [CrossRef]
  131. Pandey, S.; Awasthee, N.; Shekher, A.; Rai, L.C.; Gupta, S.C.; Dubey, S.K. Biogenic synthesis and characterization of selenium nanoparticles and their applications with special reference to antibacterial, antioxidant, anticancer and photocatalytic activity. Bioprocess Biosyst. Eng. 2021, 44, 2679–2696. [Google Scholar] [CrossRef] [PubMed]
  132. 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. 2021, 32, 351–361. [Google Scholar] [CrossRef]
  133. Kis-Csitári, J.; Kónya, Z.; Kiricsi, I. Sonochemical Synthesis of Inorganic Nanoparticles. In Functionalized Nanoscale Materials, Devices and Systems; NATO Science for Peace and Security Series B: Physics and Biophysics; Springer: Berlin/Heidelberg, Germany, 2008; pp. 369–372. [Google Scholar]
  134. Shar, A.H.; Lakhan, M.N.; Wang, J.; Ahmed, M.; Alali, K.T.; Ahmed, R.; Ali, I.; Dayo, A.Q. Facile synthesis and characterization of selenium nanoparticles by the hydrothermal approach. Dig. J. Nanomater. Biostruct. 2019, 14, 867–872. [Google Scholar]
  135. Aditha, S.K.; Kurdekar, A.D.; Chunduri, L.A.A.; Patnaik, S.; Kamisetti, V. Aqueous based reflux method for green synthesis of nanostructures: Application in CZTS synthesis. MethodsX 2016, 3, 35–42. [Google Scholar] [CrossRef]
  136. Alhawiti, A. Citric acid-mediated green synthesis of selenium nanoparticles: Antioxidant, antimicrobial, and anticoagulant potential applications. Biomass Convers. Biorefinery 2022, 1–10. [Google Scholar] [CrossRef]
  137. 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]
  138. Hien, N.Q.; Tuan, P.D.; Phu, D.V.; Quoc, L.A.; Lan, N.T.K.; Duy, N.N.; Hoa, T.T. Gamma Co-60 ray irradiation synthesis of dextran stabilized selenium nanoparticles and their antioxidant activity. Mater. Chem. Phys. 2018, 205, 29–34. [Google Scholar] [CrossRef]
  139. Clifford, D.M.; Castano, C.E.; Rojas, J.V. Supported transition metal nanomaterials: Nanocomposites synthesized by ionizing radiation. Radiat. Phys. Chem. 2017, 132, 52–64. [Google Scholar] [CrossRef]
  140. Amin, B.H.; Ahmed, H.Y.; El Gazzar, E.M.; Badawy, M.M.M. Enhancement the Mycosynthesis of Selenium Nanoparticles by Using Gamma Radiation. Dose-Response 2021, 19, 15593258211059323. [Google Scholar] [CrossRef]
  141. 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]
  142. Ayyyzhy, K.; Voronov, V.; Gudkov, S.; Rakov, I.; Simakin, A.; Shafeev, G. Laser Fabrication and Fragmentation of Selenium Nanoparticles in Aqueous Media. Phys. Wave Phenom. 2019, 27, 113–118. [Google Scholar] [CrossRef]
  143. Shafeev, G.A.; Barmina, E.V.; Pimpha, N.; Rakov, I.I.; Simakin, A.V.; Sharapov, M.G.; Uvarov, O.V.; Gudkov, S.V. Laser generation and fragmentation of selenium nanoparticles in water and their testing as an additive to fertilisers. Quantum Electron. 2021, 51, 615–618. [Google Scholar] [CrossRef]
  144. Vasileiadis, T.; Dracopoulos, V.; Kollia, M.; Sygellou, L.; Yannopoulos, S.N. Synthesis of t-Te and a-Se nanospheres using continuous wave visible light. J. Nanoparticle Res. 2019, 21, 218. [Google Scholar] [CrossRef]
  145. Vorozhtsov, A.; Goncharova, D.; Gavrilenko, E.; Nemoykina, A.; Svetlichnyi, V. Antibacterial activity of zinc oxide nanoparticles obtained by pulsed laser ablation in water and air. MATEC Web Conf. 2018, 243, 00017. [Google Scholar] [CrossRef] [Green Version]
  146. Liang, X.; Zhang, S.; Gadd, G.M.; McGrath, J.; Rooney, D.W.; Zhao, Q. Fungal-derived selenium nanoparticles and their potential applications in electroless silver coatings for preventing pin-tract infections. Regen. Biomater. 2022, 9, rbac013. [Google Scholar] [CrossRef]
  147. Fouda, A.; Hassan, S.; Eid, A.; Abdel-Rahman, M.; Hamza, M. Light enhanced the antimicrobial, anticancer, and catalytic activities of selenium nanoparticles fabricated by endophytic fungal strain, Penicillium crustosum EP-1. Sci. Rep. 2022, 12, 11834. [Google Scholar] [CrossRef]
  148. Mariadoss, A.V.A.; Saravanakumar, K.; Sathiyaseelan, A.; Naveen, K.V.; Wang, M.-H. Enhancement of anti-bacterial potential of green synthesized selenium nanoparticles by starch encapsulation. Microb. Pathog. 2022, 167, 105544. [Google Scholar] [CrossRef]
  149. Hashem, A.H.; Khalil, A.M.A.; Reyad, A.M.; Salem, S.S. Biomedical Applications of Mycosynthesized Selenium Nanoparticles Using Penicillium expansum ATTC 36200. Biol. Trace Elem. Res. 2021, 199, 3998–4008. [Google Scholar] [CrossRef]
  150. Hashem, A.H.; Salem, S.S. Green and ecofriendly biosynthesis of selenium nanoparticles using Urtica dioica (stinging nettle) leaf extract: Antimicrobial and anticancer activity. Biotechnol. J. 2022, 17, 2100432. [Google Scholar] [CrossRef]
  151. 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]
  152. Shah, V.; Medina-Cruz, D.; Vernet-Crua, A.; Truong, L.B.; Sotelo, E.; Mostafavi, E.; González, M.U.; García-Martín, J.M.; Cholula-Díaz, J.L.; Webster, T.J. Pepper-Mediated Green Synthesis of Selenium and Tellurium Nanoparticles with Antibacterial and Anticancer Potential. J. Funct. Biomater. 2023, 14, 24. [Google Scholar] [CrossRef]
