Next Article in Journal
TiO2 (Core)/Crumpled Graphene Oxide (Shell) Nanocomposites Show Enhanced Photodegradation of Carbamazepine
Next Article in Special Issue
Hydroxyapatite Particles from Simulated Body Fluids with Different pH and Their Effects on Mesenchymal Stem Cells
Previous Article in Journal
Erratum: Livashvili et al. Appearance of a Solitary Wave Particle Concentration in Nanofluids under a Light Field. Nanomaterials 2021, 11, 1291
Previous Article in Special Issue
Nanomaterials Synthesis through Microfluidic Methods: An Updated Overview
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Silver Nanoparticles Biosynthesis, Characterization, Antimicrobial Activities, Applications, Cytotoxicity and Safety Issues: An Updated Review

1
Department of Botany, Maharshi Dayanand University, Rohtak 124001, India
2
Department of Genetics and Plant Breeding, Oilseeds Section, CCS Haryana Agricultural University, Hisar 125004, India
3
Department of Botany and Plant Physiology, CCS Haryana Agricultural University, Hisar 125004, India
4
Department of Biochemistry, Punjab University, Chandigarh 160014, India
*
Author to whom correspondence should be addressed.
Nanomaterials 2021, 11(8), 2086; https://doi.org/10.3390/nano11082086
Submission received: 5 July 2021 / Revised: 6 August 2021 / Accepted: 7 August 2021 / Published: 17 August 2021

Abstract

:
Rapid advances in nanotechnology have led to its emergence as a tool for the development of green synthesized noble metal nanoparticles, especially silver nanoparticles (AgNPs), for applications in diverse fields such as human health, the environment and industry. The importance of AgNPs is because of their unique physicochemical and antimicrobial properties, with a myriad of activities that are applicable in various fields, including the pharmaceutical industry. Countries with high biodiversity require the collection and transformation of information about biological assets into processes, associations, methods and tools that must be combined with the sustainable utilization of biological diversity. Therefore, this review paper discusses the applicable studies of the biosynthesis of AgNPs and their antimicrobial activities towards microorganisms in different areas viz. medicine and agriculture. The confirmed antiviral properties of AgNPs promote their applicability for SARS-CoV-2 treatment, based on assimilating the virus’ activities with those of similar viruses via in vivo studies. In this review, an insight into the cytotoxicity and safety issues of AgNPs, along with their future prospects, is also provided.

1. Introduction

Nanotechnology represents a crucial turning point in the history of the universe. The endless desire and intelligence of humans have paved the way for groundbreaking inventions such as the internet, smartphones, rocket science, artificial intelligence and in vitro fertilization techniques, which also carry many moral issues; for this reason, nanotechnology is a field of science that was previously considered to be controversial. Overcoming these drawbacks, nanotechnology has emerged as an important tool in material science, which involves the fabrication, synthesis and manipulation of bulk molecules/particles in nanoscale dimensions. The fabricated objects, ranging in size from 1 to 100 nm, are called nanoparticles. The global applications of metal nanoparticles are due to their myriad aberrant properties. Being the smallest particle of a bulk object, nanoparticles manifest enhanced properties as more atoms are present on their surfaces, with less coordination than the bulk material [1]. Nanotechnology has unlocked new opportunities in diverse sectors such as food packaging, the environment, animal husbandry, agriculture and healthcare, making it one of the most fascinating industrial phenomena of the modern era. It is an emerging tool with which to address the challenges of making technology more sustainable and eco-friendly, with an increasing range of application [2].
Nanoparticles can be categorized into inorganic and organic classes. Inorganic nanoparticles mainly include metallic (Ag, Au), magnetic (Co, Ni) and semi-conductor (ZnO, CaSO4) types, while organic nanoparticles comprise carbon-based nanoparticles (carbon nanotubes, quantum dots) [3]. The exploitation of metal nanoparticles is beneficial due to their aberrant properties, such as their physical properties, reactivity and probable applications in diagnostics, drug delivery and antioxidant and antimicrobial studies. Aragon et al. (2015) studied the structural and hyperfine properties of Al-doped SnO2 nanoparticles manufactured via a polymer precursor approach [4]. Metal nanoparticles possess enhanced properties such as morphology, size and increased surface area in comparison to their bulk counterparts [1]. Due to the higher rate of silver commercialization, accounting for 55.4%, i.e., 313 out of 565, of all noble metal nanoparticle-based products available for consumers in the market, interest in AgNPs has grown rapidly [5]. Silver nanoparticles have been widely used for antimicrobial, anticoagulant, anticancer, orthopedic and thrombolytic purposes and in drug delivery, medical devices, sensing and diagnostics, etc. The importance of AgNPs is due to their catalytic activity, optical and thermal properties, chemical stability, thermal stability and antimicrobial activity [6]. A range of top-down and bottom-up approaches have been used to illustrate the tunable physicochemical and versatile functionality of AgNPs. Top-down approaches mainly involve evaporation–condensation processing methods, through which silver nanoparticles are constructed from large entities without atomic-level control. Meanwhile, bottom-up approaches mainly incorporate the electrochemical processing of metallic silver. This process begins with molecular components that are chemically assembled to build nanostructures using the principles of molecular recognition [7,8]. Particular attention has been paid to the inexpensive and environmentally friendly synthesis of AgNPs, which considers either the evaluation of reducing and antioxidant phytochemicals of plant origin, or a microorganism-mediated bioreduction mechanism. A range of AgNPs, including nanowires, pyramids, octahedral, tabular prisms and cubes, can be obtained by using many objects for the creation of NPs. AgNPs’ shape and size mainly depend on the response variables such as pH, temperature and concentration of Ag; meanwhile, in biological synthesis, they largely depend on the materials used for the production of AgNPs [9]. The concept of “green chemistry” has been expanding since the mid-1990s and researchers are seeking greener and more sustainable methods of AgNP synthesis. Green synthesis of AgNPs is simpler and easier, with minimal use of high-cost and potentially hazardous chemicals. Various types of bio-sources, such as bacteria, algae, fungi and plants, can be exploited, with their own numerous advantages and disadvantages for the green synthesis of AgNPs [10].
Considering the efficacy and applicability of AgNPs, this review first focuses on the green synthesis of AgNPs and the various experimental approaches used to investigate them. The antimicrobial activities of AgNPs, especially for SARS-CoV-2, along with their health and agricultural applications are also discussed. Finally, we discuss the cytotoxicity and safety issues associated with AgNPs and consider the potential of using AgNPs in future applications.

2. Green Synthesis of AgNPs

Extensive attention is being directed towards the synthesis of nanomaterials in the domain of applied physics, chemistry, catalysis and, most importantly, in diagnosis and therapeutic applications. The green synthesis of AgNPs has advantages over physical and chemical methods. The natural qualities of plant-derived secondary metabolites and single-step experimental installation stabilizes and reduces bulk silver to AgNPs. Biosynthesized approaches rely on the employment of eco-friendly chemicals or natural ingredients, which are less harmful to humans and the environment and are safe to use. During the chemical synthesis of AgNPs, the employment of expensive chemicals can carry health risks for workers and can be harmful to nature. Meanwhile, the physical approaches to nanoparticles synthesis incorporate abundant energy and force through energy-intensive processes that result in the high cost of the end-product and also environmental consequences. Nanobiotechnology, as derived from “green chemistry”, has huge possibilities for the development of novel and crucial end-products that benefit human health, the environment and industry [11]. It is estimated that one kg of raw silver costs approximately USD 14,000, while the synthesis of one kg of AgNPs would cost around USD 4 million [12]. Thus, the synthesis of silver nanoparticles via green approaches does not merely reduce the use of dangerous industrial chemicals but also the fallout from one-step manufacturing [13]. The biosynthesis of silver nanoparticles offers better control over the growth of nanoparticle crystals due to slow rate of synthesis and stabilization of the nanoparticles by dilution or steric hindrance. Depending upon the location at which the nanoparticle synthesis is carried out, the biosynthesis of nanoparticles can be divided into two main categories, namely either extracellular formation (nanoparticles form outside the cell) or intracellular formation (synthesized inside the cell). Intracellular formation of nanoparticles gives particles of appropriate shape and size, but they are not easy to separate from the reaction system in comparison to extracellular methods. Through extracellular synthesis, it is easier to understand the application of nanoparticles. Three important steps are required for the biosynthesis of AgNPs, i.e., bio-reduction of silver ions, controlled growth of the crystals to nanoparticles and stabilization of nanoparticles [14]. Microorganisms are of tremendous concern for nanoparticle synthesis; conversely, the process is susceptible to culture contagion, involves a time-consuming methodology and has less regulation over the nanoparticles’ size.

2.1. Bacterial-Mediated Synthesis of AgNPs

In the last decade, bacterial strains have been extensively used for the synthesis of inorganic nanomaterials (especially Se, Au, Ag) with appealing properties for the development of third-generation biosensors, promising diagnostic applications (cell-imaging and biolabeling, voltametric sensory devices) and for non-surface coating applications such as thin-film formation and annealing. Bacterial-mediated biosynthesized nanoparticles have also shown in vitro antimicrobial activity against pathogenic microbial strains and other properties such as antioxidant, anticoagulant, anticancer, anti-migration and antiproliferative [15]. Bacterial exposure to high metal ion concentrations reduces metal ions or forms complexes with metal ions for their survival. The metabolic pathway of some microorganisms is associated with metal ions, which, in turn, are required for their growth and are responsible for the bioconversion of metal ions into nanoparticles [14]. Various kinds of cellular transporters and oxidoreductase enzymes are involved, such as NADH-dependent nitrate reductase, and NADPH-dependent sulfite reductase flavoprotein subunit α and cysteine desulfhydrase are involved in intracellular and extracellular biocatalytic synthesis [15]. Bacterial species also have shown the ability to synthesize unique organic nanoparticles. A three-dimensional cellulose nanofibril network of aerobic acetic bacteria-like bacterial nanocellulose similar to genus Gluconacetobacter, the most competent bacteria for nanocellulose formulation, has previously been developed. Bacterial nanocellulose displays advanced purity, crystallinity and mechanical constancy in comparison to nanocrystalline cellulose and nanofibrillated cellulose [16]. Hence, bacterial nanocellulose is a nanomaterial that has attracted considerable interest for exploitation in biomedical applications (e.g., as antimicrobial agents, scaffolds for tissue engineering and for drug delivery systems) and on biosensor platforms [17]. Saeed and co-workers, in 2020, studied bacterial-mediated AgNPs and they showed noteworthy outcomes against human pathogens. The bacterial strains (Escherichia coli MF754138, Exiguobacterium aurantiacumm MF754139 and Brevundimons diminuta MF754140) were used to study secondary metabolite production. The bacterial strains used in the study exhibited numerous possibilities as antimicrobial agents against methicillin-resistant Staphylococcus aureus (MRSA) and numerous other multidrug-resistant (MDR) bacteria with a 10–28 mm zone of inhibition. The study reported an eco-friendly method for the creation of AgNPs, which will be helpful to control the nosocomial infections caused by MRSA and other human pathogens [18]. Huq reported a straightforward and eco-friendly process of AgNP formulations using Lysinbacillus xylanilyticus MAHUQ-40 to target antibiotic-resistant human microbes Vibrio parahaemolyticus and Salmonella typhimurium. Confirmation and characterization of AgNPs were performed via UV–visible spectroscopy, X-ray diffraction (XRD), Fourier-transform infrared (FTIR), field emission-transmission electron microscopy (FE-TEM) and dynamic light scattering (DLS) [19]. Ma et al. (2018) investigated AgNPs extracellularly biosynthesized with Streptomyces coelicoflavus KS-3. The AgNPs were spherical or nearly spherical in shape, followed by a small amount of shortened triangular, quadrangular and hexagonal particles extending from 2.33 to 91.3 nm in diameter. Moreover, a similar study on the cytotoxic effects of AgNPs indicated that AgNPs exhibited a good cytotoxic outcome in a dose-dependent approach against HTB-182 and A549 cells ranging in concentration from 1 to 50 µg/mL [20]. The bacterial strains used for the biosynthesis of AgNPs are summarized in Table 1.

2.2. Algal-Mediated Synthesis of AgNPs

Algae have long been exploited in food, feed, cosmetics, fertilizers, additives and pharmaceuticals. Recently, research has focused on the algal-mediated biosynthesis of nanoparticles. Algae are easy to cultivate, scalable, have rapid growth and are an excellent source of secondary metabolites. Due to these facts, interest in algal-mediated nanoparticle synthesis is increasing rapidly [29]. Algae, varying in size from microscopic (picoplankton) to macroscopic (rhodophyta), have been used for the synthesis of AgNPs. Algal strains such as Tetraselmis kochinensis, Desmodesmus and Scenedesmus have been adopted for the biological synthesis of novel metal nanoparticles. These NPs are prominently used for antimicrobial purposes and in drug delivery, electronics, catalysis and other biomedical techniques [15]. An alga hyper-accumulates heavy metals and possesses the ability to convert them into new, variable forms. Due to these appealing attributes, alga have been used to develop a range of nanomaterials. The mechanism of the algal-mediated biosynthesis of nanoparticles is characterized by the ability to control the dimensions and the phenomena of nucleation, stabilization of nanoparticles’ structure and regulation by reducing agents [30], biomolecules [31] and enzymes [32]. Merin et al. (2010) formulated AgNPs by using marine algae as a reducing and stabilizing agent [33]. A biomineralized silica cell wall called a frustule is present in the unicellular microalgae (diatoms) and forms a periodic and hierarchical 3D porous micro-nanostructure of diverse configurations. The usual functions of frustules are mechanical protection, DNA shielding from UV, biological protection and filtration and optimization of light harvesting [34,35]. Diatom frustules possess several advantages over silica materials (e.g., MCM-4), including higher biocompatibility, ease of purification and reduced toxicity. The secondary metabolites produced from algal species have been used to stabilize, cap and reduce the main metal to form metal, metal oxide or bimetallic nanoparticles. Among the diverse range of algae, red (Rhodophyceae), brown (Phaeophyceae), blue-green (Cyanophyceae), micro- and macro-green algae are the most extensively explored algae for the biosynthesis of nanoparticles to date. Till now, over 20 different green micro-algal species have been utilized for the formulation of silver nanoparticles. Algal-mediated biosynthesized AgNPs, when analyzed via different microscopic and spectroscopic techniques (SEM, XRD, FTIR, EDX, DLS), all exhibit interesting and variable physico-chemical attributes [30,36,37].
Due to the presence of various precious compounds that are responsible for the reduction and capping of nanoparticles, green macro-algae are considered bio-factories for the creation of metallic NPs. Ulva fasciata is the most common green alga used to form nanosized colloids [38]. Chaetomorpha linum is an important macro-algal green seaweed species used to formulate silver nanoparticles by prompting the reduction of Ag+ to Ag0 with the assistance of terpenoids, peptides and flavonoids. C. linum is widely acknowledged for its biological importance in the control of nutrients accessibility to its habitat [29]. Ulagesan et al. (2021) studied the biogenic preparation of AgNPs using an aqueous extract from marine red algae (Pyropia yezoensis). Spherical-shaped silver nanoparticles were confirmed, with an average crystalline size of 20–22 nm. Gram-positive and Gram-negative bacterial strains were used to study the antibacterial properties of AgNPs. The growth of Pseudomonas aeruginosa was reduced at concentration of 200 and 400 µg/mL by using Pyropia yezoensis AgNPs [39]. Algal species that have been extensively used in the past for the biosynthesis of AgNPs are listed in Table 2.

2.3. Fungal-Mediated Synthesis of AgNPs

Fungi have tremendous potential for the creation of many compounds. Around 6400 bioactive compounds are acknowledged to be formed by microscopic filamentous or non-filamentous fungal species [46]. Owing to heavy metal acceptance and the ability to internalize and bioaccumulate metals, fungal species are widely used as stabilizing and reducing agents. Furthermore, fungi can generate NPs with controlled size and morphology at a large scale in “nanofactories” [47,48]. The fungal-biosynthesized nanoparticles can be intracellular or extracellular. Although many studies have been conducted on the biosynthesis of AgNPs using fungi, the exact mechanisms have not yet been fully elucidated. It has been suggested that the extracellular formation of NPs takes place as the enzymes in the fungal filtrate act to reduce silver ions to elemental silver (Ag0) at a nanometric scale [49,50]. The enzymes most involved in the biosynthesis of metallic nanoparticles are NADH and NADH-dependent nitrate reductase [51,52]. The synthesis of endophytic AgNPs (EFNps) was performed using the aqueous extract of endophytic fungi (Lasiodiplodia theobromae) sheltered by Cinnamomum zeylanicum. Antimicrobial evaluation—for instance, minimum inhibitory concentration (MIC), minimum bacterial concentration (MBC), agar well diffusion, pyocyanin, antibiofilm and time kill curve, was performed against P. aeruginosa ATCC (27853) and an antibiotic-resistant clinical strain. Results suggested that EFNps could be a potential substitute for antibiotics to cope with the infections caused by P. aeruginosa. The fungal-mediated biosynthesis of AgNPs is a sanitary, safe, eco-friendly, reliable, inexpensive and green approach that can be used for a wide array of applications in everyday living [53]. Ammar et al. (2021) aimed for the biosynthesis, characterization and biomedical application of AgNPs via yeast metabolites. AgNPsK and AgNPsU were synthesized from the yeast strains Pichia kudriavzevii HA-NY1 and Saccharomyces uvarum HA-NY3, respectively. Highly significant inhibitory activity was shown by AgNPs hostile to Gram-positive bacteria (Staphylococcus aureus ATCC29213 and Bacillus subtilis ATCC6633) and Gram-negative bacteria (Fusarium oxysporium NRC21, Pseudomonas aeruginosa ATCC27953 and Candida tropicalis ATCC750). This study also showed the significant anticancer activity of both AgNPsK and AgNPsU against PC3 (prostate cell line), with IC50 values of 0.57, 0.50 µg mL−1 respectively, and HCT-116 (colon cell line), with IC50 values of 0.29, 0.24 µg mL−1, respectively. AgNPs were also found to be safe for the gastric profile because no ulcerogenic effects were detected in rats’ stomachs [54]. Another researcher synthesized AgNPs using Punica granatum extract as a reducing and stabilizing agent. A cubic crystalline structure with face-centered AgNPs was determined by using the XRD technique. The study revealed that AgNPs synthesized via Punica granatum extract may have a possible use in the fields of nanobiotechnology, nanomedicine and nanobiosensors [55]. The most recent studies of AgNP biosynthesis using fungal strains, along with their key attributes, are summarized in Table 3.

2.4. Plant-Mediated Synthesis of AgNPs

The plant-mediated biosynthesis of AgNPs is a straightforward practice, highly effective and with short reaction times. Due to the presence of numerous metabolites (phenols, ketones, proteins, aldehydes, amides, carboxylic acids), plants have the ability to reduce and stabilize nanoparticles. In the biosynthesis of AgNPs, almost every plant part, i.e., leaves, seeds, roots and flowers, has been used for the mining of active ingredients [10]. From close observations in recent studies on the biosynthesis of AgNPs, it can be inferred that interest is shifting significantly towards the exploitaion of medicinal flora for nanoparticle synthesis. Because of the occurrence of abundant reducing components (H+), the green leaves of medicinal plants possess the prime capacity to reduce and stabilize AgNPs [62]. Various kinds of natural compounds with proven actions against bacteria, microbes, cancerous cells and neurodegenerative disorders have been extracted from medicinal plants. Thus, the incorporation of medicinal plants into biosynthesis development could move beyond simply a green chemistry approach, enhancing the biological properties of nanoparticles [10]. In this regard, several studies have reported on the synthesis of AgNPs from medicinal plants for various applications. Jain and Mehata (2017) reported the biosynthesis of AgNPs using Ocimum sanctum leaf extract due to its exceptional antibiotic and germicidal properties. The reduction of AgNPs took place due the presence of hydroxyl and ketone groups in tulsi leaf extract and improved antibacterial activities were observed against E. coli [63]. Azadirachta indica was exploited for the biosynthesis of AgNPs and antibacterial activities against S. aureus and E. coli were observed [64]. Teucrium polium and Ocimumbasilicum were evaluated for AgNP formation, with potential antibacterial, antitumor and antioxidant activities. Biosynthesized AgNPs were also employed to study their cytotoxic activities against the HEP G2 cell line [65]. Ghramh et al. (2020) formulated spherical AgNPs of 40–45 nm using the ethanolic extract of Ruta graveolens and studied the different biological activities, immune modulation, anticancer and insecticidal potential in the extracts. The extract with AgNPs showed insecticidal activity against Culex pipiens [66]. Palithya et al. (2021) demonstrated the high negative zeta potential (−26.0 mV) value of Decaschistia crotonifolia biosynthesized Dc-AgNPs [67]. Singh et al. (2021) reported the antimicrobial activities of Carissa carandas biosynthesized AgNPs against human pathogenic bacteria [68]. Plant-mediated biosynthesized AgNPs along with their bioactivities are listed in Table 4.

3. Characterization of AgNPs

Characterization is a vital step that involves determining the size, shape, surface area, morphology, charge, dispersity and surface chemistry of nanoparticles. There is a need for concurrent development involving a wide range of techniques for a complete understanding of the system and the characterization of synthesized nanoparticles using analytical techniques such as UV–visible spectroscopy, Fourier-transform infrared spectroscopy (FTIR), X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), scanning electron microscopy (SEM), transmission electron microscopy (TEM) and dynamic light scattering (DLS). AgNP formation is verified by the appearance of a yellowish-brown color when AgNO3 is added to the culture filtrate. Confirmation can be achieved by using a culture filtrate without AgNO3 as a control. This occurs due to the excitation of plasmons on the surface of the atomic lattice of nanosized materials by light; this color change is termed surface plasmon resonance [14]. The pioneering work of Ritchie in 1950 [106] widely recognized the applications of surface plasmons in the field of surface science [107]. A strong absorption band appears in the UV–visible spectra of noble metal nanoparticles, which is usually absent in their bulk form. This is because of the collective oscillation of the conductive electrons with incident photons, and it is termed localized surface plasmon resonance (LSPR) [108]. The interaction of coherent localized oscillations with a resonant frequency largely depends on the size, geometry, dielectric environment, composition and particle–particle separation distance of NPs [109]. Because of the presence of d-d transitions exhibiting LSPR, noble metals are used for the formulation of NPs [110]. NPs exhibit a molar extinction coefficient of approximately 1011 M−1 cm−1 for absorption undergoing LSPR. However, distinct from fluorophores, plasmonic NPs do not photobleach or blink and therefore serve as intense and robust labels for immunoassays, biosensors, surface-enhanced spectroscopies and cellular imaging [111,112,113]. There is only one SPR absorption peak in the case of AgNPs, while silver nanorods (AgNRs) show two SPR absorption peaks, i.e., transverse and longitudinal [114]. Cytotoxicity and chemical degradation are the two major drawbacks of AgNPs. To overcome these drawbacks, scientists have developed a new synthetic approach for highly monodisperse polymer-coated AgNRs, which are noncytotoxic and protected against external stimuli such as heat, light and oxidation [115]. A different method of sample preparation is required for each of the characterization techniques. For ease of understanding, the fundamentals of the key approaches employed for the characterization of AgNPs are described in the following subsections.