  153. Nikam, P.B.; Salunkhe, J.D.; Minkina, T.; Rajput, V.D.; Kim, B.S.; Patil, S.V. A review on green synthesis and recent applications of red nano Selenium. Results Chem. 2022, 4, 100581. [Google Scholar] [CrossRef]
  154. ElSaied, B.E.F.; Diab, A.M.; Tayel, A.A.; Alghuthaymi, M.A.; Moussa, S.H. Potent antibacterial action of phycosynthesized selenium nanoparticles using Spirulina platensis extract. Green Process. Synth. 2021, 10, 49–60. [Google Scholar] [CrossRef]
  155. Mulla, N.A.; Otari, S.V.; Bohara, R.A.; Yadav, H.M.; Pawar, S.H. Rapid and size-controlled biosynthesis of cytocompatible selenium nanoparticles by Azadirachta indica leaves extract for antibacterial activity. Mater. Lett. 2020, 264, 127353. [Google Scholar] [CrossRef]
  156. Pansare, A.V.; Kulal, D.K.; Shedge, A.A.; Patil, V.R. hsDNA groove binding, photocatalytic activity, and in vitro breast and colon cancer cell reducing function of greener SeNPs. Dalton Trans. 2016, 45, 12144–12155. [Google Scholar] [CrossRef] [PubMed]
  157. Vyas, J.; Rana, S. Antioxidant activity and green synthesis of selenium nanoparticles using allium sativum extract. Int. J. Phytomed. 2017, 9, 634. [Google Scholar] [CrossRef] [Green Version]
  158. 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]
  159. 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]
  160. Ali, E.N.; El-Sonbaty, S.M.; Salem, F.M. Evaluation of selenium nanoparticles as a potential chemopreventive agent against lung carcinoma. Int. J. Pharmacol. Biol. Sci. 2013, 2, 38–46. [Google Scholar]
  161. 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]
  162. Yazdi, M.H.; Mahdavi, M.; Faghfuri, E.; Faramarzi, M.A.; Sepehrizadeh, Z.; Mohammad Hassan, Z.; Gholami, M.; Shahverdi, A.R. Th1 Immune Response Induction by Biogenic Selenium Nanoparticles in Mice with Breast Cancer: Preliminary Vaccine Model. Iran. J. Biotechnol. 2015, 13, 1–9. [Google Scholar] [CrossRef] [Green Version]
  163. 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]
  164. Hariharan, N.; Al-Harbi, P.; Karuppiah, S.R. Microbial synthesis of selenium nanocomposite using Saccharomyces cerevisiae and its antimicrobial activity against pathogens causing nosocomial infection. Chalcogenide Lett. 2012, 9, 509–515. [Google Scholar]
  165. Lian, S.; Diko, C.S.; Yan, Y.; Li, Z.; Zhang, H.; Ma, Q.; Qu, Y. Characterization of biogenic selenium nanoparticles derived from cell-free extracts of a novel yeast Magnusiomyces ingens. 3 Biotech 2019, 9, 221. [Google Scholar] [CrossRef]
  166. Song, X.; Chen, Y.; Sun, H.; Liu, X.; Leng, X. Physicochemical and functional properties of chitosan-stabilized selenium nanoparticles under different processing treatments. Food Chem. 2020, 331, 127378. [Google Scholar] [CrossRef] [PubMed]
  167. Ndwandwe, B.K.; Malinga, S.P.; Kayitesi, E.; Dlamini, B.C. Solvothermal synthesis of selenium nanoparticles with polygonal-like nanostructure and antibacterial potential. Mater. Lett. 2021, 304, 130619. [Google Scholar] [CrossRef]
  168. Youssef, D.M.; Alshubaily, F.A.; Tayel, A.A.; Alghuthaymi, M.A.; Al-Saman, M.A. Application of Nanocomposites from Bees Products and Nano-Selenium in Edible Coating for Catfish Fillets Biopreservation. Polymers 2022, 14, 2378. [Google Scholar] [CrossRef]
  169. 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]
  170. Niranjan, R.; Zafar, S.; Lochab, B.; Priyadarshini, R. Synthesis and Characterization of Sulfur and Sulfur-Selenium Nanoparticles Loaded on Reduced Graphene Oxide and Their Antibacterial Activity against Gram-Positive Pathogens. Nanomaterials 2022, 12, 191. [Google Scholar] [CrossRef]
  171. Dorazilová, J.; Muchová, J.; Šmerková, K.; Kočiová, S.; Diviš, P.; Kopel, P.; Veselý, R.; Pavliňáková, V.; Adam, V.; Vojtová, L. Synergistic Effect of Chitosan and Selenium Nanoparticles on Biodegradation and Antibacterial Properties of Collagenous Scaffolds Designed for Infected Burn Wounds. Nanomaterials 2020, 10, 1971. [Google Scholar] [CrossRef]
  172. Muchová, J.; Hearnden, V.; Michlovská, L.; Vištejnová, L.; Zavaďáková, A.; Šmerková, K.; Kočiová, S.; Adam, V.; Kopel, P.; Vojtová, L. Mutual influence of selenium nanoparticles and FGF2-STAB® on biocompatible properties of collagen/chitosan 3D scaffolds: In vitro and ex ovo evaluation. J. Nanobiotechnol. 2021, 19, 103. [Google Scholar] [CrossRef]
  173. Huang, T.; Holden, J.A.; Reynolds, E.C.; Heath, D.E.; O’Brien-Simpson, N.M.; O’Connor, A.J. Multifunctional Antimicrobial Polypeptide-Selenium Nanoparticles Combat Drug-Resistant Bacteria. ACS Appl. Mater. Interfaces 2020, 12, 55696–55709. [Google Scholar] [CrossRef] [PubMed]
  174. Abou Elmaaty, T.; Sayed-Ahmed, K.; Elsisi, H.; Ramadan, S.M.; Sorour, H.; Magdi, M.; Abdeldayem, S.A. Novel Antiviral and Antibacterial Durable Polyester Fabrics Printed with Selenium Nanoparticles (SeNPs). Polymers 2022, 14, 955. [Google Scholar] [CrossRef] [PubMed]
  175. Hussein, H.G.; El-Sayed, E.-S.R.; Younis, N.A.; Hamdy, A.E.H.A.; Easa, S.M. Harnessing endophytic fungi for biosynthesis of selenium nanoparticles and exploring their bioactivities. AMB Express 2022, 12, 68. [Google Scholar] [CrossRef] [PubMed]
  176. Bilek, O.; Fohlerová, Z.; Hubalek, J. Enhanced antibacterial and anticancer properties of Se-NPs decorated TiO2 nanotube film. PLoS ONE 2019, 14, 0214066. [Google Scholar] [CrossRef] [Green Version]
  177. Staats, K.; Pilz, M.; Sun, J.; Boiadjieva-Scherzer, T.; Kronberger, H.; Tobudic, S.; Windhager, R.; Holinka, J. Antimicrobial potential and osteoblastic cell growth on electrochemically modified titanium surfaces with nanotubes and selenium or silver incorporation. Sci. Rep. 2022, 12, 8298. [Google Scholar] [CrossRef]
  178. Liu, W.; Golshan, N.H.; Deng, X.; Hickey, D.J.; Zeimer, K.; Li, H.; Webster, T.J. Selenium nanoparticles incorporated into titania nanotubes inhibit bacterial growth and macrophage proliferation. Nanoscale 2016, 8, 15783–15794. [Google Scholar] [CrossRef] [PubMed]
  179. Liu, X.; Chen, D.; Su, J.; Zheng, R.; Ning, Z.; Zhao, M.; Zhu, B.; Li, Y. Selenium nanoparticles inhibited H1N1 influenza virus-induced apoptosis by ROS-mediated signaling pathways. RSC Adv. 2022, 12, 3862–3870. [Google Scholar] [CrossRef]
  180. 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 β-Thujaplicin through Reactive Oxygen Species-Mediated AKT and p53 Signaling Pathways. ACS Omega 2020, 5, 30633–30642. [Google Scholar] [CrossRef]
  181. Lin, Z.; Li, Y.; Gong, G.; Xia, Y.; Wang, C.; Chen, Y.; Hua, L.; Zhong, J.; Tang, Y.; Liu, X.; et al. Restriction of H1N1 influenza virus infection by selenium nanoparticles loaded with ribavirin via resisting caspase-3 apoptotic pathway. Int. J. Nanomed. 2018, 13, 5787–5797. [Google Scholar] [CrossRef] [Green Version]
  182. 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]
  183. Zhong, J.; Xia, Y.; Hua, L.; Liu, X.; Xiao, M.; Xu, T.; Zhu, B.; Cao, H. Functionalized selenium nanoparticles enhance the anti-EV71 activity of oseltamivir in human astrocytoma cell model. Artif. Cells Nanomed. Biotechnol. 2019, 47, 3485–3491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Lin, Z.; Li, Y.; Xu, T.; Guo, M.; Wang, C.; Zhao, M.; Chen, H.; Kuang, J.; Li, W.; Zhang, Y.; et al. Inhibition of Enterovirus 71 by Selenium Nanoparticles Loaded with siRNA through Bax Signaling Pathways. ACS Omega 2020, 5, 12495–12500. [Google Scholar] [CrossRef]
  185. Makhlof, M.E.M.; Albalwe, F.M.; Al-Shaikh, T.M.; El-Sheekh, M.M. Suppression Effect of Ulva lactuca Selenium Nanoparticles (USeNPs) on HepG2 Carcinoma Cells Resulting from Degradation of Epidermal Growth Factor Receptor (EGFR) with an Evaluation of Its Antiviral and Antioxidant Activities. Appl. Sci. 2022, 12, 11546. [Google Scholar] [CrossRef]
  186. Touliabah, H.E.; El-Sheekh, M.M.; Makhlof, M.E.M. Evaluation of Polycladia myrica mediated selenium nanoparticles (PoSeNPS) cytotoxicity against PC-3 cells and antiviral activity against HAV HM175 (Hepatitis A), HSV-2 (Herpes simplex II), and Adenovirus strain 2. Front. Mar. Sci. 2022, 9, 1092343. [Google Scholar] [CrossRef]
  187. Li, Y.; Lin, Z.; Guo, M.; Xia, Y.; Zhao, M.; Wang, C.; Xu, T.; Chen, T.; Zhu, B. Inhibitory activity of selenium nanoparticles functionalized with oseltamivir on H1N1 influenza virus. Int. J. Nanomed. 2017, 12, 5733–5743. [Google Scholar] [CrossRef] [Green Version]
  188. Zhang, H.; Li, Z.; Dai, C.; Wang, P.; Fan, S.; Yu, B.; Qu, Y. Antibacterial properties and mechanism of selenium nanoparticles synthesized by Providencia sp. DCX. Environ. Res. 2021, 194, 110630. [Google Scholar] [CrossRef] [PubMed]
  189. Mao, L.; Wang, L.; Zhang, M.; Ullah, M.; Liu, L.; Zhao, W.; Li, Y.; Ahmed, A.A.Q.; Cheng, H.; Shi, Z.; et al. In Situ Synthesized Selenium Nanoparticles-Decorated Bacterial Cellulose/Gelatin Hydrogel with Enhanced Antibacterial, Antioxidant, and Anti-Inflammatory Capabilities for Facilitating Skin Wound Healing. Adv. Healthc. Mater. 2021, 10, 2100402. [Google Scholar] [CrossRef] [PubMed]
  190. Cremonini, E.; Boaretti, M.; Vandecandelaere, I.; Zonaro, E.; Coenye, T.; Lleo, M.M.; Lampis, S.; Vallini, G. Biogenic selenium nanoparticles synthesized by Stenotrophomonas maltophilia SeITE02 loose antibacterial and antibiofilm efficacy as a result of the progressive alteration of their organic coating layer. Microb. Biotechnol. 2018, 11, 1037–1047. [Google Scholar] [CrossRef] [Green Version]
  191. Abbas, H.S.; Abou Baker, D.H.; Ahmed, E.A. Cytotoxicity and antimicrobial efficiency of selenium nanoparticles biosynthesized by Spirulina platensis. Arch. Microbiol. 2021, 203, 523–532. [Google Scholar] [CrossRef]
  192. Pekarkova, J.; Gablech, I.; Fialova, T.; Bilek, O.; Fohlerova, Z. Modifications of Parylene by Microstructures and Selenium Nanoparticles: Evaluation of Bacterial and Mesenchymal Stem Cell Viability. Front. Bioeng. Biotechnol. 2021, 9, 782799. [Google Scholar] [CrossRef]
  193. Elakraa, A.; Salah Salem, S.; El-Sayyad, G.; Salah Attia, M. Cefotaxime incorporated bimetallic silver-selenium nanoparticles: Promising antimicrobial synergism, antibiofilm activity, and bacterial membrane leakage reaction mechanism. RSC Adv. 2022, 12, 26603–26619. [Google Scholar] [CrossRef]
  194. Galkina, K.V.; Zubareva, V.M.; Kashko, N.D.; Lapashina, A.S.; Markova, O.V.; Feniouk, B.A.; Knorre, D.A. Heterogeneity of Starved Yeast Cells in IF1 Levels Suggests the Role of This Protein in vivo. Front. Microbiol. 2022, 13, 816622. [Google Scholar] [CrossRef]