3.1. UV–Visible Spectroscopy

In support of the primary characterization of NPs, UV–visible spectroscopy is an extremely practical and reliable technology. UV–visible spectroscopy assesses the synthesis and stability of AgNPs. The distinctive optical properties of AgNPs make them very interactive with particular wavelengths of light [116]. Because of the phenomenon of surface plasmon resonance (SPR), AgNPs display good absorption in the visible spectrum, with a maximum in the range of 400–500 nm. The resonant collective oscillations of conductive electrons along the transverse direction of the electromagnetic field cause SPR in the UV–visible region of the spectrum. Nanoparticles’ shape and temperature, the chemical surroundings and the dielectric constant of the medium influence the SPR band strength and bandwidth [117]. Observations of SPR peak intensity determine the sizes of various nanoparticles ranging from 2 to 100 nm. The stability of biosynthesized AgNPs was investigated for more than one year and an SPR peak of a similar wavelength was observed by means of UV–visible spectroscopy [118,119]. In this way, UV–visible spectroscopy is an efficient tool to characterize the AgNP structure kinetics and the final colloidal stability. Haiss et al. (2007) theoretically and experimentally determined the optical properties of AuNPs in aqueous solution using UV–vis spectra and established a correlation between the extinction efficiency (Qext) and d, which determines the particle concentrations (c) [120]. Albert et al. (2021) successfully performed characterization studies of ZnSO4-doped CeO2 nanoparticles using UV–visible spectra and calculated the band value using the Tauc plot method [121].

3.2. Fourier-Transform Infrared (FTIR) Spectroscopy

FTIR is a spectroscopic technique used to accurately and reproducibly estimate the signal-to-noise ratio. Minute absorbance varying on the order of 10−3 can be detected via FTIR spectroscopy [116]. The occurrence of functional groups in the biosynthesized extracts (plants or microorganisms), which are responsible for the reduction and stabilization of silver ions, can be efficiently determined with the help of the FTIR technique. FTIR is used to investigate the surface chemistry of metal nanoparticles [122]. During FTIR analysis, IR radiation is passed through the sample, resulting in some IR radiation being absorbed by the sample while some passes through it. The resulting spectrum shows absorbance and transmittance by creating a molecular imprint of the sample that represents the identity of the sample [123]. Consequently, FTIR is an appropriate, useful, cost-effective, non-invasive and modest technique to investigate the function of biomolecules in the reduction of AgNO3 to Ag. Bhambure et al. (2009) studied the characterization and stability of freeze-dried powdered AuNPs diluted with potassium bromide at a ratio of 1:1000 using the FTIR pattern, which revealed the strong interaction of AuNPs with the structural proteins of Aspergillus niger NCIM 616 [124]. Alamdari et al. (2020) performed FTIR analysis for the identification of functional groups (flavonoids, anthocyanins and cyaniding-3-glucoside) in Sambucus ebulus extract that contribute to the mechanism of bonding with ZnO NPs [125].

3.3. X-ray Diffraction (XRD)

For the assessment of both the molecular and crystal configuration of a compound, XRD is a very valuable analytical technique [126,127]. XRD determines particle sizes, the degree of crystallinity, isomorphous substitutions and the qualitative and quantitative resolution of various chemical compounds. During XRD analysis, an X-ray beam is projected onto the crystal and scattered by the atoms, which leads to the creation of diffraction patterns. The interference of scattered X-rays can be used according to Bragg’s law to identify the several characteristics of the crystal or polycrystalline material [123]. XRD confirms the crystalline nature of the NPs by defining the oxidation state of the particles as a function of time, and the measurements of XRD are generally calculated in Angstroms (1 Å = 0.1 nm). Even though XRD has a number of advantages, it also has some drawbacks, including obscurity in mounting the crystals, single conformation/binding state and low intensity of diffracted X-rays [128,129].

3.4. Scanning Electron Microscopy (SEM)

The field of nanotechnology and nanoscience has provided a driving force for the improvement of a range of high-resolution microscopic techniques. These microscopic techniques use a shaft of light of extremely energetic electrons to probe objects [130]. A scanning electron microscope is a surface imaging tool used for analyzing the diverse particle sizes, size distributions, nanomaterial shapes and the surface morphology of synthesized particles at the nanoscale [131]. The combination of SEM with energy-dispersive X-ray spectroscopy (EDX) can be used to determine the elemental composition of an AgNP sample [132]. The X-rays emanated by the sample are detected using the EDX technique and an EDX ray detector quantifies the comparative abundance of discharged X-rays vs. their energy [133]. The main advantage of SEM is that it can deliver valuable information regarding the purity and extent of particle aggregation, but the drawback is that it is not suitable to examine the interior structure of the sample.

3.5. Transmission Electron Microscopy (TEM)

TEM is used for the quantitative measurement of the particle size, morphology and distribution of nanomaterials. TEM projects an electron beam onto a sample and forms an image on a photographic plate [134]. The advantage of TEM over SEM is that it can provide an improved resolution and further critical dimensions [135,136]. The drawbacks of TEM include the need for a high vacuum and a large sample section. In TEM analysis, the sample preparation is laborious, but it tremendously useful to attain refined-quality images [116].

3.6. Dynamic Light Scattering (DLS)

Dynamic light scattering (DLS) is an important, fast, easy and non-destructive tool for characterizing nanoparticles that is used to grade the particle size at the micro- and nanometer regimes. During DLS, a laser passes through a colloidal solution, which scatters light at different intensities due to the Brownian motion, and the size is graded using the Stokes–Einstein relationship. The hydrodynamic diameter (the diameter of the hypothetical nonporous sphere that diffuses at the same rate as the particles being characterized) can be calculated from the time dependence of the scattering intensity measurements. The presence of an electrical double layer absorbed on the surfaces of nanoparticles and the capping agent/stabilizer usually affects the hydrodynamic diameter of the nanoparticles [105,137,138,139,140,141]. The mono-exponential equation is used for samples with purely monodisperse particles in DLS, where the fluctuation of intensity in scattered light is correlated against short decay intervals (τ) and the intensity autocorrelation function (ACF) is calculated [142].
G(τ) = 1 + b.e−2Dtq2τ
Here, b = constant dependent upon the instrument and settings of optics, Dt = translational diffusion coefficient and q = scattering vector.
Tomaszewska et al. (2013) studied the size distribution of chemically synthesized polydisperse silver nanoparticle colloids using DLS. The result revealed that several percent of the volume content of large NPs could screen completely the presence of smaller ones [139].

4. Antimicrobial Activities of AgNPs

The most significant medical achievement of the 20th century was antibiotics, as this discovery played a decisive role in treating patients with cancer or diabetes, and those experiencing microbial infections and surgical complications [143]. Regrettably, this achievement led to the emergence of antimicrobial resistance (AMR). The development of multidrug-resistant microorganisms (MDRM) has occurred due to the excessive use of antimicrobial agents against harmful pathogens [144]. Pathogenic resistance is a serious problem encountered by pharmacists and healthcare professionals. An individual septic with multidrug-resistant bacteria (MDR) cannot simply be cured since they need to be treated with a wide range of antibiotics [145]. According to the World Health Organization (WHO), drug-resistant pathogens cause a high mortality rate during disease pandemics [146]. Therefore, the advancement and alteration of antimicrobial compounds with improved uptake has been a major interest in recent years. During the last few decades, research has been concerned with the biomedical applications of metallic nanoparticles resulting from metals such as Ag, Au, Cu, Pt, etc. Among them, AgNPs have attracted major consideration due to their unique antimicrobial properties. The exact mechanism behind the antimicrobial action of AgNPs has not been illustrated yet, but through recent research studies, it has been found that, due to their small size, AgNPs can easily penetrate the microbial cell wall and generate reactive oxygen species (ROS) and free radicals, which leads to apoptosis [147]. The development of nanotherapeutics has received a great deal of attention, particularly regarding AgNPs because of their broad-spectrum oligodynamic properties. Dating back to 1881, the first evidence of using silver in medicine was reported to treat eye infections in neonates and, soon after, in 1901, for internal antisepsis. Currently, drugs such as silver nitrate and silver sulfadiazine are frequently used to treat wounds and dermal burns and to eradicate warts [148]. Feng et al. (2000) used electron microscopy and X-ray microanalysis to perform a mechanistic study of the antibacterial action of Ag+ on E. coli and Staphylococcus aureus. The results of electron microscopy showed an electron-light region in the center of the cells and many small, electron-dense granules either surrounding the cell wall or deposited inside the cells. Meanwhile, X-ray microanalysis revealed the existence of Ag elements and sulfur in electron-dense granules and cytoplasm. The study also suggested that DNA lost its replication potential and the protein became activated after Ag+ treatment [149]. During World War I, silver was the most used material in to treat infections among soldiers [150]. The antimicrobial applications of AuNPs are related to their non-toxicity, ease of detection, polyvalent effects, high capacity for functionalization and photothermal properties [151]. Because of their biocompatibility, functionalization and covalent bonding to low-molecular-weight chitosan, AuNPs can also be used as a vehicle for drug and vaccine delivery [152]. Usman et al. (2013) tested the antimicrobial activities of Cu–chitosan NPs (2–350 nm) against S. aureus, B. subtilis, C. albicans, P. aeruginosa and Salmonella choleraesuis [153]. Cho et al. (2004) studied the antimicrobial activities of Ag and Pt NPs against S. aureus and E. coli. using sodium dodecyl sulfate (SDS) and poly-N-vinyl-2-pyrrolidone (PVP) as stabilizing agents [154]. This section of the review is focused on the main achievements and the mechanisms of AgNPs as nanocarriers for the inhibition of a variety of microbial infections.

4.1. Antibacterial Action of AgNPs

Bacterial resistance to antibiotics is a major health issue with a massive impact globally. In the last few decades, pharma-industries have focused their concern on developing novel antibiotics with better capability to target bacterial diseases [155]. Nanoparticles have become very useful against bacterial infections considering their large surface volume area and high synergy arising from their multivalent interactions. AgNPs are the most extensively used antibacterial nanoagent because of their wide-ranging antimicrobial effectiveness against several bacteria [156]. AgNPs intermingle via the bacterial cell envelope, yet the prime cellular target remains unknown. The combined effects of AgNPs along with antibiotics lead to increased antibacterial activity against drug-resistant bacteria [157]. The combined uptake of AgNPs and popular antibiotics such as kanamycin, tetracycline, enoxacin and neomycin led to the suppression of multiple-drug-resistant S. typhimurium bacterial growth; however, this synergistic effect is not observed in the case of penicillin and ampicillin [158]. The antibacterial activity of AgNPs along with other common antibiotics such as erythromycin, vancomycin, amoxicillin, ciprofloxacin, streptomycin, tetracycline and gentamicin is improved against E. coli and S. aureus [159], while kanamycin, ampicillin and chloramphenicol showed synergistic effects against various bacterial strains including St. mutans, St. aureus, Ent. faecium and E. coli [160]. Nanomaterials possess superior antibacterial potency against Gram-positive bacteria in comparison to Gram-negative bacteria. AgNPs can constantly liberate Ag0, which may be considered a means of killing microbes. The approaches that illustrate the mechanism of antibacterial action of AgNPs are summarized in Figure 1. The adherence of silver ions to the cytoplasmic membrane and cell wall occurs because of the electrostatic affinity of Ag0 towards sulfur proteins. This leads to the disruption of the bacterial envelope by enhancing the permeability of the cytoplasmic membrane [161]. The uptake of Ag0 into the cells results in the deactivation of respiratory enzymes, the formation of ROS and interruption of ATP production [162]. ROS can play a prime role in the processes of DNA modification and cell membrane disruption. The interaction of Ag0 with the sulfur and phosphorus components of DNA results in DNA modification. Likewise, Ag0 can hinder the formulation of proteins by denaturing ribosomes in the cytoplasm [163].
Silver nanoparticles can also kill bacteria themselves, without the release of silver ions. After the attachment of AgNPs to the cell surface, they accumulate in the cell wall pits, which results in cell membrane disruption [164]. Likewise, AgNPs also disrupt the bacterial signal transduction by dephosphorylate tyrosine residues on the peptide substrates, which causes the termination of cell multiplication and apoptosis [165]. The antibacterial action of AgNPs also depends on the dissolution status of AgNPs in exposure media. The synthetic and processing factors, such as intrinsic AgNP characteristics (shape, size, capping agent) and surrounding media (organic and inorganic components), directly affect the dissolution efficacy of AgNPs [166,167,168,169]. The thick cellular wall of Gram-positive bacteria may reduce the penetration of AgNPs into their cells; this is why Gram-negative bacteria are more susceptible to AgNPs [170]. The biofilm formation in the oral environment protects bacteria from both Ag0 and AgNPs by hindering their transport. The bioavailability and mobility of AgNPs in the biofilm is determined by the AgNPs’ diffusion coefficients [171,172,173]. Ivan Sondi and B.S. Sondi (2004) also studied the biocidal activities of AgNPs against E. coli and confirmed the “pit” formation in the cell wall of this model Gram-negative bacterium. A significant increase in permeability occurred due to the accumulation of AgNPs in the bacterial membrane, resulting in cell death [174].

4.2. Antifungal Action of AgNPs

Since the primeval era, fungal infections have contributed considerably to the escalating morbidity and mortality. Studies have found that outbreaks caused by pathogenic fungi can be controlled by exploiting the fungicidal or fungistatic activity of nanoparticles [175]. Biosynthesized AgNPs stabilized with sodium dodecyl sulfate (SDS) exhibit good antifungal activity compared to fluconazole and are hostile towards phytopathogens such as Aspergilus niger, Trichophyton mentagrophytes, Fusarium semitectum, Candida glabrata, Issatchenkia orientalis, Phoma glomerata, Candida albicans and Phoma herbarum [176].
Panacek et al. (2009) demonstrated the potent antifungal activity of AgNPs stabilized with surfactants of size 25 nm, which are hostile towards four Candida strains, with an MIC value extending from 0.21 to 1.69 mg/L [177]. The antifungal activity of corn extract biosynthesized AgNPs hostile to the phytopathogenic fungus Phomopsis vexans was demonstrated by reducing the expansion of mycelium by 30–40% in Potato Dextrose Agar (PDA) medium [178]. Elgorban et al. (2016) analyzed the antifungal action of AgNPs against plant pathogenic fungi Rhizoctonia solani infecting cotton plants [179]. Similarly, the antifungal activity of Ag ions and AgNPs hostile to two plant pathogenic fungi (Bipolaris sorokiniana and Magnaporthe grisea) was reported [180].
The following mechanistic pathways are considered responsible for the antifungal activity of AgNPs: (i) fungal cells easily uptake AgNPs due to their small size, which leads to the disturbance of fungal cell walls; (ii) AgNPs perform as a source of Ag+ ions that stops DNA replication and ATP synthesis, by hydroxyl radical and ROS formulation. Because of this, the biochemical cycle of fungal cells is stopped, inducing fungal cell death. Owing the strong tendency of Ag+ for the thiol groups of the cysteine protein in fungal cells, Ag+ ions exhibit anticandidal activity through the inactivation of ATP synthesis and enzyme function, leading to cell death [181].

4.3. Antiviral Action of AgNPs

Currently, viruses are recognized as one of the most significant causative agents of human disease. Regardless of their evident structural simplicity, viruses pose a massive threat in the face of perilous diseases, such as the Spanish flu, HIV, Ebola, Marburg virus, and, lastly, the 2020 pandemic caused by COVID-19 [9]. Viral infections pose major challenges to global health, particularly given the fact that the appearance of resistant viral strains and the deleterious side effects associated with long-term use persist to slow the adoption of effectual antiviral therapies. This makes it crucial to develop safe and efficient alternatives to prevailing antiviral drugs. The recent insurgence of COVID-19, which has developed resistance towards existing antiviral drugs, has led researchers to search for new antiviral agents. The SARS-CoV-2 pandemic began in December 2019 and has killed more than 3.2 million people worldwide as of May 2021 [182]. The pathogenic nature of viruses depends on their attachment and infiltration into the host cells via the binding of viral surface constituents with ligands and proteins on the cell membrane. Thus, the preeminent approach in developing new antiviral drugs is to prevent such bindings. In the current situation, metal nanoparticles have emerged as novel antiviral agents because of their exceptional physical and chemical properties; AgNPs have become important candidates as antiviral agents. AgNPs wide range of mechanisms of attack towards their targets can reduce the microbial resistance to these nanoparticles [183]. Table 5 summarizes the most significant studies on AgNPs’ antiviral activity towards different viruses. However, details of the exact mechanisms of their action or interactions are limited so far.
The complexity of the virus structure may contribute to the limited knowledge of the mechanism of interference of nanoparticles with viruses. There are two possible methods by which AgNPs exert their antiviral activities, which are as follows: (i) AgNPs bind to the outer coating of proteins, thus suppressing the attachment of the virus to cell receptors; (ii) AgNPs bind to nucleic acid (DNA/RNA) and inhibit the replication or proliferation of the virus inside the host cells [183]. It is evident that AgNPs can alter the structure of surface proteins, thus reducing their detection and adhesion to the host receptor. Figure 2 shows the mechanism of AgNPs’ antiviral efficacy. AgNPs prevent the commencement of transmitted gastroenteritis virus (TGEV) by binding to a surface protein, S-glycoprotein [184]. Furthermore, Sharma et al. (2019) focused on the biological creation of AgNPs from medicinal plants (Tinospora cordifolia, Andrographis paniculata and Phyllanthus niruri) and evaluated their antiviral properties against chikungunya virus [208].
A crucial characteristic of nanoparticles in relation to Coronavirus (CoV) is that they are able to compete with viral binding to a receptor on the surface of a host cell. Moreover, the penetration of CoV into host cells is mediated by ACE2 receptors, particularly in the case of SARS-CoV and SARS-CoV-2 [209]. Therefore, blocking and/or lowering ACE2 levels could help to fight CoV infections as well as generate antibodies against ACE2. However, on the other hand, increased production of vasodilating angiotensin 1–7 was found to occur due to the protective action of ACE2 against viral lung damage after infection [210]. Hence, obviating COVID-19 in the host may be more helpful than fighting the virus after infection. Research studies on HIV have shown the attachment of AgNPs to disulfide bonds of the CD-4 binding domain of the glycoprotein 120 (gp120) surface proteins [201]. Similarly, AgNPs can disrupt the disulfide bonding on ACE2 and spike protein and exert their antiviral efficacy against SARS-CoV-2. The interaction of AgNPs with the ssRNA of Coronavirus improves the antiviral activity of AgNPs. The most effective antiviral efficacy of AgNPs occurs in those with a diameter of 10 nm. This was confirmed by an immunofluorescence study involving 10 nm AgNPs capped with polyvinylpyrrolidone, which completely inhibited SARS-CoV-2’s activities in comparison to 100 nm AgNPs, which did not [211]. AgNPs also showed their efficacy against key determinant proteins of pathogenicity, namely hemagglutinin (H) and neuraminidase (N). The combination of AgNPs with antiviral drugs such as zanamivir, osltamivir, FluPed and amantadine results in the production of ROS, which reduces the destruction of MDFK cells and reduces H and N proteins’ action in healthy cells [197,212,213]. Thus, it can be concluded that AgNPs’ potent effects for SARS-Co-2 could be of potential therapeutic benefit, based on the comparative in vivo reports on similar viruses.
NPs provide stabilization and the discharge of active ingredients of vaccines in the body; thus, NPs can also operate as carriers. Among certain types of respiratory diseases, NPs have been used in vaccines to treat respiratory syncytial virus (e.g., polyanhydrides), H1N1 influenza and human parainfluenza virus type 3 [214]. Further studies of the antiviral activity of AgNPs may reveal further potential in the treatment of diseases caused by a variety of viruses, especially COVID-19.

5. Applications of AgNPs

In the last few decades, it has been shown that biosynthesized AgNPs have potential uses in agriculture, health, food and industry because of their unique characteristics. A joint venture of the Food and Agriculture Organization (FAO) and WHO in 2009 presented the applications of nanotechnology in agriculture and food, with the addition of across-the-board fields such as food packaging, nanostructured ingredients, nanocoatings, nanosized biofortification and nanofiltration [215].

5.1. AgNPs in Agriculture

In adopting the 2030 plan for sustainable development, the United Nations dedicated itself to abolishing poverty and hunger and making agriculture sustainable. The global population is expected to grow to 9.8 billion by 2050; feeding such a huge population will necessitate at least a 50% rise in the production of agricultural products from 2012 levels by the mid-century [216]. Diverse nano-enabled measures are planned to progress crop production and facilitate the necessary increase in production for food, feed and fuel while practicing sustainable agriculture. NPs are considered “magic-bullets” to increase the production of agriculture as they contain nutrients, valuable genes and organic compounds that can be tailored towards precise plant structures or areas. Consequently, NPs represent elegant nanodelivery systems for agriculture administration, especially for crop nutrition. Following the poor outcomes of the “First Green Revolution” during the 1970s, there is an urgent need for a “Second Green Revolution” with an eco-friendly and more sustainable strategy. In agriculture, huge economic losses are caused by assorted plant diseases. After the first green revolution, the concept of biopesticides emerged to combat the arbitrary use and ill effects of chemical pesticides [217]. AgNPs are principally used for plant disease management because of their unique antimicrobial properties [218]. Many researchers have proposed the exploitation of the antimicrobial activities of AgNPs that are hostile towards a variety of plant pathogens.
Research has been focused on the direct applications of AgNPs in agriculture, such as root elongation, seed germination, nanofertilizers, nanopesticides and plant modification (cytotoxicity or cellular oxidative stress) in the presence of metal NPs [117,219] and indirect usage based on the antimicrobial activities of NPs [220]. To boost crop yields, the use of chemical fertilizers is key, but their effectiveness is restricted by the occurrence of leaching or volatilization, which contaminates the environment and raises the cost of production. Therefore, nanofertilizers are gaining significant attention in sustainable agriculture as a way to extend the use of slow-release fertilizers, minimize the loss of mobile nutrients and facilitate access to poorly available nutrients [221]. In agriculture, in order to collect real-time statistics on crop growth, nutrients and water availability, nanosensing devices have been developed. Moreover, to reduce the use of chemicals on seeds and crops, carbon nanotubes (CNTs) and mesoporous silica nanoparticles have been applied [8].
The formulation of traditional pesticides with polymers or metal nanoparticles is an emerging area within the pesticide industry. The advantage of nanoencapsulation is the slow and reduced release of active ingredients by the use of nanocapsules, which minimizes surplus overflow of redundant pesticides and instead uses low doses over a prolonged time period [222]. An additional benefit of nanocarriers in plant protection is their site-targeted delivery and the stability of active ingredients [223]. In agriculture, AgNPs can be used as an ecologically sound strategy to replace pesticides and synthetic additives, as small concentrations of AgNPs are effective against pathogens, without toxicity to humans. Scientists have reported the broad-spectrum phytopathogenicity of AgNPs against various phytopathogens such as Scalerotinia sclerotiorum, Fusarium culmorum, Trichoderma sp., Botrytis cinerea, Colletotrichum gloeosporioides, Rhizoctonia solani, Biploaris sorokinniana, Sphaerotheca pannasa, Phythium ultimum, Phoma and Megnaporthe grisea [224,225,226].
A study has reported the application of phytosynthesized (kaffir lime leaf extract) AgNPs for boosting the germination and starch metabolism of matured rice seeds via nanopriming technology. The formulation of elevated soluble sugar content for sustaining the seedling development was made possible as the nanopriming stimulated α-amylase activity [227]. In another study, researchers explored the role of AgNPs in the development of seed germination and its effect on the plumule and radical length of Pennisetum glaucum [228]. The study revealed the importance of biosynthesized AgNPs, which can play a significant role in controlling Acidovorax oryzae strain RS-2. AgNPs downregulate the expression of type VI secretion system allied genes, which affects the virulence of bacteria [229]. Global efforts are underway to reduce the use of dangerous substances, especially chemical pesticides, in crop production. The biological means of soil phytopathogens has been confirmed to be a suitable substitute for the use of chemical pesticides. Past experience suggests that the use of fungal strains is a reliable and sustainable approach for the synthesis of AgNPs. In a study, Trichoderma viride biosynthesized AgNPs acted as an appropriate and successful biocontrol agent against rice pathogenic fungi such as Rhizoctonia solani and Fusarium moniliforme and offered better crop and disease control [230].
Currently, improving the yield is the primary goal for plant breeders in order to meet the growing harvest demand. To accelerate crop development, advanced biotechnological techniques such as next-generation sequencing (NGS) and genome-assisted breeding (GAB) are being used for the detection and assessment of genetic variability. Alternation of gene expression via DNA and RNA isolation in plants can be used for the development of biotic and abiotic stress-resilient crops [231]. However, poor cell membrane penetration and nucleus targeting are major challenges for the progression of plant-targeted delivery systems. The meeting of biotechnology with nanotechnology could help to overcome these challenges. Nowadays, NPs are being used as an efficient gene transformation vehicle. SiO2 NPs have been used to transport DNA fragments into corn and tobacco plants, without any harmful side effects [232]. The higher DNA delivery efficiency into plants (Nicotiana tabacum, Oryza sativa and Leucaena leucocephala) by AuNPs embedded in sharp carbonaceous carriers was established [233]. Thus, through desirable modification of plants, the future demand for plant-derived products can be met using nanobiotechnology applications. Nanocoating-based materials can play a significant role in agro-machinery applications by improving their resistance to aerial oxidation-based deterioration and UV rays. A future application could be in smart machines for specific weed management through the inclusion of nanobiosensors [231].