  195. Hyrslova, I.; Kaňa, A.; Kantorova, V.; Krausova, G.; Mrvikova, I.; Doskocil, I. Selenium accumulation and biotransformation in Streptococcus, Lactococcus, and Enterococcus strains. J. Funct. Foods 2022, 92, 105056. [Google Scholar] [CrossRef]
  196. Tendenedzai, J.T.; Chirwa, E.M.N.; Brink, H.G. Enterococcus spp. Cell-Free Extract: An Abiotic Route for Synthesis of Selenium Nanoparticles (SeNPs), Their Characterisation and Inhibition of Escherichia coli. Nanomaterials 2022, 12, 658. [Google Scholar] [CrossRef]
  197. Tran, P.A.; O’Brien-Simpson, N.; Reynolds, E.C.; Pantarat, N.; Biswas, D.P.; O’Connor, A.J. Low cytotoxic trace element selenium nanoparticles and their differential antimicrobial properties against S. aureus and E. coli. Nanotechnology 2016, 27, 045101. [Google Scholar] [CrossRef]
  198. Chandramohan, S.; Sundar, K.; Muthukumaran, A. Hollow selenium nanoparticles from potato extract and investigation of its biological properties and developmental toxicity in zebrafish embryos. IET Nanobiotechnol. 2019, 13, 275–281. [Google Scholar] [CrossRef]
  199. Huang, T.; Holden, J.A.; Heath, D.E.; O’Brien-Simpson, N.M.; O’Connor, A.J. Engineering highly effective antimicrobial selenium nanoparticles through control of particle size. Nanoscale 2019, 11, 14937–14951. [Google Scholar] [CrossRef] [PubMed]
  200. Chung, S.; Zhou, R.; Webster, T. Green Synthesized BSA-Coated Selenium Nanoparticles Inhibit Bacterial Growth While Promoting Mammalian Cell Growth. Int. J. Nanomed. 2020, 15, 115–124. [Google Scholar] [CrossRef] [Green Version]
  201. Huang, T.; Kumari, S.; Herold, H.; Bargel, H.; Aigner, T.; Heath, D.; O’Brien-Simpson, N.; O’Connor, A.; Scheibel, T. Enhanced Antibacterial Activity of Se Nanoparticles Upon Coating with Recombinant Spider Silk Protein eADF4(κ16). Int. J. Nanomed. 2020, 2020, 4275–4288. [Google Scholar] [CrossRef]
  202. 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]
  203. Guisbiers, G.; Wang, Q.; Khachatryan, E.; Mimun, L.; Mendoza-Cruz, R.; Larese-Casanova, P.; Webster, T.; Nash, K. 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]
  204. Shahmoradi, S.; Shariati, A.; Amini, S.M.; Zargar, N.; Yadegari, Z.; Darban-Sarokhalil, D. The application of selenium nanoparticles for enhancing the efficacy of photodynamic inactivation of planktonic communities and the biofilm of Streptococcus mutans. BMC Res. Notes 2022, 15, 84. [Google Scholar] [CrossRef]
  205. Hernández-Díaz, J.A.; Garza-García, J.J.; León-Morales, J.M.; Zamudio-Ojeda, A.; Arratia-Quijada, J.; Velázquez-Juárez, G.; López-Velázquez, J.C.; García-Morales, S. Antibacterial Activity of Biosynthesized Selenium Nanoparticles Using Extracts of Calendula officinalis against Potentially Clinical Bacterial Strains. Molecules 2021, 26, 5929. [Google Scholar] [CrossRef]
  206. Mansouri-Tehrani, H.A.; Keyhanfar, M.; Behbahani, M.; Dini, G. Synthesis and characterization of algae-coated selenium nanoparticles as a novel antibacterial agent against Vibrio harveyi, a Penaeus vannamei pathogen. Aquaculture 2021, 534, 736260. [Google Scholar] [CrossRef]
  207. Bagheri-Josheghani, S.; Bakhshi, B. Investigation of the Antibacterial and Antibiofilm Activity of Selenium Nanoparticles against Vibrio cholerae as a Potent Therapeutics. Can. J. Infect. Dis. Med. Microbiol. 2022, 2022, 3432235. [Google Scholar] [CrossRef] [PubMed]
  208. Alghuthaymi, M.A. Antibacterial action of insect chitosan/gum Arabic nanocomposites encapsulating eugenol and selenium nanoparticles. J. King Saud Univ. Sci. 2022, 34, 102219. [Google Scholar] [CrossRef]
  209. Hosseini Bafghi, M.; Zarrinfar, H.; Darroudi, M.; Zargar, M.; Nazari, R. Green synthesis of selenium nanoparticles and evaluate their effect on the expression of ERG3, ERG11 and FKS1 antifungal resistance genes in Candida albicans and Candida glabrata. Lett. Appl. Microbiol. 2022, 74, 809–819. [Google Scholar] [CrossRef] [PubMed]
  210. El-Saadony, M.T.; Saad, A.M.; Taha, T.F.; Najjar, A.A.; Zabermawi, N.M.; Nader, M.M.; AbuQamar, S.F.; El-Tarabily, K.A.; Salama, A. Selenium nanoparticles from Lactobacillus paracasei HM1 capable of antagonizing animal pathogenic fungi as a new source from human breast milk. Saudi J. Biol. Sci. 2021, 28, 6782–6794. [Google Scholar] [CrossRef]
  211. Safaei, M.; Mozaffari, H.R.; Moradpoor, H.; Imani, M.M.; Sharifi, R.; Golshah, A. Optimization of Green Synthesis of Selenium Nanoparticles and Evaluation of Their Antifungal Activity against Oral Candida albicans Infection. Adv. Mater. Sci. Eng. 2022, 2022, 1376998. [Google Scholar] [CrossRef]
  212. Lara, H.; Guisbiers, G.; Mendoza, J.; Mimun, L.; Vincent, B.; Lopez-Ribot, J.; Nash, K. 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] [Green Version]
  213. Shahbaz, M.; Akram, A.; Raja, N.I.; Mukhtar, T.; Mehak, A.; Fatima, N.; Ajmal, M.; Ali, K.; Mustafa, N.; Abasi, F. Antifungal activity of green synthesized selenium nanoparticles and their effect on physiological, biochemical, and antioxidant defense system of mango under mango malformation disease. PLoS ONE 2023, 18, e0274679. [Google Scholar] [CrossRef]