5.2. Food and AgNPs

Nanopotentials have been found to improve food characteristics, safety and production efficiency in food industries throughout the world. The antimicrobial efficacy of AgNPs has been widely explored in the food industries. Green synthesized AgNPs are considered safer or less toxic for potential applications in food items [234]. The use of several Ag-containing food packaging materials has been approved by the USFDA in order to preserve the shelf life of food items and remain in direct contact with food items [235]. The safety and freshness of fruits and vegetables is one of the major problems faced in the agri-food sector. The inability of natural wax coatings to prevent water loss and reduce the respiration rate causes protein and weight losses during the long-term storage of fruits and vegetables. The shelf life of different food items can be significantly improved by the usage of various protective nanocoatings and appropriate packaging materials [1]. Aseptic food packaging materials can be developed by the use AgNPs, as AgNPs exhibit a broad array of antibacterial actions. Researchers increased the shelf life of fruits and vegetables by incorporating mycosynthesized (Trichoderma viride) AgNPs into sodium alginate, which prevented the protein and weight loss and also avoided microbial spoilage [236]. Similarly, another study reported the improved shelf life of chikku and grapes with the thin film of A. niger biosynthesized AgNPs in combination with sodium alginate [237].

5.3. Biomedical Applications

The unique physicochemical and antimicrobial properties of AgNPs have found a wide range of biomedical applications, such as in pharmacology, drug delivery, diagnostics, anticancer, etc. Garcia-Contreras et al. (2011) reported the exploitation of AgNPs in dental practices, such as in endodontic retrofill cement, dental implants and restoring material [238]. AgNPs’ unique toxicity profile can enable the targeting of specific vulnerabilities in cancer due to their negligible systemic toxicity. Several studies reported the beneficial role of AgNPs in cancer therapy. The toxic effects of biosynthesized AgNPs against carcinoma cells compared to non-cancer cells were studied [239]. Another study showed the significant inhibition of cancer cell lines (B16F10, A549, MCF7 and HNGC2F) after treatment with AgNPs synthesized with Butea monosperma leaf extract [240]. The toxicity profiling of AgNPs on ovarian cancer stem cells was investigated and outcomes showed that A2780 and ALDH+/CD133+ colonies were reduced significantly [241].
Inflammation is the reaction of the immune system to infections and injury to cells, which helps in tissue organization and the elimination of harmful factors [242]. An inflammatory disorder can develop due to the disruption of this complex process [243]. Anti-inflammatory compounds such as cytokinins and interleukins are produced by the primary immune organs in response to inflammatory disorders [244,245]. Biologically synthesized AgNPs possess these anti-inflammatory activities because the alkaloid or flavanoid contents function as capping agents and provide supplementary pharmacological properties. Nanosilver in higher doses has been used to facilitate the strong production of Th1 cells by the secretion of IL-2 and INF-γ, which play a crucial role in cellular immunity [243]. A research study was carried to evaluate the anti-inflammatory action of AgNPs synthesized by Leucas aspera plant. Ethanol and AgNPs of L. aspera demonstrated anti-inflammatory action against carrageenan-induced paw-edema in rats and indomethacin was used as a standard drug. Excellent anti-inflammatory activities were reported using L. aspera biosynthesized AgNPs [246]. Kim et al. (2007) used Muller Hinton agar plates to study the antimicrobial activities of AgNPs against E. coli, Staphylococcus aureus and yeast, showing that they are suitable for various medical devices and antimicrobial control systems [247].

5.4. Environment and AgNPs

The biosynthesized AgNPs have emerged as novel tools for the development of nanobiosensors to target environmental pollutants. The detection of residual pesticides in environmental samples has become possible via the surface modification of nanoparticles with an appropriate agent. For the detection of pesticides, nanotechnology-based colorimetric methods offer several advantages, such as accuracy, no need for specialized equipment, simplicity, less time consumption, etc. The specificity and sensitivity of nanoparticle-based colorimetric detection methodology is based on the chemical and molecular connections among the surface-modified nanoparticle and target pesticides [248]. The high extinction coefficient and strong surface plasmon resonance of silver (Ag) have attracted interest for the recognition of pesticide residues in the environment [249,250]. Xiong and Li described a colorimetric probe for the quantitative detection of residual pesticides by means of calixarene-modified AgNPs. The preparation of extremely firm calixarene-modified silver nanoparticles (pSC4-Ag NPs) was achieved by a one-pot synthesis scheme and they were then characterized. pSC4-Ag NPs have been proposed as a new sensor for the colorimetric recognition of residual pesticides in water, with an optunal down to a concentration of 10−7 M [251]. Similarly, Menon et al. (2013) reported a straightforward and extremely sensitive method for the detection of dimethoate pesticides in industrial wastewater via p-sulphonate calyx resorcinarene customized silver nanoparticles (pSC4R-Ag NPs) [252].

6. Cytotoxicity and Safety Issues of AgNPs

The exploitation of AgNPs is in the infancy stage and the overall evaluation of their health hazards is so far not promising, although a number of products are already on the market. In the immediate future, we expect that the use of AgNPs in human health will increase significantly because of their utility, principally in relation to pandemic control. The lethal outcomes of AgNPs in organisms depends on several factors, such as the route of exposure (penetration, concentration and duration), factors associated with vulnerable organisms and factors associated with AgNPs’ intrinsic toxicity, bioavailability and accumulation in organisms [253]. Three possible methods of consuming AgNPs are: inhalation, parenteral/dermal and oral. The cytotoxic effects of AgNPs depend on their size, tissue allocation, infiltration competence and cellular absorption [254].
Once in the human body, large AgNPs can be exhaled, while the smaller AgNPs can be deposited in the lungs and can reach different organs via the bloodstream. According to histopathological data studies, no considerable changes were found in the lungs, liver, nasal cavities or other organs regarding AgNPs with average sizes of 15–30 nm and in concentrations of 0.5–381 µg/m3 [255,256,257,258]. However, brain injuries have been reported at high concentrations of AgNPs above 2.9 mg/m3 [259].
The adhesiveness of AgNPs in biological tissues also depends on the presence of an electric charge on the NPs’ surfaces [260]. Negatively charged NPs show quite low DNA encapsulation in comparison to positively charged NPs, which attach to DNA plasmids via electrostatic contacts and increase their stability. AgNPs also form electrostatic contacts with certain blood proteins, ions and other components [261].
The level of AgNPs’ cytotoxicity in dermal injections rises with a concentration of 0.1–1000 mg/kg and size of 20–100 nm, causing cerebral [262], lung [263], renal [264] and liver lesions [265].
Oral ingestion involves an intermediate level of toxicity, where the NP dosage ranges (0.5–500 mg/L) were decisive rather than their size. Various research studies showed the following effects of NPs doses: weight loss at doses of 10 mg/kg [266], liver disorders at doses >300 mg/kg [267] and oxidative stress generated at >1000 mg/kg doses of NPs [268]. In oral ingestion, most of the applied doses of NPs are ultimately expelled in the feces; hence, accumulation is least comparable to dermal injection [269]. Vazques-Munoz et al. (2015) investigated AgNPs’ toxicity in a wide range of biological entities ranging from virus to human cell lines. This study revealed that different biological entities were inhibited within the same order of magnitude (10 µg/mL) of AgNPs. This was because of the interactions of AgNPs with fundamental components of cells and viruses alike [270].
The transfer of in vivo toxicological knowledge regarding dosages of AgNPs from rats to humans is of immense significance, because not many studies have been performed concerning human participants. Using AgNP exposure concentrations relative to deposited mass per alveolar surface area, correlation is made via multipath particle dosimetry modeling. The alveolar surface area of rats is 0.409 m2, while in the case of humans, it is 62.7 m2 [271]. Certain studies suggest that experience time is not the most important aspect to determine AgNPs’ cytotoxicity for humans, because, in certain cases, no blood or hematological changes were observed even when exposure exceeded 5 years with concentrations of 0.35 and 1.35 mg/m3 [272]. A report suggested that a jewelry manufacturer works at a much lower concentration of AgNPs (0.2–2.8 mg/L) in comparison to an individual who recovers Ag in close proximity to soluble compounds, with a concentration of 1.3–20 mg/L. Therefore, an extended experience times does not influence the health of workers; rather, exposure depends on the activity. At moderate exposure, blood AgNP levels should be in the range of 0.1–23 mg/L, whereas sporadic exposure should be around ≤0.1 mg/L [272,273]. Nevertheless, research studies on the cytotoxicity of NPs need to be expanded, particularly focusing on their effects in the lungs, since NP exposure occurs primarily through inhalation.

7. Conclusions and Future Prospects

AgNPs have long been of interest to researchers due to their customizable properties. The enhanced properties of green synthesized AgNPs can be exploited for use in almost all areas of science and everyday life, including preventing epidemics and treating contagious diseases. This article comprehensively reviewed the updated progress in AgNP biosynthesis approaches, their antimicrobial activities and their applications in various fields. The underlying concepts behind biosynthesized AgNPs, their antimicrobial activities and their cytotoxicity have been studied carefully with regard to the mechanisms, types and factors controlling the nanosynthesis process and their actions. Better-quality nanomaterials are formulated via intracellular biosynthesis approaches. Nonetheless, the downstream processing of intracellular nanoparticles is highly expensive and difficult in comparison to extracellular methods. Further studies are required to elucidate the detailed mechanisms of AgNP production to achieve pure nanoparticles, as well as to determine their stability and the functions of the metabolites produced by the microorganisms. This review also highlights the applications of AgNPs in terms of SARS-CoV-2 and suggests that future studies could play a prominent role in driving the myriad microorganisms and plants explored till date to achieve their on-field applications. The inhibitory effect of AgNPs on SARS-CoV-2 could lead to a novel scientific approach for preventing infections at an early stage. There is also emerging evidence of the ability of AgNP coatings to reduce their cytotoxicity and increase their specificity. Therefore, more research is needed on the effective doses and probable toxic effects of AgNPs to create a safe environment for humans against extremely dangerous diseases such as SARS-CoV-2. Lastly, a better understanding of the cytotoxicity and safety issues associated with AgNPs towards humans and the environment could further the limits of this technology and expand their horizon beyond laboratory use.

Author Contributions

Conceptualization, D.B., A.S., G.C., M.K. and A.R.S.; draft preparation and writing—original review, D.B., A.S., G.C., M.K., M.S., N.R., P.M. and A.R.S.; supervision and editing, A.R.S. All authors have read and agreed to the published version of the manuscript.

Funding

Did not receive any specific grant from funding agencies.

Data Availability Statement

This paper does not report any new data.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shyam, A.; Chandran, S.S.; Bini, G.; Sreelekha, E. Plant mediated synthesis of AgNPs and its applications: An overview. Inorg. Nano Met. Chem. 2020, 1–17. [Google Scholar] [CrossRef]
  2. Mishra, M.; Dashora, K.; Srivastava, A.; Fasake, V.D.; Nag, R.H. Prospects, challenges and need for regulation of nanotechnology with special reference to India. Ecotoxicol. Environ. Saf. 2019, 171, 677–682. [Google Scholar] [CrossRef] [PubMed]
  3. Tehri, N.; Vashishth, A.; Gahlaut, A.; Hooda, V. Biosynthesis, antimicrobial spectra and applications of silver nanoparticles: Current progress and future prospects. Inorg. Nano Met. Chem. 2020, 1–19. [Google Scholar] [CrossRef]
  4. Aragón, F.H.; Coaquira, J.A.H.; Villegas-Lelovsky, L.; Da Silva, S.W.; Cesar, D.F.; Nagamine, L.C.C.M.; Cohen, R.; Proupin, E.M.; Morais, P.C. Evolution of the doping regimes in the Al-doped SnO2nanoparticles prepared by a polymer precursor method. J. Phys. Condens. Matter 2015, 27, 095301. [Google Scholar] [CrossRef] [PubMed]
  5. Agnihotri, S.; Mukherji, S.; Mukherji, S. Size-controlled silver nanoparticles synthesized over the range 5–100 nm using the same protocol and their antibacterial efficacy. RSC Adv. 2014, 4, 3974–3983. [Google Scholar] [CrossRef] [Green Version]
  6. Tehri, N.; Kaur, R.; Maity, M.; Chauhan, A.; Hooda, V.; Vashishth, A.; Kumar, G. Biosynthesis, characterization, bactericidal and sporicidal activity of silver nanoparticles using the leaves extract of Litchi chinensis. Prep. Biochem. Biotechnol. 2020, 50, 865–873. [Google Scholar] [CrossRef]
  7. Gherasim, O.; Puiu, R.A.; Bîrcă, A.C.; Burdușel, A.-C.; Grumezescu, A.M. An Updated Review on Silver Nanoparticles in Biomedicine. Nanomaterials 2020, 10, 2318. [Google Scholar] [CrossRef] [PubMed]
  8. Kumar, A.; Gupta, K.; Dixit, S.; Mishra, K.; Srivastava, S. A review on positive and negative impacts of nanotechnology in agriculture. Int. J. Environ. Sci. Technol. 2018, 16, 2175–2184. [Google Scholar] [CrossRef]
  9. Mikhailova, E.O. Silver Nanoparticles: Mechanism of Action and Probable Bio-Application. J. Funct. Biomater. 2020, 11, 84. [Google Scholar] [CrossRef] [PubMed]
  10. Moradi, F.; Sedaghat, S.; Moradi, O.; Salmanabadi, S.A. Review on green nano-biosynthesis of silver nanoparticles and their biological activities: With an emphasis on medicinal plants. Inorg. Nano Met. Chem. 2020, 51, 133–142. [Google Scholar] [CrossRef]
  11. Menges, N. The Role of Green Solvents and Catalysts at the Future of Drug Design and of Synthesis. In Green Chemistry; IntechOpen: London, UK, 2018. [Google Scholar] [CrossRef] [Green Version]
  12. Castillo-Henríquez, L.; Alfaro-Aguilar, K.; Ugalde-álvarez, J.; Vega-Fernández, L.; de Oca-Vásquez, G.M.; Vega-Baudrit, J.R. Green synthesis of gold and silver nanoparticles from plant extracts and their possible applications as antimicrobial agents in the agricultural area. Nanomaterials 2020, 10, 1763. [Google Scholar] [CrossRef]
  13. Javed, B.; Ikram, M.; Farooq, F.; Sultana, T.; Mashwani, Z.-U.; Raja, N.I. Biogenesis of silver nanoparticles to treat cancer, diabetes, and microbial infections: A mechanistic overview. Appl. Microbiol. Biotechnol. 2021, 105, 2261–2275. [Google Scholar] [CrossRef]
  14. Poulose, S.; Panda, T.; Nair, P.P.; Théodore, T. Biosynthesis of Silver Nanoparticles. J. Nanosci. Nanotechnol. 2014, 14, 2038–2049. [Google Scholar] [CrossRef] [PubMed]
  15. Grasso, G.; Zane, D.; Dragone, R. Microbial Nanotechnology: Challenges and Prospects for Green Biocatalytic Synthesis of Nanoscale Materials for Sensoristic and Biomedical Applications. Nanomaterials 2019, 10, 11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Golmohammadi, H.; Morales-Narváez, E.; Naghdi, T.; Merkoçi, A. Nanocellulose in Sensing and Biosensing. Chem. Mater. 2017, 29, 5426–5446. [Google Scholar] [CrossRef]
  17. Pourreza, N.; Golmohammadi, H.; Naghdi, T.; Yousefi, H. Green in-situ synthesized silver nanoparticles embedded in bacterial cellulose nanopaper as a bionanocomposite plasmonic sensor. Biosens. Bioelectron. 2015, 74, 353–359. [Google Scholar] [CrossRef] [PubMed]
  18. Saeed, S.; Iqbal, A.; Ashraf, M.A. Bacterial-mediated synthesis of silver nanoparticles and their significant effect against pathogens. Environ. Sci. Pollut. Res. 2020, 27, 37347–37356. [Google Scholar] [CrossRef]
  19. Huq, M.A. Biogenic silver nanoparticles synthesized by Lysinibacillus xylanilyticus mahuq-40 to control antibiotic-resistant human pathogens vibrio parahaemolyticus and Salmonella typhimurium. Front. Bioeng. Biotechnol. 2020, 8, 1407. [Google Scholar] [CrossRef] [PubMed]
  20. Ma, L.; Liu, J.; Su, W.; Zeng, X.; Liu, X.; Li, W.; Deng, J.; Tang, J. Extracellular Biosynthesis, Characterization and Cytotoxic Effect of Silver Nanoparticles by Streptomyces coelicoflavus KS-3. J. Nanosci. Nanotechnol. 2018, 18, 8133–8141. [Google Scholar] [CrossRef] [PubMed]
  21. Sahoo, C.R.; Maharana, S.; Mandhata, C.P.; Bishoyi, A.K.; Paidesetty, S.K.; Padhy, R.N. Biogenic silver nanoparticle synthesis with cyanobacterium Chroococcus minutus isolated from Baliharachandi sea-mouth, Odisha, and In Vitro antibacterial activity. Saudi J. Biol. Sci. 2020, 27, 1580–1586. [Google Scholar] [CrossRef]
  22. Akter, S.; Huq, A. Biologically rapid synthesis of silver nanoparticles by Sphingobium sp. MAH-11T and their antibacterial activity and mechanisms investigation against drug-resistant pathogenic microbes. Artif. Cells Nanomed. Biotechnol. 2020, 48, 672–682. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Allam, N.G.; Ismail, G.A.; El-Gemizy, W.M.; Salem, M.A. Biosynthesis of silver nanoparticles by cell-free extracts from some bacteria species for dye removal from wastewater. Biotechnol. Lett. 2019, 41, 379–389. [Google Scholar] [CrossRef] [PubMed]
  24. Yu, X.; Li, J.; Mu, D.; Zhang, H.; Liu, Q.; Chen, G. Green synthesis and characterizations of silver nanoparticles with enhanced antibacterial properties by secondary metabolites of Bacillus subtilis (SDUM301120). Green Chem. Lett. Rev. 2021, 14, 189–202. [Google Scholar] [CrossRef]
  25. Sudarsan, S.; Shankar, M.K.; Motatis, A.K.B.; Shankar, S.; Krishnappa, D.; Mohan, C.; Rangappa, K.; Gupta, V.; Siddaiah, C. Green Synthesis of Silver Nanoparticles by Cytobacillus firmus Isolated from the Stem Bark of Terminalia arjuna and Their Antimicrobial Activity. Biomolecules 2021, 11, 259. [Google Scholar] [CrossRef] [PubMed]
  26. El-Bendary, M.A.; Abdelraof, M.; Moharam, M.E.; Elmahdy, E.M.; Allam, M.A. Potential of silver nanoparticles synthesized using low active mosquitocidal Lysinibacillus sphaericus as novel antimicrobial agents. Prep. Biochem. Biotechnol. 2021, 1–10. [Google Scholar] [CrossRef] [PubMed]
  27. Saravanan, M.; Barik, S.K.; MubarakAli, D.; Prakash, P.; Pugazhendhi, A. Synthesis of silver nanoparticles from Bacillus brevis (NCIM 2533) and their antibacterial activity against pathogenic bacteria. Microb. Pathog. 2018, 116, 221–226. [Google Scholar] [CrossRef] [PubMed]
  28. Elbeshehy, E.K.F.; Elazzazy, A.M.; Aggelis, G. Silver nanoparticles synthesis mediated by new isolates of Bacillus spp., nanoparticle characterization and their activity against Bean Yellow Mosaic Virus and human pathogens. Front. Microbiol. 2015, 6, 453. [Google Scholar] [CrossRef] [Green Version]
  29. Chaudhary, R.; Nawaz, K.; Khan, A.K.; Hano, C.; Abbasi, B.H.; Anjum, S. An Overview of the Algae-Mediated Biosynthesis of Nanoparticles and Their Biomedical Applications. Biomolecules 2020, 10, 1498. [Google Scholar] [CrossRef] [PubMed]
  30. Jena, J.; Pradhan, N.; Nayak, R.R.; Dash, B.P.; Sukla, L.B.; Panda, P.K.; Mishra, B.K. Microalga Scenedesmus sp.: A Potential Low-Cost Green Machine for Silver Nanoparticle Synthesis. J. Microbiol. Biotechnol. 2014, 24, 522–533. [Google Scholar] [CrossRef] [PubMed]
  31. Yilmaz Öztürk, B. Intracellular and extracellular green synthesis of silver nanoparticles using Desmodesmus sp.: Their antibacterial and antifungal effects. Caryologia 2019, 72, 29–43. [Google Scholar] [CrossRef]
  32. Senapati, S.; Syed, A.; Moeez, S.; Kumar, A.; Ahmad, A. Intracellular synthesis of gold nanoparticles using alga Tetraselmis kochinensis. Mater. Lett. 2012, 79, 116–118. [Google Scholar] [CrossRef]
  33. Merin, D.D.; Prakash, S.; Bhimba, B.V. Antibacterial screening of silver nanoparticles synthesized by marine micro algae. Asian Pac. J. Trop. Med. 2010, 3, 797–799. [Google Scholar] [CrossRef] [Green Version]
  34. Losic, D.; Mitchell, J.G.; Voelcker, N. Diatomaceous Lessons in Nanotechnology and Advanced Materials. Adv. Mater. 2009, 21, 2947–2958. [Google Scholar] [CrossRef]
  35. Aguirre, L.E.; Ouyang, L.; Elfwing, A.; Hedblom, M.; Wulff, A.; Inganäs, O. Diatom frustules protect DNA from ultraviolet light. Sci. Rep. 2018, 8, 5138. [Google Scholar] [CrossRef] [Green Version]
  36. Sinha, S.N.; Paul, D.; Halder, N.; Sengupta, D.; Patra, S.K. Green synthesis of silver nanoparticles using fresh water green alga Pithophora oedogonia (Mont.) Wittrock and evaluation of their antibacterial activity. Appl. Nanosci. 2014, 5, 703–709. [Google Scholar] [CrossRef] [Green Version]
  37. Khanna, P.; Kaur, A.; Goyal, D. Algae-based metallic nanoparticles: Synthesis, characterization and applications. J. Microbiol. Methods 2019, 163, 105656. [Google Scholar] [CrossRef]
  38. El-Rafie, H.; El-Rafie, M.; Zahran, M. Green synthesis of silver nanoparticles using polysaccharides extracted from marine macro algae. Carbohydr. Polym. 2013, 96, 403–410. [Google Scholar] [CrossRef] [PubMed]
  39. Ulagesan, S.; Nam, T.-J.; Choi, Y.-H. Biogenic preparation and characterization of Pyropia yezoensis silver nanoparticles (P.y AgNPs) and their antibacterial activity against Pseudomonas aeruginosa. Bioprocess Biosyst. Eng. 2020, 44, 443–452. [Google Scholar] [CrossRef]
  40. Moshfegh, A.; Jalali, A.; Salehzadeh, A.; Jozani, A.S. Biological synthesis of silver nanoparticles by cell-free extract of Polysiphonia algae and their anticancer activity against breast cancer MCF-7 cell lines. Micro Nano Lett. 2019, 14, 581–584. [Google Scholar] [CrossRef]
  41. Sharma, G.; Jasuja, N.D.; Kumar, M.; Ali, M.I. Biological Synthesis of Silver Nanoparticles by Cell-Free Extract of Spirulina platensis. J. Nanotechnol. 2015, 2015, 132675. [Google Scholar] [CrossRef] [Green Version]
  42. Rajkumar, R.; Ezhumalai, G.; Gnanadesigan, M. A green approach for the synthesis of silver nanoparticles by Chlorella vulgaris and its application in photocatalytic dye degradation activity. Environ. Technol. Innov. 2020, 21, 101282. [Google Scholar] [CrossRef]
  43. Murugesan, S.; Bhuvaneswari, S.; Sivamurugan, V. Green Synthesis, Characterization of Silver Nanoparticles of a Marine Red Alga Spyridia Fusiformis and Their Antibacterial Activity. Int. J. Pharm. Pharm. Sci. 2017, 9, 192. [Google Scholar] [CrossRef] [Green Version]
  44. Elgamouz, A.; Idriss, H.; Nassab, C.; Bihi, A.; Bajou, K.; Hasan, K.; Abu Haija, M.; Patole, S.P. Green Synthesis, Characterization, Antimicrobial, Anti-Cancer, and Optimization of Colorimetric Sensing of Hydrogen Peroxide of Algae Extract Capped Silver Nanoparticles. Nanomaterials 2020, 10, 1861. [Google Scholar] [CrossRef]
  45. LewisOscar, F.; Nithya, C.; Vismaya, S.; Arunkumar, M.; Pugazhendhi, A.; Nguyen-Tri, P.; Alharbi, S.A.; Alharbi, N.S.; Thajuddin, N. In Vitro analysis of green fabricated silver nanoparticles (AgNPs) against Pseudomonas aeruginosa PA14 biofilm formation, their application on urinary catheter. Prog. Org. Coat. 2020, 151, 106058. [Google Scholar] [CrossRef]
  46. Berdy, J. Bioactive microbial metabolites: A personal view. J. Antibiot. 2005, 58, 1–26. [Google Scholar] [CrossRef] [Green Version]
  47. Azmath, P.; Baker, S.; Rakshith, D.; Satish, S. Mycosynthesis of silver nanoparticles bearing antibacterial activity. Saudi Pharm. J. 2016, 24, 140–146. [Google Scholar] [CrossRef] [Green Version]
  48. Khan, N.T.; Jameel, J.; Rheman, S.U.A. An Overview: Biological Organisms that Serves as Nanofactories for Metallic Nanoparticles Synthesis and Fungi Being the Most Appropriate. Bioceram. Dev. Appl. 2017, 7, 101. [Google Scholar] [CrossRef] [Green Version]
  49. Ahmad, A.; Mukherjee, P.; Senapati, S.; Mandal, D.; Khan, M.; Kumar, R.; Sastry, M. Extracellular biosynthesis of silver nanoparticles using the fungus Fusarium oxysporum. Colloids Surf. B Biointerfaces 2003, 28, 313–318. [Google Scholar] [CrossRef]
  50. Elamawi, R.M.; Al-Harbi, R.E.; Hendi, A.A. Biosynthesis and characterization of silver nanoparticles using Trichoderma longibrachiatum and their effect on phytopathogenic fungi. Egypt. J. Biol. Pest Control 2018, 28, 28. [Google Scholar] [CrossRef] [Green Version]
  51. Baymiller, M.; Huang, F.; Rogelj, S. Rapid one-step synthesis of gold nanoparticles using the ubiquitous coenzyme NADH. Matters 2017, 3. [Google Scholar] [CrossRef] [Green Version]
  52. Zomorodian, K.; Pourshahid, S.; Sadatsharifi, A.; Mehryar, P.; Pakshir, K.; Rahimi, M.J.; Monfared, A.A. Biosynthesis and Characterization of Silver Nanoparticles by Aspergillus Species. BioMed Res. Int. 2016, 2016, 1–6. [Google Scholar] [CrossRef] [Green Version]
  53. Ranjani, S.; Ahmed, M.S.; Mohd, A.; Kumar, N.S.; Ruckmani, K.; Hemalatha, S. Synthesis, characterization and applications of endophytic fungal nanoparticles. Inorg. Nano Met. Chem. 2020, 51, 280–287. [Google Scholar] [CrossRef]
  54. Ammar, H.A.; El Aty, A.A.A.; El Awdan, S.A. Extracellular myco-synthesis of nano-silver using the fermentable yeasts Pichia kudriavzeviiHA-NY2 and Saccharomyces uvarumHA-NY3, and their effective biomedical applications. Bioprocess Biosyst. Eng. 2021, 44, 841–854. [Google Scholar] [CrossRef]
  55. Aygün, A.; Özdemir, S.; Gülcan, M.; Yalçın, M.S.; Uçar, M.; Şen, F. Characterization and antioxidant-antimicrobial activity of silver nanoparticles synthesized using Punica granatum extract. Int. J. Environ. Sci. Technol. 2021, 1–8. [Google Scholar] [CrossRef]
  56. Ramos, M.M.; Morais, E.D.S.; Sena, I.D.S.; Lima, A.L.; De Oliveira, F.R.; De Freitas, C.M.; Fernandes, C.P.; De Carvalho, J.C.T.; Ferreira, I.M. Silver nanoparticle from whole cells of the fungi Trichoderma spp. isolated from Brazilian Amazon. Biotechnol. Lett. 2020, 42, 833–843. [Google Scholar] [CrossRef] [PubMed]
  57. Ma, L.; Lv, S.; Tang, J.; Liu, J.; Li, W.; Deng, J.; Deng, Y.; Du, J.; Liu, X.; Zeng, X. Study on bioactive molecules involved in extracellular biosynthesis of silver nanoparticles by Penicillium aculeatum Su1. Mater. Express 2019, 9, 475–483. [Google Scholar] [CrossRef]
  58. Rodríguez-Serrano, C.; Guzmán-Moreno, J.; Ángeles-Chávez, C.; Rodríguez-González, V.; Ortega-Sigala, J.J.; Ramírez-Santoyo, R.M.; Vidales-Rodríguez, L.E. Biosynthesis of silver nanoparticles by Fusarium scirpi and its potential as antimicrobial agent against uropathogenic Escherichia coli biofilms. PLoS ONE 2020, 15, e0230275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Farrag, H.M.M.; Mostafa, F.A.A.M.; Mohamed, M.E.; Huseein, E.A.M. Green biosynthesis of silver nanoparticles by Aspergillus niger and its antiamoebic effect against Allovahlkampfia spelaea trophozoite and cyst. Exp. Parasitol. 2020, 219, 108031. [Google Scholar] [CrossRef] [PubMed]
  60. Aziz, N.; Faraz, M.; Sherwani, M.A.; Fatma, T.; Prasad, R. Illuminating the Anticancerous Efficacy of a New Fungal Chassis for Silver Nanoparticle Synthesis. Front. Chem. 2019, 7, 65. [Google Scholar] [CrossRef] [Green Version]
  61. Alghuthaymi, M.A.; Almoammar, H.; Rai, M.; Said-Galiev, E.; Abd-Elsalam, K.A. Myconanoparticles: Synthesis and their role in phytopathogens management. Biotechnol. Biotechnol. Equip. 2015, 29, 221–236. [Google Scholar] [CrossRef]
  62. Mousavi, S.M.; Hashemi, S.A.; Ghasemi, Y.; Atapour, A.; Amani, A.M.; Dashtaki, A.S.; Babapoor, A.; Arjmand, O. Green synthesis of silver nanoparticles toward bio and medical applications: Review study. Artif. Cells Nanomed. Biotechnol. 2018, 46, S855–S872. [Google Scholar] [CrossRef] [Green Version]
  63. Jain, S.; Mehata, M.S. Medicinal Plant Leaf Extract and Pure Flavonoid Mediated Green Synthesis of Silver Nanoparticles and their Enhanced Antibacterial Property. Sci. Rep. 2017, 7, 1–13. [Google Scholar] [CrossRef] [PubMed]
  64. Ahmed, S.; Saifullah, A.M.; Swami, B.L.; Ikram, S. Green synthesis of silver nanoparticles using Azadirachta indica aqueous leaf extract. J. Radiat. Res. Appl. Sci. 2016, 9, 1–7. [Google Scholar] [CrossRef] [Green Version]
  65. Moradi, F.; Sedaghat, S.; Arab-Salmanabadi, S.; Moradi, O. Biosynthesis of silver-montmorillonite nanocomposites using Ocimum Basilicum and Teucrium Polium; A comparative study. Mater. Res. Express 2019, 6, 125008. [Google Scholar] [CrossRef]
  66. Ghramh, H.A.; Ibrahim, E.H.; Kilnay, M.; Ahmad, Z.; Alhag, S.K.; Khan, K.A.; Taha, R.; Asiri, F.M. Silver Nanoparticle Production by Ruta graveolens and Testing Its Safety, Bioactivity, Immune Modulation, Anticancer, and Insecticidal Potentials. Bioinorg. Chem. Appl. 2020, 2020, 1–11. [Google Scholar] [CrossRef]
  67. Palithya, S.; Gaddam, S.A.; Kotakadi, V.S.; Penchalaneni, J.; Challagundla, V.N. Biosynthesis of silver nanoparticles using leaf extract of Decaschistia crotonifolia and its antibacterial, antioxidant, and catalytic applications. Green Chem. Lett. Rev. 2021, 14, 137–152. [Google Scholar] [CrossRef]
  68. Singh, R.; Hano, C.; Nath, G.; Sharma, B. Green biosynthesis of silver nanoparticles using leaf extract of Carissa carandas L. and their antioxidant and antimicrobial activity against human pathogenic bacteria. Biomolecules 2021, 11, 299. [Google Scholar] [CrossRef] [PubMed]
  69. Mohammed, A.E.; Al-Qahtani, A.; Al-Mutairi, A.; Al-Shamri, B.; Aabed, K.F. Antibacterial and Cytotoxic Potential of Biosynthesized Silver Nanoparticles by Some Plant Extracts. Nanomaterials 2018, 8, 382. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Agarwal, H.; Kumar, S.V.; RajeshKumar, S. Antidiabetic effect of silver nanoparticles synthesized using lemongrass (Cymbopogon Citratus) through conventional heating and microwave irradiation approach. J. Microbiol. Biotechnol. Food Sci. 2018, 7, 371–376. [Google Scholar] [CrossRef]
  71. Al-Otibi, F.; Al-Ahaidib, R.A.; Alharbi, R.I.; Al-Otaibi, R.M.; Albasher, G. Antimicrobial Potential of Biosynthesized Silver Nanoparticles by Aaronsohnia factorovskyi Extract. Molecules 2020, 26, 130. [Google Scholar] [CrossRef]
  72. Alghoraibi, I.; Soukkarieh, C.; Zein, R.; Alahmad, A.; Walter, J.-G.; Daghestani, M. Aqueous extract of Eucalyptus camaldulensis leaves as reducing and capping agent in biosynthesis of silver nanoparticles. Inorg. Nano Met. Chem. 2020, 50, 895–902. [Google Scholar] [CrossRef]
  73. Raj, S.; Singh, H.; Trivedi, R.; Soni, V. Biogenic synthesis of AgNPs employing Terminalia arjuna leaf extract and its efficacy towards catalytic degradation of organic dyes. Sci. Rep. 2020, 10, 1–10. [Google Scholar] [CrossRef]
  74. Panja, S.; Choudhuri, I.; Khanra, K.; Pati, B.; Bhattacharyya, N. Biological and Photocatalytic Activity of Silver Nanoparticle Synthesized from Ehretia laevis Roxb. Leaves Extract. Nano Biomed. Eng. 2020, 12, 104–113. [Google Scholar] [CrossRef]
  75. Nilavukkarasi, M.; Vijayakumar, S.; Kumar, S.P. Biological synthesis and characterization of silver nanoparticles with Capparis zeylanica L. leaf extract for potent antimicrobial and anti proliferation efficiency. Mater. Sci. Energy Technol. 2020, 3, 371–376. [Google Scholar] [CrossRef]