  214. Lazcano-Ramírez, H.G.; Garza-García, J.J.O.; Hernández-Díaz, J.A.; León-Morales, J.M.; Macías-Sandoval, A.S.; García-Morales, S. Antifungal Activity of Selenium Nanoparticles Obtained by Plant-Mediated Synthesis. Antibiotics 2023, 12, 115. [Google Scholar] [CrossRef]
  215. El-Saadony, M.T.; Saad, A.M.; Najjar, A.A.; Alzahrani, S.O.; Alkhatib, F.M.; Shafi, M.E.; Selem, E.; Desoky, E.-S.M.; Fouda, S.E.E.; El-Tahan, A.M.; et al. The use of biological selenium nanoparticles to suppress Triticum aestivum L. crown and root rot diseases induced by Fusarium species and improve yield under drought and heat stress. Saudi J. Biol. Sci. 2021, 28, 4461–4471. [Google Scholar] [CrossRef]
  216. Bafghi, M.H.; Nazari, R.; Darroudi, M.; Zargar, M.; Zarrinfar, H. The effect of biosynthesized selenium nanoparticles on the expression of CYP51A and HSP90 antifungal resistance genes in Aspergillus fumigatus and Aspergillus flavus. Biotechnol. Prog. 2022, 38, e3206. [Google Scholar] [CrossRef]
  217. Shahbaz, M.; Fatima, N.; Mashwani, Z.-u.-R.; Akram, A.; Haq, E.U.; Mehak, A.; Abasi, F.; Ajmal, M.; Yousaf, T.; Raja, N.I.; et al. Effect of Phytosynthesized Selenium and Cerium Oxide Nanoparticles on Wheat (Triticum aestivum L.) against Stripe Rust Disease. Molecules 2022, 27, 8149. [Google Scholar] [CrossRef]
  218. Salem, M.F.; Abd-Elraoof, W.A.; Tayel, A.A.; Alzuaibr, F.M.; Abonama, O.M. Antifungal application of biosynthesized selenium nanoparticles with pomegranate peels and nanochitosan as edible coatings for citrus green mold protection. J. Nanobiotechnol. 2022, 20, 182. [Google Scholar] [CrossRef]
  219. Hashem, A.H.; Abdelaziz, A.M.; Askar, A.A.; Fouda, H.M.; Khalil, A.M.A.; Abd-Elsalam, K.A.; Khaleil, M.M. Bacillus megaterium-Mediated Synthesis of Selenium Nanoparticles and Their Antifungal Activity against Rhizoctonia solani in Faba Bean Plants. J. Fungi 2021, 7, 195. [Google Scholar] [CrossRef] [PubMed]
  220. Joshi, S.M.; 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] [PubMed] [Green Version]
  221. Filipović, N.; Ušjak, D.; Milenković, M.T.; Zheng, K.; Liverani, L.; Boccaccini, A.R.; Stevanović, M.M. Comparative Study of the Antimicrobial Activity of Selenium Nanoparticles with Different Surface Chemistry and Structure. Front. Bioeng. Biotechnol. 2021, 8, 624621. [Google Scholar] [CrossRef] [PubMed]
  222. Tuyen, N.N.K.; Huy, V.K.; Duy, N.H.; An, H.; Nam, N.T.H.; Dat, N.M.; Huong, Q.T.T.; Trang, N.L.P.; Anh, N.D.P.; Thy, L.T.M.; et al. Green synthesis of selenium nanorods using Muntigia calabura leaf extract: Effect of pH on characterization and bioactivities. Res. Sq. 2023, 1, 1–19. [Google Scholar] [CrossRef]
  223. 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]
  224. Cheren’kaja, T.V.; Borisova, L.A.; Aleksandrova, I.V.; Kosolapov, D.A. The etiological structure of bacteremia and sepsis causative agents in patients with intensive care in an emergency hospital. Neotlozhnaya Meditsinskaja Pomos. 2013, 16, 33–36. (In Russian) [Google Scholar]
  225. Lesnichaya, M.V.; Malysheva, S.F.; Belogorlova, N.A.; Graskova, I.A.; Gazizova, A.V.; Perfilyeva, A.I.; Nozhkina, O.A.; Sukhov, B.G. Synthesis and antimicrobial activity of arabinogalactan-stabilized selenium nanoparticles from sodium bis(2-phenylethyl)diselenophosphinate. Russ. Chem. Bull. 2019, 68, 2245–2251. [Google Scholar] [CrossRef]
  226. Sahoo, B.; Leena Panigrahi, L.; Jena, S.; Jha, S.; Arakha, M. Oxidative stress generated due to photocatalytic activity of biosynthesized selenium nanoparticles triggers cytoplasmic leakage leading to bacterial cell death. RSC Adv. 2023, 13, 11406–11414. [Google Scholar] [CrossRef] [PubMed]
  227. Mühling, M.; Bradford, A.; Readman, J.W.; Somerfield, P.J.; Handy, R.D. An investigation into the effects of silver nanoparticles on antibiotic resistance of naturally occurring bacteria in an estuarine sediment. Mar. Environ. Res. 2009, 68, 278–283. [Google Scholar] [CrossRef] [Green Version]
  228. 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] [Green Version]
  229. Garg, A.; Singh, C.; Pradhan, D.; Ghosh, G.; Rath, G. Topical application of nanoparticles integrated supramolecular hydrogels for the potential treatment of seborrhoeic dermatitis. Pharm. Dev. Technol. 2020, 25, 748–756. [Google Scholar] [CrossRef]
  230. Parsamehr, N.; Rezaie, S.; Khodavaisy, S.; Salari, S.; Hadizadeh, S.; Kord, M.; Ayatollahi Mousavi, S.A. Effect of biogenic selenium nanoparticles on ERG11 and CDR1 gene expression in both fluconazole-resistant and -susceptible Candida albicans isolates. Curr. Med. Mycol. 2017, 3, 16–20. [Google Scholar] [CrossRef] [Green Version]
  231. Nagajyothi, P.C.; Sreekanth, T.V.M.; Tettey, C.O.; Jun, Y.I.; Mook, S.H. Characterization, antibacterial, antioxidant, and cytotoxic activities of ZnO nanoparticles using Coptidis Rhizoma. Bioorganic Med. Chem. Lett. 2014, 24, 4298–4303. [Google Scholar] [CrossRef]
  232. Rao, K.; Imran, M.; Jabri, T.; Ali, I.; Perveen, S.; Shafiullah; Ahmed, S.; Shah, M.R. Gum tragacanth stabilized green gold nanoparticles as cargos for Naringin loading: A morphological investigation through AFM. Carbohydr. Polym. 2017, 174, 243–252. [Google Scholar] [CrossRef]
  233. Song, Z.; Hrbek, J.; Osgood, R. Formation of TiO2 Nanoparticles by Reactive-Layer-Assisted Deposition and Characterization by XPS and STM. Nano Lett. 2005, 5, 1327–1332. [Google Scholar] [CrossRef]
  234. Das, B.; Dash, S.K.; Mandal, D.; Ghosh, T.; Chattopadhyay, S.; Tripathy, S.; Das, S.; Dey, S.K.; Das, D.; Roy, S. Green synthesized silver nanoparticles destroy multidrug resistant bacteria via reactive oxygen species mediated membrane damage. Arab. J. Chem. 2017, 10, 862–876. [Google Scholar] [CrossRef] [Green Version]
  235. Bell, N.C.; Minelli, C.; Tompkins, J.; Stevens, M.M.; Shard, A.G. Emerging Techniques for Submicrometer Particle Sizing Applied to Stöber Silica. Langmuir 2012, 28, 10860–10872. [Google Scholar] [CrossRef] [PubMed]
  236. Ehara, K.; Sakurai, H. Metrology of airborne and liquid-borne nanoparticles: Current status and future needs. Metrologia 2010, 47, S83. [Google Scholar] [CrossRef]
  237. Henriquez, R.R.; Ito, T.; Sun, L.; Crooks, R.M. The resurgence of Coulter counting for analyzing nanoscale objects. Analyst 2004, 129, 478–482. [Google Scholar] [CrossRef]
  238. Neville, F.; Broderick, M.J.F.; Gibson, T.; Millner, P.A. Fabrication and Activity of Silicate Nanoparticles and Nanosilicate-Entrapped Enzymes Using Polyethyleneimine As a Biomimetic Polymer. Langmuir 2011, 27, 279–285. [Google Scholar] [CrossRef] [PubMed]
  239. Gudkov, S.V.; Astashev, M.E.; Baimler, I.V.; Uvarov, O.V.; Voronov, V.V.; Simakin, A.V. Laser-Induced Optical Breakdown of an Aqueous Colloidal Solution Containing Terbium Nanoparticles: The Effect of Oxidation of Nanoparticles. J. Phys. Chem. B 2022, 126, 5678–5688. [Google Scholar] [CrossRef]
  240. Singh, V.; Shrivastava, A.; Wahi, N. Biosynthesis of silver nanoparticles by plants crude extracts and their characterization using UV, XRD, TEM and EDX. Afr. J. Biotechnol. 2015, 14, 2554–2567. [Google Scholar] [CrossRef] [Green Version]
  241. Naraginti, S.; Li, Y. Preliminary investigation of catalytic, antioxidant, anticancer and bactericidal activity of green synthesized silver and gold nanoparticles using Actinidia deliciosa. J. Photochem. Photobiol. B Biol. 2017, 170, 225–234. [Google Scholar] [CrossRef]
  242. Zad, Z.R.; Davarani, S.S.H.; Taheri, A.; Bide, Y. A yolk shell Fe3O4@PA-Ni@Pd/Chitosan nanocomposite -modified carbon ionic liquid electrode as a new sensor for the sensitive determination of fluconazole in pharmaceutical preparations and biological fluids. J. Mol. Liq. 2018, 253, 233–240. [Google Scholar] [CrossRef]
  243. Mittal, A.K.; Chisti, Y.; Banerjee, U.C. Synthesis of metallic nanoparticles using plant extracts. Biotechnol. Adv. 2013, 31, 346–356. [Google Scholar] [CrossRef] [PubMed]
  244. Wang, T.; Lin, J.; Chen, Z.; Megharaj, M.; Naidu, R. Green synthesized iron nanoparticles by green tea and eucalyptus leaves extracts used for removal of nitrate in aqueous solution. J. Clean. Prod. 2014, 83, 413–419. [Google Scholar] [CrossRef]
  245. Sana, S.S.; Dogiparthi, L.K. Green synthesis of silver nanoparticles using Givotia moluccana leaf extract and evaluation of their antimicrobial activity. Mater. Lett. 2018, 226, 47–51. [Google Scholar] [CrossRef]
  246. Venkateswarlu, S.; Natesh Kumar, B.; Prasad, C.H.; Venkateswarlu, P.; Jyothi, N.V.V. Bio-inspired green synthesis of Fe3O4 spherical magnetic nanoparticles using Syzygium cumini seed extract. Phys. B Condens. Matter 2014, 449, 67–71. [Google Scholar] [CrossRef]
  247. Burmistrov, D.E.; Serov, D.A.; Simakin, A.V.; Baimler, I.V.; Uvarov, O.V.; Gudkov, S.V. A Polytetrafluoroethylene (PTFE) and Nano-Al2O3 Based Composite Coating with a Bacteriostatic Effect against E. coli and Low Cytotoxicity. Polymers 2022, 14, 4764. [Google Scholar] [CrossRef] [PubMed]
  248. Bai, K.; Hong, B.; Huang, W.; He, J. Selenium-Nanoparticles-Loaded Chitosan/Chitooligosaccharide Microparticles and Their Antioxidant Potential: A Chemical and In Vivo Investigation. Pharmaceutics 2020, 12, 43. [Google Scholar] [CrossRef] [Green Version]
  249. Shurygina, I.A.; Trukhan, I.S.; Dremina, N.N.; Shurygin, M.G. Selenium Nanoparticles. In Nanotechnology in Medicine; Wiley: Hoboken, NJ, USA, 2021; pp. 47–66. [Google Scholar]
  250. Kumar, C.M.V.; Karthick, V.; Inbakandan, D.; Kumar, V.G.; Rene, E.R.; Dhas, T.S.; Ravi, M.; Sowmiya, P.; Das, C.G.A. Effect of selenium nanoparticles induced toxicity on the marine diatom Chaetoceros gracilis. Process Saf. Environ. Prot. 2022, 163, 200–209. [Google Scholar] [CrossRef]
  251. Crisan, M.C.; Teodora, M.; Lucian, M. Copper Nanoparticles: Synthesis and Characterization, Physiology, Toxicity and Antimicrobial Applications. Appl. Sci. 2021, 12, 141. [Google Scholar] [CrossRef]
  252. Ruparelia, J.P.; Chatterjee, A.K.; Duttagupta, S.P.; Mukherji, S. Strain specificity in antimicrobial activity of silver and copper nanoparticles. Acta Biomater. 2008, 4, 707–716. [Google Scholar] [CrossRef]
  253. Hou, J.; Wang, X.; Hayat, T.; Wang, X. Ecotoxicological effects and mechanism of CuO nanoparticles to individual organisms. Environ. Pollut. 2017, 221, 209–217. [Google Scholar] [CrossRef]