  76. Hamidi, A.; Yazdi, M.E.T.; Amiri, M.S.; Hosseini, H.A.; Darroudi, M. Biological synthesis of silver nanoparticles in Tribulus terrestris L. extract and evaluation of their photocatalyst, antibacterial, and cytotoxicity effects. Res. Chem. Intermed. 2019, 45, 2915–2925. [Google Scholar] [CrossRef]
  77. Parvathiraja, C.; Shailajha, S.; Shanavas, S.; Gurung, J. Biosynthesis of silver nanoparticles by Cyperus pangorei and its potential in structural, optical and catalytic dye degradation. Appl. Nanosci. 2020, 11, 477–491. [Google Scholar] [CrossRef]
  78. Çalhan, S.D.; Gündoğan, M. Biosynthesis of silver nanoparticles using Onosma sericeum Willd. and evaluation of their catalytic properties and antibacterial and cytotoxic activity. Turk. J. Chem. 2020, 44, 1587–1600. [Google Scholar] [CrossRef] [PubMed]
  79. Chandirika, J.U.; Annadurai, G. Biosynthesis and characterization of silver nanoparticles using leaf extract Abutilon indicum. Glob. J. Biotechnol. Biochem. 2018, 13, 7–11. [Google Scholar] [CrossRef]
  80. Rasool, S.; Raza, M.A.; Manzoor, F.; Kanwal, Z.; Riaz, S.; Iqbal, M.J.; Naseem, S. Biosynthesis, characterization and anti-dengue vector activity of silver nanoparticles prepared from Azadirachta indica and Citrullus colocynthis. R. Soc. Open Sci. 2020, 7, 200540. [Google Scholar] [CrossRef] [PubMed]
  81. Siraj, M.; Shah, Z.A.; Ullah, S.; Bibi, H.; Suleman, M.; Zia, A.; Masood, T.; Iqbal, Z.; Iqbal, M. Biosynthesized silver nanoparticles from shoot and seed extracts of Asphodelus tenufolius for heavy metal sensing. Science 2020, 46, 697. [Google Scholar] [CrossRef]
  82. Aravind, M.; Ahmad, A.; Ahmad, I.; Amalanathan, M.; Naseem, K.; Mary, S.M.M.; Parvathiraja, C.; Hussain, S.; Algarni, T.S.; Pervaiz, M.; et al. Critical green routing synthesis of silver NPs using jasmine flower extract for biological activities and photocatalytical degradation of methylene blue. J. Environ. Chem. Eng. 2020, 9, 104877. [Google Scholar] [CrossRef]
  83. Patra, J.K.; Das, G.; Shin, H.-S. Facile green biosynthesis of silver nanoparticles using Pisum sativum L. outer peel aqueous extract and its antidiabetic, cytotoxicity, antioxidant, and antibacterial activity. Int. J. Nanomed. 2019, 14, 6679–6690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Azizi, M.; Sedaghat, S.; Tahvildari, K.; Derakhshi, P.; Ghaemi, A. Green biosynthesis of silver nanoparticles with Eryngium caucasicum Trautv aqueous extract. Inorg. Nano Met. Chem. 2020, 50, 429–436. [Google Scholar] [CrossRef]
  85. Das, B.; De, A.; Podder, S.; Das, S.; Ghosh, C.K.; Samanta, A. Green biosynthesis of silver nanoparticles using Dregea volubilis flowers: Characterization and evaluation of antioxidant, antidiabetic and antibacterial activity. Inorg. Nano Met. Chem. 2020, 51, 1066–1079. [Google Scholar] [CrossRef]
  86. Rasheed, T.; Bilal, M.; Iqbal, H.; Li, C. Green biosynthesis of silver nanoparticles using leaves extract of Artemisia vulgaris and their potential biomedical applications. Colloids Surf. B Biointerfaces 2017, 158, 408–415. [Google Scholar] [CrossRef] [PubMed]
  87. Kanniah, P.; Radhamani, J.; Chelliah, P.; Muthusamy, N.; Balasingh, E.J.J.S.; Thangapandi, J.R.; Balakrishnan, S.; Shanmugam, R. Green Synthesis of Multifaceted Silver Nanoparticles Using the Flower Extract of Aerva lanata and Evaluation of Its Biological and Environmental Applications. ChemistrySelect 2020, 5, 2322–2331. [Google Scholar] [CrossRef]
  88. Mehwish, H.M.; Rajoka, M.S.R.; Xiong, Y.; Cai, H.; Aadil, R.M.; Mahmood, Q.; He, Z.; Zhu, Q. Green synthesis of a silver nanoparticle using Moringa oleifera seed and its applications for antimicrobial and sun-light mediated photocatalytic water detoxification. J. Environ. Chem. Eng. 2021, 9, 105290. [Google Scholar] [CrossRef]
  89. Ekennia, A.C.; Uduagwu, D.N.; Nwaji, N.N.; Olowu, O.J.; Nwanji, O.L.; Ejimofor, M.; Sonde, C.U.; Oje, O.O.; Igwe, D.O. Green synthesis of silver nanoparticles using leaf extract of Euphorbia sanguine: An In Vitro study of its photocatalytic and melanogenesis inhibition activity. Inorg. Nano Met. Chem. 2021, 1–9. [Google Scholar] [CrossRef]
  90. Parthiban, E.; Manivannan, N.; Ramanibai, R.; Mathivanan, N. Green synthesis of silver-nanoparticles from Annona reticulata leaves aqueous extract and its mosquito larvicidal and anti-microbial activity on human pathogens. Biotechnol. Rep. 2018, 21, e00297. [Google Scholar] [CrossRef]
  91. Jalilian, F.; Chahardoli, A.; Sadrjavadi, K.; Fattahi, A.; Shokoohinia, Y. Green synthesized silver nanoparticle from Allium ampeloprasum aqueous extract: Characterization, antioxidant activities, antibacterial and cytotoxicity effects. Adv. Powder Technol. 2020, 31, 1323–1332. [Google Scholar] [CrossRef]
  92. Gupta, S.D.; Kohli, P. LED Exposure Modulates the Biosynthesis of Silver Nanoparticles from Root Tuber Extract of Chlorophytum borivilianum and their Phytotoxicty. J. Clust. Sci. 2020, 1–14. [Google Scholar] [CrossRef]
  93. Jahan, I.; Erci, F.; Isildak, I. Microwave-Assisted Green Synthesis of Non-Cytotoxic Silver Nanoparticles Using the Aqueous Extract of Rosa santana (rose) Petals and Their Antimicrobial Activity. Anal. Lett. 2019, 52, 1860–1873. [Google Scholar] [CrossRef]
  94. Chandhirasekar, K.; Thendralmanikandan, A.; Thangavelu, P.; Nguyen, B.-S.; Nguyen, T.-A.; Sivashanmugan, K.; Nareshkumar, A.; Nguyen, V.-H. Plant-extract-assisted green synthesis and its larvicidal activities of silver nanoparticles using leaf extract of Citrus medica, Tagetes lemmonii, and Tarenna asiatica. Mater. Lett. 2020, 287, 129265. [Google Scholar] [CrossRef]
  95. Hamedi, S.; Shojaosadati, S.A. Rapid and green synthesis of silver nanoparticles using Diospyros lotus extract: Evaluation of their biological and catalytic activities. Polyhedron 2019, 171, 172–180. [Google Scholar] [CrossRef]
  96. Huang, F.; Long, Y.; Liang, Q.; Purushotham, B.; Swamy, M.K.; Duan, Y. Safed Musli (Chlorophytum borivilianum L.) Callus-Mediated Biosynthesis of Silver Nanoparticles and Evaluation of their Antimicrobial Activity and Cytotoxicity against Human Colon Cancer Cells. J. Nanomater. 2019, 2019, 1–8. [Google Scholar] [CrossRef]
  97. Nayaka, S.; Bhat, M.P.; Chakraborty, B.; Pallavi, S.S.; Airodagi, D.; Muthuraj, R.; Halaswamy, H.M.; Dhanyakumara, S.B.; Shashiraj, K.N.; Kupaneshi, K.N.S.A.C. Seed Extract-mediated Synthesis of Silver Nanoparticles from Putranjiva roxburghii Wall., Phytochemical Characterization, Antibacterial Activity and Anticancer Activity Against MCF-7 Cell Line. Indian J. Pharm. Sci. 2020, 82, 260–269. [Google Scholar] [CrossRef]
  98. Azhar, N.A.; Ghozali, S.Z.; Abu Bakar, S.A.; Lim, V.; Ahmad, N.H. Suppressing growth, migration, and invasion of human hepatocellular carcinoma HepG2 cells by Catharanthus roseus-silver nanoparticles. Toxicol. Vitr. 2020, 67, 104910. [Google Scholar] [CrossRef] [PubMed]
  99. Vanti, G.L.; Kurjogi, M.; Basavesha, K.N.; Teradal, N.L.; Masaphy, S.; Nargund, V.B. Synthesis and antibacterial activity of Solanum torvum mediated silver nanoparticle against Xanthomonas axonopodis pv. punicae and Ralstonia solanacearum. J. Biotechnol. 2020, 309, 20–28. [Google Scholar] [CrossRef]
  100. Shah, Z.; Hassan, S.; Shaheen, K.; Khan, S.A.; Gul, T.; Anwar, Y.; Al-Shaeri, M.A.; Khan, M.; Khan, R.; Haleem, M.A.; et al. Synthesis of AgNPs coated with secondary metabolites of Acacia nilotica: An efficient antimicrobial and detoxification agent for environmental toxic organic pollutants. Mater. Sci. Eng. C 2020, 111, 110829. [Google Scholar] [CrossRef] [PubMed]
  101. Brito, T.K.; Viana, R.L.S.; Moreno, C.J.G.; Barbosa, J.D.S.; Júnior, F.L.D.S.; de Medeiros, M.J.C.; Melo-Silveira, R.F.; Almeida-Lima, J.; Pontes, D.D.L.; Silva, M.S.; et al. Synthesis of Silver Nanoparticle Employing Corn Cob Xylan as a Reducing Agent with Anti-Trypanosoma cruzi Activity. Int. J. Nanomed. 2020, 15, 965–979. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Sankarganesh, P.; Kumar, A.G.; Parthasarathy, V.; Joseph, B.; Priyadharsini, G.; Anbarasan, R. Synthesis of Murraya koenigii Mediated Silver Nanoparticles and Their In Vitro and In Vivo Biological Potential. J. Inorg. Organomet. Polym. Mater. 2021, 1–9. [Google Scholar] [CrossRef]
  103. Vijayakumar, S.; Divya, M.; Vaseeharan, B.; Chen, J.; Biruntha, M.; Silva, L.P.; Durán-Lara, E.F.; Shreema, K.; Ranjan, S.; Dasgupta, N. Biological Compound Capping of Silver Nanoparticle with the Seed Extracts of Blackcumin (Nigella sativa): A Potential Antibacterial, Antidiabetic, Anti-inflammatory, and Antioxidant. J. Inorg. Organomet. Polym. Mater. 2020, 31, 624–635. [Google Scholar] [CrossRef]
  104. Ahmad, S.; Tauseef, I.; Haleem, K.S.; Khan, K.; Shahzad, M.; Ali, M.; Sultan, F. Synthesis of silver nanoparticles using leaves of Catharanthus roseus and their antimicrobial activity. Appl. Nanosci. 2019, 10, 4459–4464. [Google Scholar] [CrossRef]
  105. Gurunathan, S.; Han, J.W.; Kwon, D.-N.; Kim, J.-H. Enhanced antibacterial and anti-biofilm activities of silver nanoparticles against Gram-negative and Gram-positive bacteria. Nanoscale Res. Lett. 2014, 9, 373. [Google Scholar] [CrossRef] [Green Version]
  106. Ritchie, R.H. Plasma Losses by Fast Electrons in Thin Films. Phys. Rev. 1957, 106, 874–881. [Google Scholar] [CrossRef]
  107. Barnes, W.L.; Dereux, A.; Ebbesen, T.W. Surface plasmon subwavelength optics. Nat. Cell Biol. 2003, 424, 824–830. [Google Scholar] [CrossRef]
  108. Haes, A.; Van Duyne, R.P. A unified view of propagating and localized surface plasmon resonance biosensors. Anal. Bioanal. Chem. 2004, 379, 920–930. [Google Scholar] [CrossRef]
  109. Petryayeva, E.; Krull, U.J. Localized surface plasmon resonance: Nanostructures, bioassays and biosensing—A review. Anal. Chim. Acta 2011, 706, 8–24. [Google Scholar] [CrossRef] [PubMed]
  110. Liz-Marzán, L.M. Tailoring Surface Plasmons through the Morphology and Assembly of Metal Nanoparticles. Langmuir 2006, 22, 32–41. [Google Scholar] [CrossRef]
  111. Jensen, T.R.; Malinsky, M.D.; Haynes, C.; Van Duyne, R.P. Nanosphere Lithography: Tunable Localized Surface Plasmon Resonance Spectra of Silver Nanoparticles. J. Phys. Chem. B 2000, 104, 10549–10556. [Google Scholar] [CrossRef]
  112. Yguerabide, J.; Yguerabide, E.E. Light-Scattering Submicroscopic Particles as Highly Fluorescent Analogs and Their Use as Tracer Labels in Clinical and Biological Applications. Anal. Biochem. 1998, 262, 157–176. [Google Scholar] [CrossRef]
  113. Choi, Y.; Kang, T.; Lee, L.P. Plasmon Resonance Energy Transfer (PRET)-based Molecular Imaging of Cytochromecin Living Cells. Nano Lett. 2008, 9, 85–90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Jiang, Z.; Wen, G.; Luo, Y.; Zhang, X.; Liu, Q.; Liang, A. A new silver nanorod SPR probe for detection of trace benzoyl peroxide. Sci. Rep. 2014, 4, 5323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Zhuo, X.; Henriksen-Lacey, M.; de Aberasturi, D.J.; Sánchez-Iglesias, A.; Liz-Marzán, L.M. Shielded Silver Nanorods for Bioapplications. Chem. Mater. 2020, 32, 5879–5889. [Google Scholar] [CrossRef]
  116. Zhang, X.-F.; Liu, Z.-G.; Shen, W.; Gurunathan, S. Silver nanoparticles: Synthesis, characterization, properties, applications, and therapeutic approaches. Int. J. Mol. Sci. 2016, 17, 1534. [Google Scholar] [CrossRef] [PubMed]
  117. Vijayaraghavan, K.; Ashokkumar, T. Plant-mediated biosynthesis of metallic nanoparticles: A review of literature, factors affecting synthesis, characterization techniques and applications. J. Environ. Chem. Eng. 2017, 5, 4866–4883. [Google Scholar] [CrossRef]
  118. Henglein, A. Physicochemical properties of small metal particles in solution: Microelectrode reactions, chemisorption, composite metal particles, and the atom-to-metal transition. J. Phys. Chem. 1993, 97, 5457–5471. [Google Scholar] [CrossRef]
  119. Sastry, M.; Mayya, K.; Bandyopadhyay, K. pH Dependent changes in the optical properties of carboxylic acid derivatized silver colloidal particles. Colloids Surf. A Physicochem. Eng. Asp. 1997, 127, 221–228. [Google Scholar] [CrossRef]
  120. Haiss, W.; Thanh, N.T.K.; Aveyard, J.; Fernig, D. Determination of Size and Concentration of Gold Nanoparticles from UV−Vis Spectra. Anal. Chem. 2007, 79, 4215–4221. [Google Scholar] [CrossRef] [PubMed]
  121. Albert, H.M.; Lohitha, T.; Alagarsamy, K.; Gonsago, C.; Vishwakarma, V. Performance of ZnSO4 doped CeO2 nanoparticles and their antibacterial mechanism. Mater. Today Proc. 2021. [Google Scholar] [CrossRef]
  122. Akintelu, S.A.; Bo, Y.; Folorunso, A.S. A Review on Synthesis, Optimization, Mechanism, Characterization, and Antibacterial Application of Silver Nanoparticles Synthesized from Plants. J. Chem. 2020, 2020, 3189043. [Google Scholar] [CrossRef]
  123. Rajeshkumar, S.; Bharath, L.V. Mechanism of plant-mediated synthesis of silver nanoparticles—A review on biomolecules involved, characterisation and antibacterial activity. Chem. Biol. Interact. 2017, 273, 219–227. [Google Scholar] [CrossRef]
  124. Bhambure, R.; Bule, M.; Shaligram, N.; Kamat, M.; Singhal, R. Extracellular Biosynthesis of Gold Nanoparticles using Aspergillus niger—Its Characterization and Stability. Chem. Eng. Technol. 2009, 32, 1036–1041. [Google Scholar] [CrossRef]
  125. Alamdari, S.; Ghamsari, M.S.; Lee, C.; Han, W.; Park, H.-H.; Tafreshi, M.J.; Afarideh, H.; Ara, M.H.M. Preparation and Characterization of Zinc Oxide Nanoparticles Using Leaf Extract of Sambucus ebulus. Appl. Sci. 2020, 10, 3620. [Google Scholar] [CrossRef]
  126. Waseda, Y.; Matsubara, E.; Shinoda, K. X-ray Diffraction Crystallography; Springer: Berlin/Heidelberg, Germany, 2011. [Google Scholar]
  127. Das, R.; Nath, S.S.; Chakdar, D.; Gope, G.; Bhattacharjee, R. Preparation of silver nanoparticles and their characterization. J. Nanotechnol. 2009, 1–6. [Google Scholar] [CrossRef]
  128. Cao, G.; Wang, Y. Nanostructures and Nanomaterials; World Scientific Series in Nanoscience and Nanotechnology; World Scientific: Singapore, 2011; Volume 2, ISBN 978-981-4322-50-8. [Google Scholar]
  129. Chapman, H.N.; Fromme, P.; Barty, A.; White, T.A.; Kirian, R.; Aquila, A.; Hunter, M.S.; Schulz, J.; DePonte, D.P.; Weierstall, U.; et al. Femtosecond X-ray protein nanocrystallography. Nat. Cell Biol. 2011, 470, 73–77. [Google Scholar] [CrossRef] [PubMed]
  130. Pawley, J. The Development of Field-Emission Scanning Electron Microscopy for Imaging Biological Surfaces. Scanning 1997, 19, 324–336. [Google Scholar] [PubMed]
  131. Yao, H.; Kimura, K. Field emission scanning electron microscopy for structural characterization of 3d gold nanoparticle superlattices. In Modern Research and Educational Topics in Microscopy; Méndez-Vilas, A., Díaz, J., Eds.; Formatex Research Center: Badajoz, Spain, 2007; pp. 568–575. [Google Scholar]
  132. Anandalakshmi, K.; Venugobal, J.; Ramasamy, V. Characterization of silver nanoparticles by green synthesis method using Pedalium murex leaf extract and their antibacterial activity. Appl. Nanosci. 2015, 6, 399–408. [Google Scholar] [CrossRef] [Green Version]
  133. Pasricha, R.; Bala, T.; Biradar, A.V.; Umbarkar, S.; Sastry, M. Synthesis of Catalytically Active Porous Platinum Nanoparticles by Transmetallation Reaction and Proposition of the Mechanism. Small 2009, 5, 1467–1473. [Google Scholar] [CrossRef] [PubMed]
  134. Asoro, M.A.; Kovar, D.; Ferreira, P.J. In Situ Transmission Electron Microscopy Observations of Sublimation in Silver Nanoparticles. ACS Nano 2013, 7, 7844–7852. [Google Scholar] [CrossRef]
  135. Lin, P.-C.; Lin, S.; Wang, P.C.; Sridhar, R. Techniques for physicochemical characterization of nanomaterials. Biotechnol. Adv. 2014, 32, 711–726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Hall, J.B.; Dobrovolskaia, M.A.; Patri, A.K.; McNeil, S.E. Characterization of nanoparticles for therapeutics. Nanomedicine 2007, 2, 789–803. [Google Scholar] [CrossRef]
  137. Lim, J.; Yeap, S.P.; Che, H.X.; Low, S.C. Characterization of magnetic nanoparticle by dynamic light scattering. Nanoscale Res. Lett. 2013, 8, 381. [Google Scholar] [CrossRef] [Green Version]
  138. Jang, M.-H.; Lee, S.; Hwang, Y.S. Characterization of Silver Nanoparticles under Environmentally Relevant Conditions Using Asymmetrical Flow Field-Flow Fractionation (AF4). PLoS ONE 2015, 10, e0143149. [Google Scholar] [CrossRef] [PubMed]
  139. Tomaszewska, E.; Soliwoda, K.; Kądzioła-Długołęcka, K.; Tkacz-Szczesna, B.; Celichowski, G.; Cichomski, M.; Szmaja, W.; Grobelny, J. Detection Limits of DLS and UV-Vis Spectroscopy in Characterization of Polydisperse Nanoparticles Colloids. J. Nanomater. 2013, 2013, 1–10. [Google Scholar] [CrossRef] [Green Version]
  140. Meva, F.E.; Ntoumba, A.A.; Kedi, P.B.E.; Tchoumbi, E.; Schmitz, A.; Schmolke, L.; Klopotowski, M.; Moll, B.; Kökcam, D.; Mpondo, E.A.M.; et al. Silver and palladium nanoparticles produced using a plant extract as reducing agent, stabilized with an ionic liquid: Sizing by X-ray powder diffraction and dynamic light scattering. J. Mater. Res. Technol. 2019, 8, 1991–2000. [Google Scholar] [CrossRef]
  141. Mourdikoudis, S.; Pallares, R.M.; Thanh, N.T.K. Characterization techniques for nanoparticles: Comparison and complementarity upon studying nanoparticle properties. Nanoscale 2018, 10, 12871–12934. [Google Scholar] [CrossRef] [Green Version]
  142. Bhattacharjee, S. DLS and zeta potential—What they are and what they are not? J. Control. Release 2016, 235, 337–351. [Google Scholar] [CrossRef] [PubMed]
  143. Salomoni, R.; Léo, P.; Montemor, A.; Rinaldi, B.; Rodrigues, M. Antibacterial effect of silver nanoparticles in Pseudomonas aeruginosa. Nanotechnol. Sci. Appl. 2017, 10, 115–121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Fan, W.; Yung, B.; Huang, P.; Chen, X. Nanotechnology for Multimodal Synergistic Cancer Therapy. Chem. Rev. 2017, 117, 13566–13638. [Google Scholar] [CrossRef]
  145. Supraja, N.; Prasad, T.; Soundariya, M.; Babujanarthanam, R. Synthesis, characterization and dose dependent antimicrobial and anti-cancerous activity of phycogenic silver nanoparticles against human hepatic carcinoma (HepG2) cell line. AIMS Environ. Sci. 2016, 3, 425–440. [Google Scholar] [CrossRef]
  146. Das, C.A.; Kumar, V.G.; Dhas, T.S.; Karthick, V.; Govindaraju, K.; Joselin, J.M.; Baalamurugan, J. Antibacterial activity of silver nanoparticles (biosynthesis): A short review on recent advances. Biocatal. Agric. Biotechnol. 2020, 27, 101593. [Google Scholar] [CrossRef]
  147. Prasher, P.; Singh, M.; Mudila, H. Oligodynamic Effect of Silver Nanoparticles: A Review. BioNanoScience 2018, 8, 951–962. [Google Scholar] [CrossRef]
  148. Annamalai, J.; Nallamuthu, T. Green synthesis of silver nanoparticles: Characterization and determination of antibacterial potency. Appl. Nanosci. 2015, 6, 259–265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Feng, Q.L.; Wu, J.; Chen, G.Q.; Cui, F.Z.; Kim, T.N.; Kim, J.O. A mechanistic study of the antibacterial effect of silver ions on Escherichia coli and Staphylococcus aureus. J. Biomed. Mater. Res. 2000, 52, 662–668. [Google Scholar] [CrossRef]
  150. Esmaeillou, M.; Zarrini, G.; Rezaee, M.A.; Mojarrad, J.S.; Bahadori, A. Vancomycin Capped with Silver Nanoparticles as an Antibacterial Agent against Multi-Drug Resistance Bacteria. Adv. Pharm. Bull. 2017, 7, 479–483. [Google Scholar] [CrossRef] [Green Version]
  151. Dizaj, S.M.; Lotfipour, F.; Barzegar-Jalali, M.; Zarrintan, M.H.; Adibkia, K. Antimicrobial activity of the metals and metal oxide nanoparticles. Mater. Sci. Eng. C 2014, 44, 278–284. [Google Scholar] [CrossRef]