  254. Tang, H.; Xu, M.; Luo, J.; Zhao, L.; Ye, G.; Shi, F.; Lv, C.; Chen, H.; Wang, Y.; Li, Y. Liver toxicity assessments in rats following sub-chronic oral exposure to copper nanoparticles. Environ. Sci. Eur. 2019, 31, 30. [Google Scholar] [CrossRef] [Green Version]
  255. Ferro, C.; Florindo, H.F.; Santos, H.A. Selenium Nanoparticles for Biomedical Applications: From Development and Characterization to Therapeutics. Adv. Healthc. Mater. 2021, 10, 2100598. [Google Scholar] [CrossRef] [PubMed]
  256. Ramamurthy, C.H.; 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]
  257. 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. 2016, 424, 13–33. [Google Scholar] [CrossRef]
  258. 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. [Google Scholar] [CrossRef]
  259. Liu, W.; Li, X.; Wong, Y.-S.; Zheng, W.; Zhang, Y.; Cao, W.; Chen, T. Selenium Nanoparticles as a Carrier of 5-Fluorouracil to Achieve Anticancer Synergism. ACS Nano 2012, 6, 6578–6591. [Google Scholar] [CrossRef]
  260. 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]
  261. Qiu, Y.; Chen, X.; Chen, Z.; Zeng, X.; Yue, T.; Yuan, Y. Effects of Selenium Nanoparticles on Preventing Patulin-Induced Liver, Kidney and Gastrointestinal Damage. Foods 2022, 11, 749. [Google Scholar] [CrossRef]
  262. 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]
  263. 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]
  264. Abdel Moneim, 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] [Green Version]
  265. Yang, L.; Wang, W.; Chen, J.; Wang, N.; Zheng, G. A comparative study of resveratrol and resveratrol-functional selenium nanoparticles: Inhibiting amyloid β aggregation and reactive oxygen species formation properties. J. Biomed. Mater. Res. Part A 2018, 106, 3034–3041. [Google Scholar] [CrossRef]
  266. Mahmoudvand, H.; Shakibaie, M.; Tavakoli, R.; Jahanbakhsh, S.; Sharifi, I. In Vitro Study of Leishmanicidal Activity of Biogenic Selenium Nanoparticles against Iranian Isolate of Sensitive and Glucantime-Resistant Leishmania tropica. Iran. J. Parasitol. 2014, 9, 452–460. [Google Scholar]
  267. Wlodkowic, D.; Telford, W.; Skommer, J.; Darzynkiewicz, Z. Apoptosis and Beyond: Cytometry in Studies of Programmed Cell Death. Methods Cell Biol. 2011, 103, 55–98. [Google Scholar]
  268. Malekifard, F.; Tavassoli, M.; Vaziri, K. In Vitro Assessment Antiparasitic Effect of Selenium and Copper Nanoparticles on Giardia deodenalis Cyst. Iran. J. Parasitol. 2020, 15, 411. [Google Scholar] [CrossRef]
  269. Arafa, F.M.; Mogahed, N.M.F.H.; Eltarahony, M.M.; Diab, R.G. Biogenic selenium nanoparticles: Trace element with promising anti-toxoplasma effect. Pathog. Glob. Health 2023, 5, 1–16. [Google Scholar] [CrossRef]
  270. Mahmoudvand, H.; Fasihi Harandi, M.; 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]
  271. Mohammed, S.; Ali, A.A. Effect of selenium nanoparticles against protoscoleces of Echinococcus granulosus in vitro and hydatid cysts in mice. Iraqi J. Vet. Sci. 2022, 36, 195–202. [Google Scholar] [CrossRef]
  272. Sarhan, M.H.; Farghaly, A.; Abd El-Aal, N.F.; Mohammed Farag, S.; Ahmed Ali, A.; Farag, T.I. Egyptian propolis and selenium nanoparticles against murine trichinosis: A novel therapeutic insight. J. Helminthol. 2022, 96, e50. [Google Scholar] [CrossRef]
Figure 1. Distribution by year of the publications available in the PubMed database by search keywords ‘selenium nanoparticles antimicrobial activity’ (222 articles); ‘selenium nanoparticles antibacterial activity’ (147 articles); ‘selenium nanoparticles antifungal activity’ (34 articles); ‘selenium nanoparticles antiviral activity’ (25 articles).
Figure 1. Distribution by year of the publications available in the PubMed database by search keywords ‘selenium nanoparticles antimicrobial activity’ (222 articles); ‘selenium nanoparticles antibacterial activity’ (147 articles); ‘selenium nanoparticles antifungal activity’ (34 articles); ‘selenium nanoparticles antiviral activity’ (25 articles).
Materials 16 05363 g001
Figure 2. Possible commercial products based on SeNPs (References are given in the text).
Figure 2. Possible commercial products based on SeNPs (References are given in the text).
Materials 16 05363 g002
Figure 3. General approaches to SeNPs synthesis (references are given in the text).
Figure 3. General approaches to SeNPs synthesis (references are given in the text).
Materials 16 05363 g003
Figure 4. Influence of methods for selenium nanoparticle synthesis on the size and antibacterial properties of nanoparticles: (a) The ratio of the use of various methods for selenium nanoparticle synthesis according to the literature; (b) Size distribution of selenium nanoparticles for different types of synthesis; (c) Distribution of values of the minimum inhibitory concentration of selenium nanoparticles in different types of synthesis. Each symbol means the MIC value taken from a separate publication. Colors correspond to synthesis methods: orange crosses—laser ablation, blue triangles—microwave method, cyan triangles—chemical reduction, green squares—biological reduction, orange triangles—gamma irradiation.