  152. Tiwari, P.M.; Vig, K.; Dennis, V.A.; Singh, S.R. Functionalized Gold Nanoparticles and Their Biomedical Applications. Nanomaterials 2011, 1, 31–63. [Google Scholar] [CrossRef]
  153. El Zowalaty, M.; Ibrahim, N.A.; Salama, M.; Shameli, K.; Usman, M.; Zainuddin, N. Synthesis, characterization, and antimicrobial properties of copper nanoparticles. Int. J. Nanomed. 2013, 8, 4467–4479. [Google Scholar] [CrossRef] [Green Version]
  154. Cho, K.-H.; Park, J.-E.; Osaka, T.; Park, S.-G. The study of antimicrobial activity and preservative effects of nanosilver ingredient. Electrochim. Acta 2005, 51, 956–960. [Google Scholar] [CrossRef]
  155. Surwade, P.; Ghildyal, C.; Weikel, C.; Luxton, T.; Peloquin, D.; Fan, X.; Shah, V. Augmented antibacterial activity of ampicillin with silver nanoparticles against methicillin-resistant Staphylococcus aureus (MRSA). J. Antibiot. 2018, 72, 50–53. [Google Scholar] [CrossRef]
  156. Tang, S.; Zheng, J. Antibacterial Activity of Silver Nanoparticles: Structural Effects. Adv. Health Mater. 2018, 7, e1701503. [Google Scholar] [CrossRef] [PubMed]
  157. Yuwen, L.; Sun, Y.-T.; Tan, G.-L.; Xiu, W.-J.; Zhang, Y.; Weng, L.; Teng, Z.; Wang, L.-H. MoS2@polydopamine-Ag nanosheets with enhanced antibacterial activity for effective treatment of Staphylococcus aureus biofilms and wound infection. Nanoscale 2018, 10, 16711–16720. [Google Scholar] [CrossRef]
  158. Deng, H.; McShan, D.; Zhang, Y.; Sinha, S.S.; Arslan, Z.; Ray, P.C.; Yu, H. Mechanistic Study of the Synergistic Antibacterial Activity of Combined Silver Nanoparticles and Common Antibiotics. Environ. Sci. Technol. 2016, 50, 8840–8848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Saratale, G.D.; Saratale, R.G.; Benelli, G.; Kumar, G.; Pugazhendhi, A.; Kim, D.-S.; Shin, H.-S. Anti-diabetic Potential of Silver Nanoparticles Synthesized with Argyreia nervosa Leaf Extract High Synergistic Antibacterial Activity with Standard Antibiotics against Foodborne Bacteria. J. Clust. Sci. 2017, 28, 1709–1727. [Google Scholar] [CrossRef]
  160. Hwang, I.-S.; Hwang, J.H.; Choi, H.; Kim, K.-J.; Lee, D.G. Synergistic effects between silver nanoparticles and antibiotics and the mechanisms involved. J. Med. Microbiol. 2012, 61, 1719–1726. [Google Scholar] [CrossRef] [Green Version]
  161. Khorrami, S.; Zarrabi, A.; Khaleghi, M.; Danaei, M.; Mozafari, M.R. Selective cytotoxicity of green synthesized silver nanoparticles against the MCF-7 tumor cell line and their enhanced antioxidant and antimicrobial properties. Int. J. Nanomed. 2018, 13, 8013–8024. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Ramkumar, V.S.; Pugazhendhi, A.; Gopalakrishnan, K.; Sivagurunathan, P.; Saratale, G.D.; Dung, T.N.B.; Kannapiran, E. Biofabrication and characterization of silver nanoparticles using aqueous extract of seaweed Enteromorpha compressa and its biomedical properties. Biotechnol. Rep. 2017, 14, 1–7. [Google Scholar] [CrossRef]
  163. Durán, N.; Nakazato, G.; Seabra, A.B. Antimicrobial activity of biogenic silver nanoparticles, and silver chloride nanoparticles: An overview and comments. Appl. Microbiol. Biotechnol. 2016, 100, 6555–6570. [Google Scholar] [CrossRef]
  164. Liao, C.; Li, Y.; Tjong, S.C. Bactericidal and Cytotoxic Properties of Silver Nanoparticles. Int. J. Mol. Sci. 2019, 20, 449. [Google Scholar] [CrossRef] [Green Version]
  165. Li, L.; Li, L.; Zhou, X.; Yu, Y.; Li, Z.; Zuo, D.; Wu, Y. Silver nanoparticles induce protective autophagy via Ca2+/CaMKKβ/AMPK/mTOR pathway in SH-SY5Y cells and rat brains. Nanotoxicology 2019, 13, 369–391. [Google Scholar] [CrossRef]
  166. Noronha, V.; Paula, A.J.; Durán, G.; Galembeck, A.; Cogo-Muller, K.; Franz-Montan, M.; Durán, N. Silver nanoparticles in dentistry. Dent. Mater. 2017, 33, 1110–1126. [Google Scholar] [CrossRef]
  167. Shanmuganathan, R.; MubarakAli, D.; Prabakar, D.; Muthukumar, H.; Thajuddin, N.; Kumar, S.; Pugazhendhi, A. An enhancement of antimicrobial efficacy of biogenic and ceftriaxone-conjugated silver nanoparticles: Green approach. Environ. Sci. Pollut. Res. 2017, 25, 10362–10370. [Google Scholar] [CrossRef] [PubMed]
  168. Khorrami, S.; Najafabadi, F.J.; Zarepour, A.; Zarrabi, A. Is Astragalus gossypinus Honey a Natural Antibacterial and Cytotoxic Agent? An Investigation on A. gossypinus Honey Biological Activity and Its Green Synthesized Silver Nanoparticles. BioNanoScience 2019, 9, 603–610. [Google Scholar] [CrossRef]
  169. Jacob, J.M.; John, M.S.; Jacob, A.; Abitha, P.; Kumar, S.S.; Rajan, R.; Natarajan, S.; Pugazhendhi, A. Bactericidal coating of paper towels via sustainable biosynthesis of silver nanoparticles using Ocimum sanctum leaf extract. Mater. Res. Express 2018, 6, 045401. [Google Scholar] [CrossRef]
  170. Meikle, T.G.; Dyett, B.P.; Strachan, J.B.; White, J.; Drummond, C.J.; Conn, C.E. Preparation, Characterization, and Antimicrobial Activity of Cubosome Encapsulated Metal Nanocrystals. ACS Appl. Mater. Interfaces 2020, 12, 6944–6954. [Google Scholar] [CrossRef]
  171. Saravanan, M.; Arokiyaraj, S.; Lakshmi, T.; Pugazhendhi, A. Synthesis of silver nanoparticles from Phenerochaete chrysosporium (MTCC-787) and their antibacterial activity against human pathogenic bacteria. Microb. Pathog. 2018, 117, 68–72. [Google Scholar] [CrossRef] [PubMed]
  172. Yin, I.X.; Yu, O.Y.; Zhao, I.S.; Mei, M.L.; Li, Q.-L.; Tang, J.; Chu, C.-H. Developing biocompatible silver nanoparticles using epigallocatechin gallate for dental use. Arch. Oral Biol. 2019, 102, 106–112. [Google Scholar] [CrossRef] [PubMed]
  173. Pugazhendhi, A.; Prabakar, D.; Jacob, J.M.; Karuppusamy, I.; Saratale, R.G. Synthesis and characterization of silver nanoparticles using Gelidium amansii and its antimicrobial property against various pathogenic bacteria. Microb. Pathog. 2018, 114, 41–45. [Google Scholar] [CrossRef]
  174. Sondi, I.; Salopek-Sondi, B. Silver nanoparticles as antimicrobial agent: A case study on E. coli as a model for Gram-negative bacteria. J. Colloid Interface Sci. 2004, 275, 177–182. [Google Scholar] [CrossRef]
  175. Sharmin, S.; Rahaman, M.; Sarkar, C.; Atolani, O.; Islam, M.T.; Adeyemi, O.S. Nanoparticles as antimicrobial and antiviral agents: A literature-based perspective study. Heliyon 2021, 7, e06456. [Google Scholar] [CrossRef]
  176. Kowalczyk, P.; Szymczak, M.; Maciejewska, M.; Laskowska, M.; Ostaszewski, R.; Skiba, G.; Franiak-Pietryga, I. All That Glitters Is Not Silver—A New Look at Microbiological and Medical Applications of Silver Nanoparticles. Int. J. Mol. Sci. 2021, 22, 854. [Google Scholar] [CrossRef]
  177. Panáček, A.; Kolář, M.; Večeřová, R.; Prucek, R.; Soukupová, J.; Krystof, V.; Hamal, P.; Zboril, R.; Kvítek, L. Antifungal activity of silver nanoparticles against Candida spp. Biomaterials 2009, 30, 6333–6340. [Google Scholar] [CrossRef] [PubMed]
  178. Khan, M.; Khan, A.U.; Alam, M.J.; Park, S.; Alam, M. Biosynthesis of silver nanoparticles and its application against phytopathogenic bacterium and fungus. Int. J. Environ. Anal. Chem. 2019, 100, 1390–1401. [Google Scholar] [CrossRef]
  179. Elgorban, A.M.; El-Samawaty, A.E.-R.M.; Yassin, M.A.; Sayed, S.R.; Adil, S.; Elhindi, K.M.; Bakri, M.; Khan, M. Antifungal silver nanoparticles: Synthesis, characterization and biological evaluation. Biotechnol. Biotechnol. Equip. 2015, 30, 56–62. [Google Scholar] [CrossRef] [Green Version]
  180. Jo, Y.-K.; Kim, B.H.; Jung, G. Antifungal Activity of Silver Ions and Nanoparticles on Phytopathogenic Fungi. Plant Dis. 2009, 93, 1037–1043. [Google Scholar] [CrossRef] [Green Version]
  181. Koduru, J.R.; Kailasa, S.K.; Bhamore, J.R.; Kim, K.-H.; Dutta, T.; Vellingiri, K. Phytochemical-assisted synthetic approaches for silver nanoparticles antimicrobial applications: A review. Adv. Colloid Interface Sci. 2018, 256, 326–339. [Google Scholar] [CrossRef]
  182. WHO (World Health Organization). Coronavirus Disease (COVID-19). Available online: https://www.who.int/emergencies/diseases/novel-coronavirus-2019?gclid=CjwKCAjwhMmEBhBwEiwAXwFoETN8HYVlB-WOMiWVh8S0XnewihrOxP2vCwiVUP36e_7g1ha5xb-mFxoCl-QQAvD_BwE (accessed on 5 May 2021).
  183. Salleh, A.; Naomi, R.; Utami, N.D.; Mohammad, A.W.; Mahmoudi, E.; Mustafa, N.; Fauzi, M.B. The Potential of Silver Nanoparticles for Antiviral and Antibacterial Applications: A Mechanism of Action. Nanomaterials 2020, 10, 1566. [Google Scholar] [CrossRef]
  184. Lv, X.; Wang, P.; Bai, R.; Cong, Y.; Suo, S.; Ren, X.; Chen, C. Inhibitory effect of silver nanomaterials on transmissible virus-induced host cell infections. Biomaterials 2014, 35, 4195–4203. [Google Scholar] [CrossRef] [PubMed]
  185. Azarudeen, R.M.S.T.; Govindarajan, M.; Amsath, A.; Muthukumaran, U.; Benelli, G. Single-Step Biofabrication of Silver Nanocrystals Using Naregamia alata: A Cost Effective and Eco-Friendly Control Tool in the Fight against Malaria, Zika Virus and St. Louis Encephalitis Mosquito Vectors. J. Clust. Sci. 2016, 28, 179–203. [Google Scholar] [CrossRef] [Green Version]
  186. Balagna, C.; Perero, S.; Percivalle, E.; Nepita, E.V.; Ferraris, M. Virucidal effect against coronavirus SARS-CoV-2 of a silver nanocluster/silica composite sputtered coating. Open Ceram. 2020, 1, 100006. [Google Scholar] [CrossRef]
  187. Galdiero, S.; Rai, M.; Gade, A.; Falanga, A.; Incoronato, N.; Russo, L.; Galdiero, M.; Gaikwad, S.; Ingle, A. Antiviral activity of mycosynthesized silver nanoparticles against herpes simplex virus and human parainfluenza virus type 3. Int. J. Nanomed. 2013, 8, 4303–4314. [Google Scholar] [CrossRef] [Green Version]
  188. Dhanasezhian, A.; Srivani, S.; Govindaraju, K.; Parija, P.; Sasikala, S.; Ramesh Kumar, M.R. Anti-herpes simplex virus (HSV-1 and HSV-2) activity of biogenic gold and silver nanoparticles using seaweed Sargassum wightii. Indian J. Geo-Mar. Sci. 2019, 48, 1252–1257. [Google Scholar]
  189. Morris, D.; Ansar, M.; Speshock, J.; Ivanciuc, T.; Qu, Y.; Casola, A.; Garofalo, R.P. Antiviral and Immunomodulatory Activity of Silver Nanoparticles in Experimental RSV Infection. Viruses 2019, 11, 732. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Mayorga, J.L.C.; Randazzo, W.; Fabra, M.J.; Lagaron, J.; Aznar, R.; Sánchez, G. Antiviral properties of silver nanoparticles against norovirus surrogates and their efficacy in coated polyhydroxyalkanoates systems. LWT 2017, 79, 503–510. [Google Scholar] [CrossRef] [Green Version]
  191. Zhang, R.; Lin, Z.; Lui, V.C.; Wong, K.K.; Tam, P.K.; Lee, P.; Lok, C.N.; Lamb, J.R.; Chen, Y.; Xia, H. Silver nanoparticle treatment ameliorates biliary atresia syndrome in rhesus rotavirus inoculated mice. Nanomed. Nanotechnol. Biol. Med. 2017, 13, 1041–1050. [Google Scholar] [CrossRef] [PubMed]
  192. Orlowski, P.; Tomaszewska, E.; Gniadek, M.; Bąska, P.; Nowakowska, J.; Sokołowska, J.; Nowak-Życzyńska, Z.; Donten, M.; Celichowski, G.; Grobelny, J.; et al. Tannic Acid Modified Silver Nanoparticles Show Antiviral Activity in Herpes Simplex Virus Type 2 Infection. PLoS ONE 2014, 9, e104113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Orłowski, P.; Kowalczyk, A.; Tomaszewska, E.; Soliwoda, K.; Węgrzyn, A.; Grzesiak, J.; Celichowski, G.; Grobelny, J.; Eriksson, K.; Krzyzowska, M. Antiviral Activity of Tannic Acid Modified Silver Nanoparticles: Potential to Activate Immune Response in Herpes Genitalis. Viruses 2018, 10, 524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Aziz, A.T.; Alshehri, M.A.; Alanazi, N.A.; Panneerselvam, C.; Trivedi, S.; Maggi, F.; Sut, S.; Dall'Acqua, S. Phytochemical analysis of Rhazya stricta extract and its use in fabrication of silver nanoparticles effective against mosquito vectors and microbial pathogens. Sci. Total Environ. 2019, 700, 134443. [Google Scholar] [CrossRef]
  195. Kumar, S.D.; Singaravelu, G.; Ajithkumar, S.; Murugan, K.; Nicoletti, M.; Benelli, G. Mangrove-Mediated Green Synthesis of Silver Nanoparticles with High HIV-1 Reverse Transcriptase Inhibitory Potential. J. Clust. Sci. 2016, 28, 359–367. [Google Scholar] [CrossRef]
  196. Lin, Z.; Li, Y.; Guo, M.; Xu, T.; Wang, C.; Zhao, M.; Wang, H.; Chen, T.; Zhu, B. The inhibition of H1N1 influenza virus-induced apoptosis by silver nanoparticles functionalized with zanamivir. RSC Adv. 2017, 7, 742–750. [Google Scholar] [CrossRef] [Green Version]
  197. Li, Y.; Lin, Z.; Zhao, M.; Xu, T.; Wang, C.; Yinghua, L.; Wang, H.; Xia, H.M.; Zhu, B. Silver Nanoparticle Based Codelivery of Oseltamivir to Inhibit the Activity of the H1N1 Influenza Virus through ROS-Mediated Signaling Pathways. ACS Appl. Mater. Interfaces 2016, 8, 24385–24393. [Google Scholar] [CrossRef] [PubMed]
  198. Yang, X.X.; Li, C.M.; Huang, C.Z. Curcumin modified silver nanoparticles for highly efficient inhibition of respiratory syncytial virus infection. Nanoscale 2016, 8, 3040–3048. [Google Scholar] [CrossRef] [PubMed]
  199. Haggag, E.G.; Elshamy, A.M.; Rabeh, M.A.; Gabr, N.M.; Salem, M.; Youssif, K.A.; Samir, A.; Bin Muhsinah, A.; Alsayari, A.; Abdelmohsen, U.R. Antiviral potential of green synthesized silver nanoparticles of Lampranthus coccineus and Malephora lutea. Int. J. Nanomed. 2019, 14, 6217–6229. [Google Scholar] [CrossRef] [Green Version]
  200. Elumalai, D.; Hemavathi, M.; Deepaa, C.V.; Kaleena, P.K. Evaluation of phytosynthesised silver nanoparticles from leaf extracts of Leucas aspera and Hyptis suaveolens and their larvicidal activity against malaria, dengue and filariasis vectors. Parasite Epidemiol. Control 2017, 2, 15–26. [Google Scholar] [CrossRef] [PubMed]
  201. Elechiguerra, J.L.; Burt, J.L.; Morones-Ramirez, J.R.; Camacho-Bragado, A.; Gao, X.; Lara, H.H.; Yacaman, M.J. Interaction of silver nanoparticles with HIV-1. J. Nanobiotechnology 2005, 3, 6. [Google Scholar] [CrossRef] [Green Version]
  202. Sanchez-Guzman, D.; Le Guen, P.; Villeret, B.; Sola, N.; Le Borgne, R.; Guyard, A.; Kemmel, A.; Crestani, B.; Sallenave, J.-M.; Garcia-Verdugo, I. Silver nanoparticle-adjuvanted vaccine protects against lethal influenza infection through inducing BALT and IgA-mediated mucosal immunity. Biomaterials 2019, 217, 119308. [Google Scholar] [CrossRef] [PubMed]
  203. El-Mohamady, R.S.; Ghattas, T.; Zawrah, M.; El-Hafeiz, Y.A. Inhibitory effect of silver nanoparticles on bovine herpesvirus-1. Int. J. Vet. Sci. Med. 2018, 6, 296–300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Etemadzade, M.; Ghamarypour, A.; Zabihollahi, R.; Shabbak, G.; Shirazi, M.; Sahebjamee, H.; Vaziri, A.Z.; Assadi, A.; Ardestani, M.S.; Shandiz, S.A.S.; et al. Synthesis and evaluation of antiviral activities of novel sonochemical silver nanorods against HIV and HSV viruses. Asian Pac. J. Trop. Dis. 2016, 6, 854–858. [Google Scholar] [CrossRef]
  205. Tremiliosi, G.C.; Simoes, L.G.P.; Minozzi, D.T.; Santos, R.I.; Vilela, D.C.B.; Durigon, E.L.; Machado, R.R.G.; Medina, D.S.; Ribeiro, L.K.; Rosa, I.L.V.; et al. Ag nanoparticles-based antimicrobial polycotton fabrics to prevent the transmission and spread of SARS-CoV-2. bioRxiv 2020. [Google Scholar] [CrossRef]
  206. Mori, Y.; Ono, T.; Miyahira, Y.; Nguyen, V.Q.; Matsui, T.; Ishihara, M. Antiviral activity of silver nanoparticle/chitosan composites against H1N1 influenza A virus. Nanoscale Res. Lett. 2013, 8, 93. [Google Scholar] [CrossRef] [Green Version]
  207. Sujitha, V.; Murugan, K.; Paulpandi, M.; Panneerselvam, C.; Suresh, U.; Roni, M.; Nicoletti, M.; Higuchi, A.; Madhiyazhagan, P.; Subramaniam, J.; et al. Green-synthesized silver nanoparticles as a novel control tool against dengue virus (DEN-2) and its primary vector Aedes aegypti. Parasitol. Res. 2015, 114, 3315–3325. [Google Scholar] [CrossRef]
  208. Sharma, V.; Kaushik, S.; Pandit, P.; Dhull, D.; Yadav, J.P.; Kaushik, S. Green synthesis of silver nanoparticles from medicinal plants and evaluation of their antiviral potential against chikungunya virus. Appl. Microbiol. Biotechnol. 2018, 103, 881–891. [Google Scholar] [CrossRef] [PubMed]
  209. Li, W.; Hulswit, R.; Kenney, S.P.; Widjaja, I.; Jung, K.; Alhamo, M.A.; van Dieren, B.; van Kuppeveld, F.J.M.; Saif, L.J.; Bosch, B.-J. Broad receptor engagement of an emerging global coronavirus may potentiate its diverse cross-species transmissibility. Proc. Natl. Acad. Sci. USA 2018, 115, E5135–E5143. [Google Scholar] [CrossRef] [Green Version]
  210. Imai, Y.; Kuba, K.; Rao, S.; Huan, Y.; Guo, F.; Guan, B.; Yang, P.; Sarao, R.; Wada, T.; Leong-Poi, H.; et al. Angiotensin-converting enzyme 2 protects from severe acute lung failure. Nat. Cell Biol. 2005, 436, 112–116. [Google Scholar] [CrossRef] [PubMed]
  211. Jeremiah, S.S.; Miyakawa, K.; Morita, T.; Yamaoka, Y.; Ryo, A. Potent antiviral effect of silver nanoparticles on SARS-CoV-2. Biochem. Biophys. Res. Commun. 2020, 533, 195–200. [Google Scholar] [CrossRef] [PubMed]
  212. Alafeef, M.; Moitra, P.; Pan, D. Nano-enabled sensing approaches for pathogenic bacterial detection. Biosens. Bioelectron. 2020, 165, 112276. [Google Scholar] [CrossRef]
  213. Li, Y.; Lin, Z.; Zhao, M.; Guo, M.; Xu, T.; Wang, C.; Xia, H.; Zhu, B. Reversal of H1N1 influenza virus-induced apoptosis by silver nanoparticles functionalized with amantadine. RSC Adv. 2016, 6, 89679–89686. [Google Scholar] [CrossRef]
  214. Lee, L.A.; Wang, Q. Adaptations of nanoscale viruses and other protein cages for medical applications. Nanomed. Nanotechnol. Biol. Med. 2006, 2, 137–149. [Google Scholar] [CrossRef]
  215. Takeuchi, M.T.; Kojima, M.; Luetzow, M. State of the art on the initiatives and activities relevant to risk assessment and risk management of nanotechnologies in the food and agriculture sectors. Food Res. Int. 2014, 64, 976–981. [Google Scholar] [CrossRef]
  216. Kah, M.; Tufenkji, N.; White, J.C. Nano-enabled strategies to enhance crop nutrition and protection. Nat. Nanotechnol. 2019, 14, 532–540. [Google Scholar] [CrossRef]
  217. Ghormade, V.; Deshpande, M.V.; Paknikar, K. Perspectives for nano-biotechnology enabled protection and nutrition of plants. Biotechnol. Adv. 2011, 29, 792–803. [Google Scholar] [CrossRef] [PubMed]
  218. Mishra, S.; Singh, H.B. Biosynthesized silver nanoparticles as a nanoweapon against phytopathogens: Exploring their scope and potential in agriculture. Appl. Microbiol. Biotechnol. 2014, 99, 1097–1107. [Google Scholar] [CrossRef] [PubMed]
  219. Cox, A.; Venkatachalam, P.; Sahi, S.; Sharma, N. Reprint of: Silver and titanium dioxide nanoparticle toxicity in plants: A review of current research. Plant Physiol. Biochem. 2017, 110, 33–49. [Google Scholar] [CrossRef]
  220. Marchiol, L.; Mattiello, A.; Pošćić, F.; Giordano, C.; Musetti, R. In Vivo synthesis of nanomaterials in plants: Location of silver nanoparticles and plant metabolism. Nanoscale Res. Lett. 2014, 9, 1–11. [Google Scholar] [CrossRef] [Green Version]
  221. Usman, M.; Farooq, M.; Wakeel, A.; Nawaz, A.; Alam Cheema, S.; Rehman, H.U.; Ashraf, I.; Sanaullah, M. Nanotechnology in agriculture: Current status, challenges and future opportunities. Sci. Total Environ. 2020, 721, 137778. [Google Scholar] [CrossRef]
  222. Agrawal, S.; Rathore, P. Review Article Nanotechnology Pros and Cons to Agriculture: A Review. Int. J. Curr. Microbiol. Appl. Sci. 2014, 3, 43–55. [Google Scholar] [CrossRef]
  223. Nair, R.; Varghese, S.H.; Nair, B.G.; Maekawa, T.; Yoshida, Y.; Kumar, D.S. Nanoparticulate material delivery to plants. Plant Sci. 2010, 179, 154–163. [Google Scholar] [CrossRef]
  224. Abd-Elsalam, K.A.; Prasad, R. Nanobiotechnology Applications in Plant Protection; Springer: Berlin, Germany, 2018; ISBN 978-3-319-91160-1. [Google Scholar]
  225. Park, H.-J.; Kim, S.-H.; Kim, H.-J.; Choi, S.-H. A New Composition of Nanosized Silica-Silver for Control of Various Plant Diseases. Plant Pathol. J. 2006, 22, 295–302. [Google Scholar] [CrossRef]
  226. Gajbhiye, M.; Kesharwani, J.; Ingle, A.; Gade, A.; Rai, M. Fungus-mediated synthesis of silver nanoparticles and their activity against pathogenic fungi in combination with fluconazole. Nanomed. Nanotechnol. Biol. Med. 2009, 5, 382–386. [Google Scholar] [CrossRef]
  227. Mahakham, W.; Sarmah, A.K.; Maensiri, S.; Theerakulpisut, P. Nanopriming technology for enhancing germination and starch metabolism of aged rice seeds using phytosynthesized silver nanoparticles. Sci. Rep. 2017, 7, 1–21. [Google Scholar] [CrossRef] [PubMed]
  228. Sable, S.V.; Ranade, S.; Joshi, S. Role of AgNPs in the enhancement of seed germination and its effect on plumule and radicle length of Pennisetum glaucum. IET Nanobiotechnology 2018, 12, 922–926. [Google Scholar] [CrossRef] [PubMed]
  229. Ali, K.A.; Yao, R.; Wu, W.; Masum, M.I.; Luo, J.; Wang, Y.; Zhang, Y.; An, Q.; Sun, G.; Li, B. Biosynthesis of silver nanoparticle from pomelo (Citrus Maxima) and their antibacterial activity against acidovorax oryzae RS-2. Mater. Res. Express 2020, 7, 015097. [Google Scholar] [CrossRef]
  230. Manikandaselvi, S.; Sathya, V.; Vadivel, V.; Sampath, N.; Brindha, P. Evaluation of bio control potential of AgNPs synthesized from Trichoderma viride. Adv. Nat. Sci. Nanosci. Nanotechnol. 2020, 11, 035004. [Google Scholar] [CrossRef]
  231. Acharya, A.; Pal, P.K. Agriculture nanotechnology: Translating research outcome to field applications by influencing environmental sustainability. NanoImpact 2020, 19, 100232. [Google Scholar] [CrossRef]
  232. Galbraith, D.W. Nanobiotechnology: Silica breaks through in plants. Nat. Nanotechnol. 2007, 2, 272–273. [Google Scholar] [CrossRef]
  233. Vijayakumar, P.S.; Abhilash, O.U.; Khan, B.M.; Prasad, B.L.V. Nanogold-Loaded Sharp-Edged Carbon Bullets as Plant-Gene Carriers. Adv. Funct. Mater. 2010, 20, 2416–2423. [Google Scholar] [CrossRef]
  234. Shankar, S.; Chorachoo, J.; Jaiswal, L.; Voravuthikunchai, S.P. Effect of reducing agent concentrations and temperature on characteristics and antimicrobial activity of silver nanoparticles. Mater. Lett. 2014, 137, 160–163. [Google Scholar] [CrossRef]
  235. Jaiswal, L.; Shankar, S.; Rhim, J.-W. Applications of nanotechnology in food microbiology. Methods Microbiol. 2019, 46, 43–60. [Google Scholar] [CrossRef]
  236. Fayaz, A.M.; Balaji, K.; Girilal, M.; Kalaichelvan, P.T.; Venkatesan, R. Mycobased Synthesis of Silver Nanoparticles and Their Incorporation into Sodium Alginate Films for Vegetable and Fruit Preservation. J. Agric. Food Chem. 2009, 57, 6246–6252. [Google Scholar] [CrossRef]
  237. Moharekar, S.; Bora, P.; Kapre, V.; Uplane, M.; Daithankar, V.; Patil, B.; Moharekar, S.; Moharekar, S.T. Exploitation of Aspergillus niger for synthesis of silver nanoparticles and their use to improve shelf life of fruits and toxic dye degradation. IJIPSR 2014, 2, 2106–2118. [Google Scholar]
  238. García-Contreras, R.; Argueta-Figueroa, L.; Mejía-Rubalcava, C.; Jiménez-Martínez, R.; Cuevas-Guajardo, S.; Sánchez-Reyna, P.A.; Zeron, H.M. Perspectives for the use of silver nanoparticles in dental practice. Int. Dent. J. 2011, 61, 297–301. [Google Scholar] [CrossRef] [PubMed]
  239. Gurunathan, S.; Jeong, J.-K.; Han, J.W.; Zhang, X.-F.; Park, J.H.; Kim, J.-H. Multidimensional effects of biologically synthesized silver nanoparticles in Helicobacter pylori, Helicobacter felis, and human lung (L132) and lung carcinoma A549 cells. Nanoscale Res. Lett. 2015, 10, 1–17. [Google Scholar] [CrossRef] [Green Version]
  240. Patra, S.; Mukherjee, S.; Barui, A.K.; Ganguly, A.; Sreedhar, B.; Patra, C.R. Green synthesis, characterization of gold and silver nanoparticles and their potential application for cancer therapeutics. Mater. Sci. Eng. C 2015, 53, 298–309. [Google Scholar] [CrossRef]
  241. Choi, Y.-J.; Park, J.-H.; Han, J.W.; Kim, E.; Jae-Wook, O.; Lee, S.Y.; Kim, J.-H.; Gurunathan, S. Differential Cytotoxic Potential of Silver Nanoparticles in Human Ovarian Cancer Cells and Ovarian Cancer Stem Cells. Int. J. Mol. Sci. 2016, 17, 2077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  242. Ricciotti, E.; FitzGerald, G.A. Prostaglandins and inflammation. Arterioscler. Thromb. Vasc. Biol. 2011, 31, 986–1000. [Google Scholar] [CrossRef] [PubMed]
  243. Mani, A.K. Evaluation of In-vitro Anti-Inflammatory Activity of Silver Nanoparticles Synthesised using Piper Nigrum Extract. J. Nanomed. Nanotechnol. 2015, 6, 1. [Google Scholar] [CrossRef] [Green Version]
  244. Satyavani, K.; Gurudeeban, S.; Ramanathan, T.; Balasubramanian, T. Biomedical potential of silver nanoparticles synthesized from calli cells of Citrullus colocynthis (L.) Schrad. J. Nanobiotechnol. 2011, 9, 43. [Google Scholar] [CrossRef] [Green Version]
  245. Jacob, S.J.P.; Finub, J.; Narayanan, A. Synthesis of silver nanoparticles using Piper longum leaf extracts and its cytotoxic activity against Hep-2 cell line. Colloids Surf. B Biointerfaces 2012, 91, 212–214. [Google Scholar] [CrossRef] [PubMed]
  246. Sri Kumaran, N.; Vijayaraj, R. BIosynthesis of silver nano particles from Leucas aspera (willd.) link and its anti-inflammatory potential against carrageen induced paw edema in rats. Int. J. Pharm. Sci. Res. 2017, 8, 2588–2593. [Google Scholar] [CrossRef]
  247. Kim, J.S.; Kuk, E.; Yu, K.N.; Kim, J.-H.; Park, S.; Lee, H.J.; Kim, S.H.; Park, Y.K.; Park, Y.H.; Hwang, C.-Y.; et al. Antimicrobial effects of silver nanoparticles. Nanomed. Nanotechnol. Biol. Med. 2007, 3, 95–101. [Google Scholar] [CrossRef] [PubMed]
  248. Singh, R.; Thakur, P.; Thakur, A.; Kumar, H.; Chawla, P.; Rohit, J.V.; Kaushik, R.; Kumar, N. Colorimetric sensing approaches of surface-modified gold and silver nanoparticles for detection of residual pesticides: A review. Int. J. Environ. Anal. Chem. 2020, 1–17. [Google Scholar] [CrossRef]
  249. Zeng, S.; Baillargeat, D.; Ho, H.-P.; Yong, K.-T. Nanomaterials enhanced surface plasmon resonance for biological and chemical sensing applications. Chem. Soc. Rev. 2014, 43, 3426–3452. [Google Scholar] [CrossRef]
  250. Zhang, D.; Liu, Q. Biosensors and bioelectronics on smartphone for portable biochemical detection. Biosens. Bioelectron. 2016, 75, 273–284. [Google Scholar] [CrossRef] [PubMed]
  251. Xiong, D.; Li, H. Colorimetric detection of pesticides based on calixarene modified silver nanoparticles in water. Nanotechnology 2008, 19, 465502. [Google Scholar] [CrossRef]
  252. Menon, S.K.; Modi, N.R.; Pandya, A.; Lodha, A. Ultrasensitive and specific detection of dimethoate using a p-sulphonato-calix[4]resorcinarene functionalized silver nanoprobe in aqueous solution. RSC Adv. 2013, 3, 10623. [Google Scholar] [CrossRef]
  253. Pilaquinga, F.; Morey, J.; Torres, M.; Seqqat, R.; de las Nieves Piña, M. Silver nanoparticles as a potential treatment against SARS-CoV-2: A review. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnology 2021, e1707. [Google Scholar] [CrossRef]
  254. Khandelwal, N.; Kaur, G.; Kumar, N.; Tiwari, A. Application of silver nanoparticles in viral inhibition: A new hope for antivirals. Dig. J. Nanomater. Biostructures 2014, 9, 175–186. [Google Scholar]
  255. Song, K.S.; Sung, J.H.; Ji, J.H.; Lee, J.H.; Lee, J.S.; Ryu, H.R.; Lee, J.K.; Chung, Y.H.; Park, H.M.; Shin, B.S.; et al. Recovery from silver-nanoparticle-exposure-induced lung inflammation and lung function changes in Sprague Dawley rats. Nanotoxicology 2012, 7, 169–180. [Google Scholar] [CrossRef]
  256. Kim, J.S.; Sung, J.H.; Ji, J.H.; Song, K.S.; Lee, J.H.; Kang, C.S.; Yu, I.J. In vivo Genotoxicity of Silver Nanoparticles after 90-day Silver Nanoparticle Inhalation Exposure. Saf. Health Work 2011, 2, 34–38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Ji, J.H.; Jung, J.H.; Kim, S.S.; Yoon, J.-U.; Park, J.D.; Choi, B.S.; Chung, Y.H.; Kwon, I.H.; Jeong, J.; Han, B.S.; et al. Twenty-Eight-Day Inhalation Toxicity Study of Silver Nanoparticles in Sprague-Dawley Rats. Inhal. Toxicol. 2007, 19, 857–871. [Google Scholar] [CrossRef] [PubMed]
  258. Hyun, J.; Lee, B.; Ryu, H.; Sung, J.; Chung, K.; Yu, I. Effects of repeated silver nanoparticles exposure on the histological structure and mucins of nasal respiratory mucosa in rats. Toxicol. Lett. 2008, 182, 24–28. [Google Scholar] [CrossRef]
  259. Kwon, J.-T.; Minai-Tehrani, A.; Hwang, S.-K.; Kim, J.-E.; Shin, J.-Y.; Yu, K.-N.; Chang, S.-H.; Kim, D.-S.; Kwon, Y.-T.; Choi, I.-J.; et al. Acute Pulmonary Toxicity and Body Distribution of Inhaled Metallic Silver Nanoparticles. Toxicol. Res. 2012, 28, 25–31. [Google Scholar] [CrossRef] [Green Version]
  260. Sharma, S.; Mukkur, T.; Benson, H.A.; Chen, Y. Pharmaceutical Aspects of Intranasal Delivery of Vaccines Using Particulate Systems. J. Pharm. Sci. 2009, 98, 812–843. [Google Scholar] [CrossRef] [PubMed]
  261. Kumar, M.; Behera, A.K.; Lockey, R.F.; Zhang, J.; Bhullar, G.; De La Cruz, C.P.; Chen, L.-C.; Leong, K.; Huang, S.-K.; Mohapatra, S.S. Intranasal Gene Transfer by Chitosan–DNA Nanospheres Protects BALB/c Mice Against Acute Respiratory Syncytial Virus Infection. Hum. Gene Ther. 2002, 13, 1415–1425. [Google Scholar] [CrossRef] [PubMed]
  262. Rahman, M.; Wang, J.; Patterson, T.; Saini, U.; Robinson, B.; Newport, G.; Murdock, R.; Schlager, J.; Hussain, S.; Ali, S. Expression of genes related to oxidative stress in the mouse brain after exposure to silver-25 nanoparticles. Toxicol. Lett. 2009, 187, 15–21. [Google Scholar] [CrossRef] [PubMed]
  263. Wang, X.; Ji, Z.; Chang, C.H.; Zhang, H.; Wang, M.; Liao, Y.-P.; Lin, S.; Meng, H.; Li, R.; Sun, B.; et al. Use of Coated Silver Nanoparticles to Understand the Relationship of Particle Dissolution and Bioavailability to Cell and Lung Toxicological Potential. Small 2013, 10, 385–398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Tang, J.; Xiong, L.; Wang, S.; Wang, J.; Liu, L.; Li, J.; Yuan, F.; Xi, T. Distribution, Translocation and Accumulation of Silver Nanoparticles in Rats. J. Nanosci. Nanotechnol. 2009, 9, 4924–4932. [Google Scholar] [CrossRef] [PubMed]
  265. Tiwari, D.K.; Jin, T.; Behari, J. Dose-dependent in-vivo toxicity assessment of silver nanoparticle in Wistar rats. Toxicol. Mech. Methods 2010, 21, 13–24. [Google Scholar] [CrossRef]
  266. Shahare, B.; Yashpal, M. Toxic effects of repeated oral exposure of silver nanoparticles on small intestine mucosa of mice. Toxicol. Mech. Methods 2013, 23, 161–167. [Google Scholar] [CrossRef]
  267. Kim, Y.S.; Kim, J.S.; Cho, H.S.; Rha, D.S.; Kim, J.M.; Park, J.D.; Choi, B.S.; Lim, R.; Chang, H.K.; Chung, Y.H.; et al. Twenty-Eight-Day Oral Toxicity, Genotoxicity, and Gender-Related Tissue Distribution of Silver Nanoparticles in Sprague-Dawley Rats. Inhal. Toxicol. 2008, 20, 575–583. [Google Scholar] [CrossRef]
  268. Adeyemi, O.S.; Faniyan, T.O. Antioxidant status of rats administered silver nanoparticles orally. J. Taibah Univ. Med. Sci. 2014, 9, 182–186. [Google Scholar] [CrossRef] [Green Version]
  269. Gaillet, S.; Rouanet, J.-M. Silver nanoparticles: Their potential toxic effects after oral exposure and underlying mechanisms—A review. Food Chem. Toxicol. 2015, 77, 58–63. [Google Scholar] [CrossRef]
  270. Vazquez-Munoz, R.; Borrego, B.; Juarez-Moreno, K.O.; García-García, M.; Mota-Morales, J.; Bogdanchikova, N.; Huerta-Saquero, A. Toxicity of silver nanoparticles in biological systems: Does the complexity of biological systems matter? Toxicol. Lett. 2017, 276, 11–20. [Google Scholar] [CrossRef] [PubMed]
  271. Ji, J.; Yu, I.J. Estimation of human equivalent exposure from rat inhalation toxicity study of silver nanoparticles using multi-path particle dosimetry model. Toxicol. Res. 2012, 1, 206–210. [Google Scholar] [CrossRef]
  272. Lee, J.H.; Mun, J.; Park, J.D.; Yu, I.J. A health surveillance case study on workers who manufacture silver nanomaterials. Nanotoxicology 2011, 6, 667–669. [Google Scholar] [CrossRef]
  273. Armitage, S.A. The determination of silver in whole blood and its application to biological monitoring of occupationally exposed groups. Ann. Occup. Hyg. 1996, 40, 331–338. [Google Scholar] [CrossRef]
Figure 1. The mechanism of antibacterial action of AgNPs illustrated as (1) disruption of cell wall and membrane by silver ions released from AgNPs; (2) AgNPs inhibiting ATP synthesis; (3) mitochondrial dysfunction caused by Ag+ released from AgNPs; (4) ribosomal degradation by AgNPs; (5) AgNPs damaging nucleic acid by incorporating Ag+.
Figure 1. The mechanism of antibacterial action of AgNPs illustrated as (1) disruption of cell wall and membrane by silver ions released from AgNPs; (2) AgNPs inhibiting ATP synthesis; (3) mitochondrial dysfunction caused by Ag+ released from AgNPs; (4) ribosomal degradation by AgNPs; (5) AgNPs damaging nucleic acid by incorporating Ag+.
Nanomaterials 11 02086 g001
Figure 2. The possible mechanism of AgNPs’ antiviral efficacy.
Figure 2. The possible mechanism of AgNPs’ antiviral efficacy.
Nanomaterials 11 02086 g002
Table 1. Bacterial strains used for the biosynthesis of AgNPs, their characterization and activities.
Table 1. Bacterial strains used for the biosynthesis of AgNPs, their characterization and activities.
Sr.
No.
Organism Taken
for AgNPs
SizeShapeCharacterizationSynthesis ConditionsActivity StudiedReference
1.Escherichia coli,
Exiguobacterium aurantiacumm,
Brevundimonas diminuta
5–50SphericalUV, TEM,
XRD, FTIR,
SEM
At pH 9, incubation temp. 37 °C, for 24 h and 72 h in darkAntibacterial
Gram-positive
Bacillus subtilis,
Staphylococcus aureus,
Bacillus cereus
Gram-negative
Staphylococcus,
Pseudomonas aeruginosa,
Klebsiella pneumonia,
Escherichia coli,
Salmonella typhi,
Enterobacter vermicularris
[18]
2.Cyanobacterium,
Chroococcus
minutus
UV, SEM,
FTIR, XRD,
SEM-EDX
Incubated for 24–56 h at 40 °CAntibacterial
Escherichia coli,
Staphylococcus aureus,
Pseudomonas aeruginosa
[21]
3.Lysinibacillus
xylanilyticus
strain
MAHAQ-40
8–30SphericalUV, FTIR,
XRD, DSL,
FE-TEM
Incubated for 48 h at 30 °CAntibacterial
Salmonella typhimulium,
Vibrio parahaemolyticus
[19]
4.Sphingobium sp.
MAH-11
7–22SphericalSAED, XRD, FTIRIncubated for 24 h at 30 °CAntibacterial
Staphylococcus aureus,
Escherichia coli,
Pseudomonas aeruginosa
[22]
5.Bacillus pumilus,
Bacillus
paralicheniformis,
Sphigomonaspaucimobilis
4–20Spherical,
Oval
XRD, TEM, FTIR Cytotoxicity
Vigna radiata
[23]
6.Streptomyces
coelicoflavus
KS-3
2.33–91.3Spherical, truncated, triangular, quadrangular and hexagonalXRD, TEM,
FTIR, DLS,
EDX
At pH 7, incubated for 72 h at 32 °C in darkCytotoxicity
Carcinoma cells (HTB-182),
Adeno-carcinoma cells (A549)
[20]
7.Bacillus subtilis
(SDUM301120)
2–26SphericalHRTEM,
UV, XRD
EDX, FTIR
At pH 9, incubated for 12 hAntibacterial
Escherichia coli ATCC 25922,
Staphylococcus aureus ATCC29213,
Vibrio parahaemolyticus ATCC 17802,
Acinetobacter baumanni ATCC 19606
[24]
8.Cytobacillus
fimus
14.23SphericalUV, FTIR,
XRD, SEM,
DLS
Incubated for 24 h at 37 °CAntibacterial
Escherichia coli,
Staphylococcus aureus
Antifungal
Magnaporthe grisea
[25]
9.Lysinibacillus
sphaerius
14–21Spherical,
hexagonal,
cuboidal,
rod-shaped,
irregular
UV, TEM,
DLS, FTIR
Incubated at room temp. for 48 hAntibacterial
Gram-negative
E. coli ATCC 25955,
Pseudo aeruginosa ATCC 10145,
Klebsiella pneumonia,
Proteus vulgaris,
Salmonella typhimurium,
Enterobacter aerogennes,
Shigella dysenteriae
Gram-positive
Bacillus subtilis ATCC 6633,
Staphylococcus aureus NRRL
Antifungal
Yeast, filamentous fungi
Virucidal
Rotavirus
Cytotoxicity
Epithelial cell MA 104
[26]
10.Bacillus brevis
(NCIM 2533)
41–68SphericalUV, FTIR,
TLC, SEM,
AFM
Incubate for 2 h at room temp.Antibacterial
Salmonella typhai,
Staphylococcus aureus
[27]
11.Bacillus pumilus,
B. persicus,
B. licheniformis
77–92Spherical, triangular, hexagonalUV–Vis, FTIR, TEM, EDX, DLS Human pathogenic bacteria
Bean yellow mosaic virus
[28]
Table 2. Algal species used for the biosynthesis of AgNPs, their characterization and activities.
Table 2. Algal species used for the biosynthesis of AgNPs, their characterization and activities.
Sr.
No.
Organism Taken
for AgNPs
Size
(nm)
ShapeCharacterizationSynthesis ConditionsActivity StudiedReferences
1.Pyropiayozoensis20–22SphericalFTIR, XRD,
SEM, TEM
Incubated at 35 °C in dark for 15–20 minAntibacterial
Pseudomonas aeruginosa,
Staphylococcus aureus
[39]
2.Polysiphonia25SphericalFTIR, SEM, TEM, EDXStirring at room temp. for 2 hAnticancer
MCG-7 cell line
[40]
3.Spirulina platensis30–50SphericalFTIR, UV,
SEM, TEM
At pH range 4.7–5.0, incubated for 10 min at 60 °C Antibacterial
Escherichia coli, MTCC-9721,
Proteus vulgaris, MTCC-7299,
Klebsiella pneumoniae, MTCC-9751,
Staphylococcus aureus, MTCC-9542,
S. epidermidis, MTCC-2639,
Bacillus cereus, MTCC-9017
[41]
4.Chlorella valgaris55.06–61.89SphericalFTIR, XRD, FESEM, UV,
DLS
Incubated at room temp. for 24 hPhotocatalytic dye degradation[42]
5.Spyridia fusiformis5–50Spherical, triangular, pseudo-spherical, rectangular FTIR, TEM,
XRD, HR-TEM
Incubated at room temp.Antibacterial
Escherichia coli (ATCC 10798),
Klebsiella pneumaniae (ATCC 31488),
Staphylococcus aureus (ATCC 10832D-5),
Pseudomonas aeruginosa
(ATCC 207)
[43]
6.Noctiluca valgaris4.5SphericalDSL, SEM,
EDS, UV,
HRTEM
Anticancer
MDA-MB-231
Antibacterial
Escherichia coli ATCC25922,
Staphylococcus aureus ATCC29213
[44]
7.Spirulina platensis29Spherical and dispersedUV, SEM, TEM, DSL,
XRD, FTIR
Under sunlight for 10–20 min at pH 7, incubated at room temp.Anti-biofilm
Pseudomonas aeruginosa PA14
[45]
Table 3. Fungal species used for the biosynthesis of AgNPs, their characterization and activities.
Table 3. Fungal species used for the biosynthesis of AgNPs, their characterization and activities.
Sr.
No.
Organism
Taken
for AgNPs
Size
(nm)
ShapeCharacterizationSynthesis ConditionsActivity StudiedReference
1.Trichoderma
spp.
RoundUV, TEM,
FTIR
pH 5–7
Incubated for 3–9 days at 32 °C
Antibacterial
Gram-positive
Staohylococcus aureus ATCC 6538,
Enterococcus faecalis ATCC 29212
Gram-negative
E. coli ATCC 8939,
Pseudomonas aeruginosa ATCC25853
[56]
2.Penicillium
aculeatum
Su1
4–55SphericalUV, TEM,
DLS, EDX,
XRD, FTIR
Incubated for 72 h in dark at 32 °CEnzyme activity
Nitrate reductase
Protein identification
[57]
3.Trichoderma
longibrachiatum
10SphericalUV, TEM,
FTIR, DLS
Incubated for 48 h in dark at 22–33 °CAntifungal
Fusarium verticillioides,
Fusarium moniliforme,
Penicillum brevicompactum,
Heminthosporium oryzae,
Pyricularia grisea
[50]
4.Fusarium
scirpi
2–20Quasi-
spherical
UV, XRD,
STEM, HRTEM,
EDX
Incubated for 72 h at 28 °CAntimicrobial
Uropathogenic
E. coli
[58]
5.Punica
gramatum
5–45SphericalUV, FTIR,
XPS, XRD,
TEM
Incubated for 2 days at 90 °CDPPH,
DNA cleavage,
Antibacterial
Gram-negative
L. pneumophila,
P. aeruginosa,
E. coli
Gram-positive
E. hirae,
B. cereus,
S. aureus
[55]
6.Pichia kudriavzevii HA-NY2
Saccharomyces
uvarum
HA-NY3
12.4–30Cubic,
spherical
UV, TEM,
FTIR, DLS
Incubated for 72 h in dark at 30 °CAntibacterial
Gram-positive
Bacillus subtilis ATCC 6633,
Staphylococcus aureus ATCC 29213
Gram-negative
Pseudomonas aeruginosa ATCC 27953,
Candida tropicalis ATCC 750,
Fusarium oxysporium NRC21
Anti-inflammatory
Paw edema
[54]
7.Aspergillus niger10.31SphericalFTIR, TEM,
UV
Incubated for 72 h in dark at 28 °CAnti-amoebic
Allovahlkampfia spelaea
[59]
8.Piriformospora
indica
6–15SphericalUV, SEM,
EDX, TEM,
FTIR, XRD
At pH 6, incubated for 72 h at 28 °CDPPH
Anticancer
Human breast
Adenocarcinoma (MCF-7),
Human cervical carcinoma (HeLa),
Human liver hepatocellular carcinoma
(HepG2),
Embryonic kidney cell
(HEK-2930)
Antiproliferative
MCF-7, HeLa, HepG2
[60]
9.Trichoderma5–50Spherical,
oval
SEM, EDS,
TEM, XRD,
FTIR
At pH 7, incubated for 1 h at 25 °CAntifungal