Figure 4. Influence of methods for selenium nanoparticle synthesis on the size and antibacterial properties of nanoparticles: (a) The ratio of the use of various methods for selenium nanoparticle synthesis according to the literature; (b) Size distribution of selenium nanoparticles for different types of synthesis; (c) Distribution of values of the minimum inhibitory concentration of selenium nanoparticles in different types of synthesis. Each symbol means the MIC value taken from a separate publication. Colors correspond to synthesis methods: orange crosses—laser ablation, blue triangles—microwave method, cyan triangles—chemical reduction, green squares—biological reduction, orange triangles—gamma irradiation.
Materials 16 05363 g004
Figure 5. Dependence of the selenium nanoparticles’ effective concentration on their size in the study of antiviral activity (all results presented in the graph use selenium nanoparticles synthesized with the chemical reduction method). Each symbol means the MIC value taken from a separate publication. Colors correspond to type of susceptible microorganisms: red circles—antiviral activity. The green straight is the trend line. Data from Table 2 were used in Figure 4 and Figure 5.
Figure 5. Dependence of the selenium nanoparticles’ effective concentration on their size in the study of antiviral activity (all results presented in the graph use selenium nanoparticles synthesized with the chemical reduction method). Each symbol means the MIC value taken from a separate publication. Colors correspond to type of susceptible microorganisms: red circles—antiviral activity. The green straight is the trend line. Data from Table 2 were used in Figure 4 and Figure 5.
Materials 16 05363 g005
Figure 6. Antibacterial action of selenium nanoparticles. (a) Distribution of minimum inhibitory concentration (MIC) values of selenium nanoparticles for the three most common studied species of bacteria: B. subtilis (orange triangles), S. aureus (purple triangles), E. coli (green circles); (bd) Dependence of the minimum inhibitory concentration of selenium nanoparticles on their size for the study of antibacterial activity: against B. subtilis (MIC values are blue circles, trend line is orange)—(b); S. aureus (MIC values are orange circles, trend line is purple)—(c); and E. coli bacteria (MIC values are red circles, trend line is green)—(d). Each symbol means the MIC value taken from a separate publication. Data from Table 2 were used in this figure.
Figure 6. Antibacterial action of selenium nanoparticles. (a) Distribution of minimum inhibitory concentration (MIC) values of selenium nanoparticles for the three most common studied species of bacteria: B. subtilis (orange triangles), S. aureus (purple triangles), E. coli (green circles); (bd) Dependence of the minimum inhibitory concentration of selenium nanoparticles on their size for the study of antibacterial activity: against B. subtilis (MIC values are blue circles, trend line is orange)—(b); S. aureus (MIC values are orange circles, trend line is purple)—(c); and E. coli bacteria (MIC values are red circles, trend line is green)—(d). Each symbol means the MIC value taken from a separate publication. Data from Table 2 were used in this figure.
Materials 16 05363 g006
Figure 7. Dependence of the minimum inhibitory concentration of selenium nanoparticles on their size in the study of antifungal activity. Preparations of nanoparticles synthesized using gamma irradiation are marked in blue, purple—using the method of laser ablation in a liquid, pink—using biosynthesis.
Figure 7. Dependence of the minimum inhibitory concentration of selenium nanoparticles on their size in the study of antifungal activity. Preparations of nanoparticles synthesized using gamma irradiation are marked in blue, purple—using the method of laser ablation in a liquid, pink—using biosynthesis.
Materials 16 05363 g007
Figure 8. Mechanisms of SeNPs’ antibacterial (a), antifungal (b) and antiviral (c) actions. Effect of SeNP functionalization on antibacterial activity of SeNPs (d). The arrows show the causal relationship of events. Red and black arrows indicate effects on microorganisms. Green arrows show effects on eukaryotic cells.
Figure 8. Mechanisms of SeNPs’ antibacterial (a), antifungal (b) and antiviral (c) actions. Effect of SeNP functionalization on antibacterial activity of SeNPs (d). The arrows show the causal relationship of events. Red and black arrows indicate effects on microorganisms. Green arrows show effects on eukaryotic cells.
Materials 16 05363 g008aMaterials 16 05363 g008b
Table 1. Examples of recent patents on SeNPs.
Table 1. Examples of recent patents on SeNPs.
Patent IDPatent NameReference
1US9624237B2Oridonin functionalized selenium nanoparticles and method of preparation thereof[73]
2US8445026B2Selenium nanoparticles with improved biological effects[90]
3US10807920B2Trichoderma-derived selenium nanoparticle foliar fertilizer for reducing crop fungal diseases and toxic contamination[75]
4US9259005B2Antipathogenic surfaces having selenium nanoclusters[57]
5CN111214460AFolic acid-chitosan-nano-selenium tumor-targeted drug delivery system and preparation method thereof[67]
6RU2798268C1Method of obtaining a veterinary drug based on non-specific immunoglobulins and colloidal particles of selenium for the correction of the immune system[52]
7JP2011501977AMethod for producing hydrous tissue paper having antibacterial and antifungal functions[56]
8KR101120635B1Method for cultivating high quality and functional vegetable fruit[80]
9US20220339187A1Protein-bound nano-selenium and preparation method and application method thereof [91]
10US10875235B2Bactericidal surface patterns[58]
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

Serov, D.A.; Khabatova, V.V.; Vodeneev, V.; Li, R.; Gudkov, S.V. A Review of the Antibacterial, Fungicidal and Antiviral Properties of Selenium Nanoparticles. Materials 2023, 16, 5363. https://doi.org/10.3390/ma16155363

AMA Style

Serov DA, Khabatova VV, Vodeneev V, Li R, Gudkov SV. A Review of the Antibacterial, Fungicidal and Antiviral Properties of Selenium Nanoparticles. Materials. 2023; 16(15):5363. https://doi.org/10.3390/ma16155363

Chicago/Turabian Style

Serov, Dmitry A., Venera V. Khabatova, Vladimir Vodeneev, Ruibin Li, and Sergey V. Gudkov. 2023. "A Review of the Antibacterial, Fungicidal and Antiviral Properties of Selenium Nanoparticles" Materials 16, no. 15: 5363. https://doi.org/10.3390/ma16155363

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