Sclerotinia sclerotiorum
[61]
10.Cinnamomum
zeylanium
76Spherical,
oval
SEM, EDS,
TEM, XRD,
FTIR, EDX
Incubated for 24 h at room temp.Antibacterial
Staphylococcus aureus,
E. coli,
Pseudomonas aeruginosa
[53]
Table 4. Plant species used for the biosynthesis of AgNPs, their characterization and activities.
Table 4. Plant species used for the biosynthesis of AgNPs, their characterization and activities.
Sr.
No.
PlantsPartSize (nm)ShapeCharacterizationReducing
Agent
Synthesis
Conditions
Activities
Studied
References
1Ruta graveolensLeaves40–45 nmSphericalFTIR,
SEM
Alcohol, phenol, primary amine, sec. amine, azide, carbamide, allene, ketenimine, alkane, alkene, aldehyde, ester, amine, halo compoundsIncubated at room temp. for 24 hAnticancer
Antibacterial
Escherichia coli,
Proteus mirabilis,
Shigella flexneri,
Staphylococcus aureus
Insecticidal
Culex pipiens
[66]
2Phoenix dactylifera,
Ferula
asafetida,
Acacia nilotica
Fruit67.8–155.7 nmSphericalFT-IR
FE-SEM, TEM, Zeta potential
Alcohol, phenol or glycoside, amide, aromatic nitrileIncubated at room temp. for 48 hAntibacterial
Escherichia coli,
Staphylococcus aureus,
Pseudomonas aeruginosa
Anticancer
[69]
3Cymbopogon citrates (lemon grass)Leaves50–80 nmSphericalUV–Vis, XRD, FTIR, AFM, SEM, TEMAmine, phenol, alkane, alkylIncubated at room temp. Antidiabetic[70]
4Aaronsohnia factorovskyiStem,
leaves,
flower
104–140 nmSphericalUV–Vis, FT-IR, FE-SEM, GC–MSCarboxylic acid, alkyne, thiocynante, aronatic compound, alkene, isothiocynateIncubated in sunlight for 30 minAntibacterial
Staphylococcus aureus,
Bacillus subtilis,
Pseudomonas aeruginosa, Escherichia coli
Antifungal
Fusarium oxysporum,
Fusarium solani,
Helminthosporum rostratum, Alternaria alternate
[71]
5Eucalyptus camaldulensisLeaves16–28 nmSphericalUV–Vis, SEM, FTIR, XRD, EDX, DLS, ZetaAlcohol, phenol, flavonoid, flavones, catechin, ester, ether, alkane, carboxylic acid, primary amine, aldehyde Incubated at 25 °C in dark for 24 hAntioxidant[72]
6Terminalia arjunaLeaves10–50 nmSphericalUV–Vis, FTIR, TEM, FE-SEM, XRDHalo compounds, amine, alkyne, alcohol, phenolIncubated at 40–45 °CCatalytic degradation of organic dyes
methyl orange, methylene blue, Congo red and 4-Nitrophenol
[73]
7Ehretialaevis-Roxb.Leaves25–30 nmSphericalUV–Vis, FTIR, TEM, XRD, EDX, ZetaAlcohol or phenol, alkenes, primary amines, alkanes, alkyl halidesIncubated at 90 °C for 1 hAntimicrobial
Bacillus subtilis,
Escherichia coli,
Enterococcus faecalis,
Pseudomonas arginosa
Larvicidal
Culex quinquefasciatus
Anticancer
[74]
8Capparis zeylania L.Leaves23 nmSphericalUV–Vis, FTIR, XRD, SEM, TEMAlkynes, phosphine group, aliphatic ester, amine, nh group, carbonyl group, hydroxyl groupIncubated in dark conditions at 37 °CAntibacterial
Enterococcus faecalis,
Staphylococcus epidermidis,
Salmonella paratyphi,
Shigella dysenteriae
Antifungal
Candida albicans,
Aspergillus niger
Antiproliferative
[75]
9Tribulus terrestris L.Shoot~25 nmSphericalUV–Vis, TEM, DLS, XRD Kept in dark at room temp.Antibacterial
Bacillus subtilis,
Escherichia coli,
Pseudomonas aeruginosa,
Staphylococcus aureus
Photocatalytic activity
Cytotoxicity activity
[76]
10Cyperus pangoreiLeaves32–60 nmSphericalUV–Vis, FTIR, XRD, TEM, EDXXPSCarboxylic acid, phenol group, cyclohexane ringKept in an oven at 100 °C for 1 hPhotocatalytic activity of Ag NPs
against dye
Rhodamine B
[77]
11Onosmasericeum Wild.Root10–50 nmSphericalUV–Vis,
FT-IR,
FESEM, EDAX,
TEM,
XRD
Alcohol or phenolpH 8, temp. 25 °CAntibacterial
Acinetobacter baumannii,
Bacillus subtilis,
Escherichia coli,
Staphylococcus aureus,
Aeromonas hydrophila
Cytotoxicity
MCF-7 breast cancer cell line
Catalytic effect
2-nitrobenzenamine
[78]
12Abutilon indicumLeaves50–100 nmSphericalUV–Vis, FTIR, SEMAmines, alcohol, ketones, aldehyde, etc.Incubated at room temp.Antibacterial
Escherichia coli,
Streptococcus aurous
[79]
13DecaschistiacrotonifoliaLeaves12–18 nmSphericalUV–Vis, FTIR, TEM, XRDAlcohol, carboxylic acid, ester, ether, phenols, alkanes, amides, alkaloid, etc.Incubated at room temp.Antimicrobial
Escherichia coli,
Staphylococcus aureus, Klebsiella pneumonia,
Bacillus subtilis
Antioxidant
Photocatalytic activity of dyes
Cotton blue
Congo red
4-nitrophenol
[67]
14Azadirachta indica,
Citrullus colocynthis
Leaves
Fruit
17 nm
26 nm
SphericalUV–Vis, FTIR,
SEM,
EDX,
XRD
Polyphenols, aromatic terpenoid, flavonoids, alkene, ether, amines, aldehydes, ketones, carboxylic acid70 °C temp.Larvicidal
Aedes aegypti
[80]
15Litchi chinensisLeaves5–15 nmSphericalUV–Vis, FTIR, TEMPhenolic compound, alcohol, etc.95 °C temp. Bactericidal and sporicidal
Bacillus subtillus
[6]
16Asphodelus tenufoliusShoot,
Seed
58.2 nm
51.6 nm
Spherical and polydispersedUV–Vis, FTIR, SEMAlkanes, alkyne, alkene, carboxylic acid, alcohol/ phenolspH 5.5, temp. 30 °C [81]
17Jasmine flowerFlower40 nm UV–Vis, FTIR, SEM, TEMPhenolic compounds, alcohol, phenol, alkyl, etc.Incubated for 2 h at 110 °C Antimicrobial
Escherichia coli,
Staphylococcus aureus
Photocatalytic degradation of methylene blue dye
[82]
18Pisum sativum L.Seed10–25 nmSphericalUV–Vis, FTIR, XRD, SEMPhenol, alkynes, amines, alkyl halides, etc.24 h incubationAntioxidant
Antidiabetic
Cytotoxicity
Antibacterial
Escherichia coli,
Enterococcus faecium,
Streptococcus typhimurium and Streptococcus entrica
[83]
19Eryngium caucasicumTrautvLeaves10–20 nmSphericalUV–Vis, FTIR, XRD, TEM, SEMPhenols, amines, alcohol, carboxylic acid, ester, ether, terpenoid, flavonoids, tanninsIncubated at 80 °C for 8 hAntibacterial
Escherichia coli,
Staphylococcus aureus
[84]
20DregavolubilisFlower8.59–19.18 nmSphericalUV–Vis, FTIR, FESEM, EDX, HRTEMPolyphenol, phenolic acid, polysaccharides, flavones, amideIncubated at room temp. Antioxidant
Antidiabetic
Antibacterial
Bacillus subtilis,
Escherichia coli,
Pseudomonas aeruginosa,
Staphylococcus aureus
[85]
21Carissa carandas L.Leaves30–35 nm UV–Vis, FTIR, XRDAlkenes, methoxy group, alkynesIncubated at temp. 25 and 60 °CAntioxidant
Antibacterial
Salmonella typhimurium,
Enterobacter faecalis,
Shigella flexneri,
Citrobacter spp.
Gonococci spp.
[68]
22Artemisia vulgarisLeaves25 nmSphericalUV–Vis, FTIR, SEM, EDX, TEM, AFMPhenolic group, phenols, aromatic amines, carbonyl groupsIncubated at room temp. for 2 h Antioxidant
Anticancer
Antimicrobial
Escherichia coli,
Pseudomonas aeruginosa,
Staphylococcus aureus,
Klebsiella pneumoniae,
Haemophillus influenza
[86]
23AervalanataFlower90 nm UV–Vis, SEM, AFM, FTIR, TEM, XRDAlkenes, secondary amines, carboxylic acid ether, ester alcohol Anticancer
Antibacterial
Bacillus subtilis,
Escherichia coli,
Klebsiella planticola,
Streptococcus faecalis
Photocatalytic activity
Methylene blue
[87]
24Moringa oleiferaSeed4.0 nmSphericalUV–Vis, SEM, TEMFTIR, XRDPolyphenolic or flavonoid compounds, alkanes, alkenes, primary alcoholpH 11.0, temp. 60 °CAntimicrobial
Escherichia coli,
Pseudomonas aeruginosa,
Salmonella enterica typhimurium
Photocatalytic activity
Methylene blue, orange red,
4-nitrophenol
[88]
25Euphorbia sanguineLeaves20–28.8 nmSphericalUV–Vis, SEM, TEM, FTIRAmines, hydroxyl groupIncubated at room temp.Photocatalytic activity
Congo red dye
Melanogenesis inhibition activity
[89]
26Annona reticulataLeaves7.67–8.34 nmsphericalUV–Vis, FTRI, TEMXRDChloride group, anhydride group, methyl group, carbonyl group, alkane, amideIncubated for 2 h in dark at room temp.Larvicidal
Aedes aegypti
Antibacterial
Escherichia coli,
Pseudomonas aeruginosa,
Staphylococcus aureus,
Bacillus cereus
Antifungal
Candida albicans
[90]
27Allium ampeloprasumAerial part2.3–27 nmSphericalUV–Vis, FTIR, XRD, TEMAlcohol, phenolic compound, methyl, methylene, methoxy group, carboxylic acid, ester, ether, aliphatic amineTemp. 35–37 °CAnticancer
Antibacterial
Escherichia coli,
Pseudomonas aeruginosa,
Staphylococcus aureus
Antioxidant
[91]
28Chlorophytum borivilianumRoot tuber2.4–19.4 nmSphericalUV–Vis, FTIR, XRD, TEMHydroxyl group, aldehyde group, amide, aromatic etherpH 4.6, at room temp.Phytotoxicity
during seedling growth of Peltophorumpterocarpum
[92]
29Rosa Santana (Rose)Petal6.52–25.24 nmSphericalUV–Vis, FTIR, XRD, TEMHydroxyl, alkyl, alkyne, halogenated compoundIncubated at 90 °C for 25 minAntibacterial
Escherichia coli,
Staphylococcus aureus
Cytotoxicity
against mouse fibroblast cell line by XTT assay
[93]
30Citrus medica,
Tagetes lemmonii,
Tarenna asiatica
Leaves40–220 nm,
30–120 nm,
60–350 nm
Spherical, cuboidUV–Vis, FTIR, SEM, XRDAromatic amines, alcohols, carboxylic acids, esters, ether Larvicidal
Aedes aegypti
[94]
31Diospyros lotusLeaves10–25 nmSphericalUV–Vis, FTIR, TEM, XRD, SEMPhenolic, aromatic alkenes,
aliphatic hydrocarbon chain, amine
Phytochemical screening
Antibacterial
Escherichia coli
Anticoagulant
Catalytic activity
Methylene blue
[95]
32Chlorophytum borivilianum L. (Safed musli)CallusAVG. 52 nm UV–Vis, FTIR, XRD, AFMAmine, phenol,
hydroxyl, alkyl, alkenes, alkynes
Incubated at room temp. for 5 hAntibacterial
Escherichia coli,
Pseudomonas aeruginosa,
Staphylococcus aureus,
Bacillus subtilis
Antifungal
Candida albicans
Cytotoxicity
Colon cancer cell line
[96]
33Putranjivaroxburghi WallSeed13–69 nmSphericalUV–Vis, AEM, XRD, FTIR, TEMAliphatic primary amine, alkane, aldehydes, alkene, amines, sulfoxide, alcohol/phenols, phosphine, alkyl halide, alkyne, StaphylococcuspH 8.5Phytochemical analysis
Anticancer
Antibacterial
Escherichia coli,
Staphylococcus aureus,
Streptococcus pneumoniae, Enterococcus faecalis
[97]
34Cataharanthus roseusLeaves TEM Incubated in dark at room temp. for 24 hAnticancer [98]
35Solanum turvumFruits27 nmSphericalUV–Vis, FTIR, DLS, SEM, EDS, TEMAlcoholic and phenolic compound, alcohol and ether, polyphenols, amide, carbon chlorideIncubated at room temp. for 12 h, pH 6Antibacterial
Xanthomonas axonopodispv. punicae,
Ralstoniasolanacaerum
Phytotoxicity study
Vigna unguiculata
[99]
36Acacia niloticaStem27–50 nmSphericalUV–Vis, FTIR, XRD, SEM, TEM, XPS, DLS, etc.Phenol/carboxylic acid, secondary alcoholTemp. 40–50 °C for 5 hAntibacterial
Methicillin resistance,
Staphylococcus aureus
Antifungal
Candida albicans
Reduction of pollutant
4-nitrophenol,
2-nitrophenol
Degradation of dyes
Congo red, methylene blue, methyl orange
[100]
37Corn cobsXylanAvg. 55.3 nmSpherical, triangular, square, ovalUV–Vis, AFM, EDS, DLS, SEM, FTIR, RAMAN spectroscopy, flow cytometryMonosaccharide, carboxylic group, beta-glycosidic bond, hydroxyl group Incubated for 24 h in dark conditionsAnti-parasitic activity
Trypanosoma cruzi
Cytotoxicity
[101]
38MurrayakoenigiiLeavesAvg. 42 nmSphericalUV–Vis, FTIR, SEMHydroxyl group, ketones, aromatic compounds, quinoneIncubated at room temp. for 2 hAntibacterial
Escherichia coli,
Pseudomonas aeruginosa,
Staphylococcus aureus,Klebsiella pneumoniae
Cytotoxicity
In vitro
In vivo
[102]
39Nigella sativa
(Black cumin)
SeedAvg. 34 nmSphericalUV–Vis, XRD, FTIR, TEM, EDX, GC–MSHydroxyl, amide, alkenes, alcohol, aldehydes, ketones or carboxylic acid Antidiabetic
Anti-inflammatory
Antioxidant
Antibacterial
Escherichia coli,
Pseudomonas aeruginosa,
Staphylococcus aureus,
Listeria monocytogenes
[103]
40Catharanthus roseusLeaves40–60 nmBunch formFTIR, EDX, SEM Antibacterial
Shigella dysenteriae,
Klebsiella pneumoniae,
Bacillus anthraces,
Staphylococcus aureus,
Pseudomonas aeruginosa
[104]
41Allopylus cobsLeaves2–100 nmCrystallineUV–Vis, XRD, FTIR, TEM, XPS, DLSMethyl, amide, free amino, carboxylate Antibacterial
Pseudomonas aeruginosa,
Shigella flexneri,
Staphylococcus aureus,
Streptococcus pneumonia
Anti-biofilm
[105]
Table 5. Antiviral efficacy of AgNPs on different viruses, their characterization and activities.
Table 5. Antiviral efficacy of AgNPs on different viruses, their characterization and activities.
Sr.
No.
VirusFamilySource of AgNPsAgNPs
Size
CompositionSynthesis
Conditions
Mechanism of ActionReference
1.CoronavirusCoronaviridaePure Ag Nanowire and colloid10–20 nm3–13 µg/mLAgNPs procured from the Institute for Health and Consumer Protection (IHCP, a Joint Research Centre
of European Commission located in Italy)
AgNPs decreased cell apoptosis through activation of p38/mitochondria/caspase-3 signaling in ST cells [184]
2.Malaria, Nile Virus, ZikaFlaviviridaeNaregamiaalata5–35 nm6–30 µg/mLIncubated for 10 min at room temp.AgNPs’ passage through the insect cuticle and into individual cells interferes with molting and other physiological processes [185]
3.SARS-CoV-2Coronaviridae 200 nm
coating
100–200 µL Coating reduces the titers of SARS-CoV-2 to zero [186]
4. Herpes Simplex Virus
and Human Parainfluenza Virus Type 3
Herpesviridae
Paramyxoviridae
Alternaria species 46 nm0.1–10 µg/mLFungus suspended in distilled water for 48 h AgNPs control viral infectivity by
blocking interaction of the virus with the cell
[187]
5. Herpes Simplex Virus
Human Parainfluenza Virus Type 3
Herpesviridae
Paramyxoviridae
Fusarium. oxysporum 20 nm0.1–10 µg/mLFungus suspended in distilled water for 48 h AgNPs may block an early event before stable
binding of the virus with the cell membrane, but it is likely that the nanoparticles interact directly with the viral envelope or its proteins and behave as virucidal agents
[187]
6. Herpes Simplex Virus
and Human Parainfluenza Virus Type 3
Herpesviridae
Paramyxoviridae
Curvularia species 30 nm0.1–10 µg/mLFungus suspended in distilled water for 48 h Interference with replication at the post-entry phase [187]
7. Herpes Simplex Virus (HSV-I,II) HerpesviridaeSargassum withtii 0.5–5 µg/mLSeaweed powder treated with 90 mL of 10 mM AgNO3 solution for 15 h under stirring conditionsAgNPs possess size-dependent interaction and the ability to block virus attachment and entry [188]
8.Respiratory syncytial virusPneumoviridaePoly-vinylpyrolidone (PVP) coated silver nanospheres8–12 nm1 mg/mL and
2–4 mg/kg of mice
Procured from
NanoComposix Inc. (San Diego, CA, USA)
AgNPs attached to surface glycol proteins and interfered with RSV’s ability to initiate attachment with the proper receptors, preventing fusion of the virus to the host cell [189]
9.Norovirus Surrogates 1 g/Kg 4–10 nm21 mg/LPHBV18 suspended in ultrapure Milli-Q water and then mixed with sodium borohydride and AgNO3AgNPs reduce certain no. of NK cells [190]
10.Rhesus Rotavirus Reoviridae Collagen 10 nm50 µL (0.4 mM)The AgNP–collagen mixture gelled inside the abdominal cavity at temp. 37 °CThe virus load reduced in the liver due to increase in NK cells and T cells together [191]
11.Herpes Simplex Virus Type 2Herpesviridae Tannic-acid-modified AgNP 10–65 nm2.5–5 µg/mL Mixture of sodium citrate (4%) and tannic acid (5%) added to AgNO3 and stirredAgNPs interact with the virion’s surface and create a physical obstacle, impairing interaction with
the viral receptors on the cell surface
[192]
12.Herpes Simplex Virus 2 (HSV-2)Herpesviridae Tannic-acid-modified AgNP 20–40 nm5 µg/mLReducing agent added to
aqueous solution of silver nitrate and heated to boiling point
Tannic acid has been shown to inhibit the attachment
of viruses to host cells
[193]
13.Zika VirusFlaviviridae Rhazyastricta 20–40 nm5–120 µg/mLDried powdered leaf extract mixed with 1 mM AgNO3 at room temp.Due to minute size, they effectively penetrate the infectious agent [194]
14.HIV-1 Retroviridae Rhizophoralamarckii 12–28 nm0.25–1 µg/mL 3 mL of extract reacted with 47 mL of 1 mM AgNO3 and incubated for 6 hHIV-1 reverse transcriptase inhibitory activity [195]
15.H1N1 InfluenzaeOrthomyxoviridae Zanamivir AgNP 2–3 nm2.5 µg/mL Constant magnetic stirring for 30 min at room temp.Zanamivir is neuraminidase (NA) inhibitor and binds with NA pocket to disturb enzyme reaction [196]
16.H1N1 InfluenzaeOrthomyxoviridaeOseltamivir2–3 nm2.5 μg/mL0.1 mL Vit. C added to 4 mL AgNO3 at room temp.Blocks the release of new virions from the cell’
smembrane and becomes resistant to the influenza A virus
[197]
17.Respiratory Syncytial Virus (RSV) Paramyxoviridae Curcuma longa 0.23 nm 0.008–0.24 nM Curcumin dissolved in DMSO and added to ultra-pure water. Vigorous stirring of AgNO3 (10 mM) at 100 °CAgNPs could prevent the virus from entering into cells and its replication [198]
18.HSV-1, HAV-10, and
CoxB4 virus
Herpesviridae Lampranthus
coccineus
10.12–27.89 nm10–40 µg/mL Aqueous extract added to 1
mM silver nitrate in the ratio 2:10 and kept in water bath for 10 min at 60 °C
Interacts with herpes simplex thymidine kinase,
hepatitis A 3c proteinase and Coxsackie virus B4 3c protease
[199]
19.HSV-1, HAV-10 andCoxB4 virusHerpesviridae Malephora lutea F. Aizoaceae 8.91–14.48 nm10–40 µg/mL Aqueous extract added to 1
mM silver nitrate in the ratio 2:10 and kept in water bath for 10 min at 60 °C
Interacts with herpes simplex thymidine kinase,
hepatitis A 3c proteinase and Coxsackie virus B4 3c protease
[199]
20.Dengue VirusFlaviviridae Leucasaspera
Hyptissuaveolens
7–22 nm TEM
22–43 nm FRSEM
2–10 mg/L2 mL of fresh extract added to 98 mL of aqueous silver nitrate (1 mM) solution and incubated at 28 °C for 60 minThe surface reactivity facilitated by capping makes these functionalized NPs a promising tool for vector control [200]
21.HIV-1Retroviridae PVP/BSA-coated AgNPs 1–10 nm25 µg/mL AgNPs procured from Nanotechnologies, Inc.AgNPs bind with GP120 subunit of viral envelope glycoprotein [201]
22.InfluenzaOrthomyxoviridae Lipoic acid 8–12 nm0.5–5 µg/mL AgNPs procured from
Nanocomposix company
AgNPs induced neutrophil and monocyte recruitment and increased the levels of KC (CXCL-1), IL-12 and IL-6, as soon as
4 h after AgNP injection
[202]
23.Herpesvirus-1Herpesviridae Chemical reduction 15–50 nm24 µg/mL The solution prepared for two-fold serial dilutions with varied concentrationAgNPs at nontoxic concentrations were capable of inhibiting BoHV-1 when administered prior to viral infection [203]
24.HSV-1, HIVHerpesviridae
Retroviridae
Sonochemical method 1–10 nm1–10 µmol/mL Severe stirring at room temperatureInteract with viral envelope glycoprotein [204]
25.SARS-CoV-2CoronaviridaePolycotton AgNP-CS10–28 nm5%NanoxTecnologia S.A.–São Carlos/SP-Brazil)Binding of AgNPs with sulfur residues from the virus’s surface glycoproteins, preventing interaction with the receptor and its entry into the host cell [205]
26.H1N1 Influenza A VirusOrthomyxoviridaeAgNP–chitosan 5.5–12.9 nm62–77 µg/mL Aqueous medium at room temperatureVirion and composite interacted as the NP size is very small and cause degradation of virus [206]
27.Dengue virus (DEN-2)FlaviridaeMoringa oleifera100 nm20 µg/mL Toxic action of M. oleifera AgNPs against A. aegypti may be linked to lectin content, which is able to affect digestive and detoxifying enzymes[207]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bamal, D.; Singh, A.; Chaudhary, G.; Kumar, M.; Singh, M.; Rani, N.; Mundlia, P.; Sehrawat, A.R. Silver Nanoparticles Biosynthesis, Characterization, Antimicrobial Activities, Applications, Cytotoxicity and Safety Issues: An Updated Review. Nanomaterials 2021, 11, 2086. https://doi.org/10.3390/nano11082086

AMA Style

Bamal D, Singh A, Chaudhary G, Kumar M, Singh M, Rani N, Mundlia P, Sehrawat AR. Silver Nanoparticles Biosynthesis, Characterization, Antimicrobial Activities, Applications, Cytotoxicity and Safety Issues: An Updated Review. Nanomaterials. 2021; 11(8):2086. https://doi.org/10.3390/nano11082086

Chicago/Turabian Style

Bamal, Deepak, Anoop Singh, Gaurav Chaudhary, Monu Kumar, Manjeet Singh, Neelam Rani, Poonam Mundlia, and Anita R. Sehrawat. 2021. "Silver Nanoparticles Biosynthesis, Characterization, Antimicrobial Activities, Applications, Cytotoxicity and Safety Issues: An Updated Review" Nanomaterials 11, no. 8: 2086. https://doi.org/10.3390/nano11082086

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