Next Article in Journal
Antiprotozoal Activity of Thymoquinone (2-Isopropyl-5-methyl-1,4-benzoquinone) for the Treatment of Leishmania major-Induced Leishmaniasis: In Silico and In Vitro Studies
Next Article in Special Issue
pH Alteration in Plant-Mediated Green Synthesis and Its Resultant Impact on Antimicrobial Properties of Silver Nanoparticles (AgNPs)
Previous Article in Journal
Clostridioides difficile Infection Treatment and Outcome Disparities in a National Sample of United States Hospitals
Previous Article in Special Issue
Development of Wash-Durable Antimicrobial Cotton Fabrics by In Situ Green Synthesis of Silver Nanoparticles and Investigation of Their Antimicrobial Efficacy against Drug-Resistant Bacteria
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Silver Nanoparticle-Based Therapy: Can It Be Useful to Combat Multi-Drug Resistant Bacteria?

by
Eva M. Mateo
1,* and
Misericordia Jiménez
2
1
Department of Microbiology and Ecology, Faculty of Medicine and Odontology, Universitat de Valencia, E-46010 Valencia, Spain
2
Department of Microbiology and Ecology, Faculty of Biological Sciences, Universitat de Valencia, E-46100 Valencia, Spain
*
Author to whom correspondence should be addressed.
Antibiotics 2022, 11(9), 1205; https://doi.org/10.3390/antibiotics11091205
Submission received: 23 July 2022 / Revised: 1 September 2022 / Accepted: 5 September 2022 / Published: 6 September 2022
(This article belongs to the Special Issue Antimicrobial Silver in Medicinal Applications)

Abstract

:
The present review focuses on the potential use of silver nanoparticles in the therapy of diseases caused by antibiotic-resistant bacteria. Such bacteria are known as “superbugs”, and the most concerning species are Acinetobacter baumannii, Pseudomonas aeruginosa, Staphylococcus aureus (methicillin and vancomycin-resistant), and some Enterobacteriaceae. According to the World Health Organization (WHO), there is an urgent need for new treatments against these “superbugs”. One of the possible approaches in the treatment of these species is the use of antibacterial nanoparticles. After a short overview of nanoparticle usage, mechanisms of action, and methods of synthesis of nanoparticles, emphasis has been placed on the use of silver nanoparticles (AgNPs) to combat the most relevant emerging resistant bacteria. The toxicological aspects of the AgNPs, both in vitro using cell cultures and in vivo have been reviewed. It was found that toxic activity of AgNPs is dependent on dose, size, shape, and electrical charge. The mechanism of action of AgNPs involves interactions at various levels such as plasma membrane, DNA replication, inactivation of protein/enzymes necessary, and formation of reactive oxygen species (ROS) leading to cell death. Researchers do not always agree in their conclusions on the topic and more work is needed in this field before AgNPs can be effectively applied in clinical therapy to combat multi-drug resistant bacteria.

Graphical Abstract

1. Introduction

The World Health Organization (WHO) has published a list of the emerging bacteria or bacterial families that pose the greatest threat to human health because they are resistant to many antibiotics and for which there is a very urgent need for new treatments [1,2,3,4]. The list ranks Acinetobacter baumannii, Pseudomonas aeruginosa and Enterobacteriaceae (Klebsiella pneumoniae, Escherichia coli, Enterobacter spp., Serratia spp., Proteus spp., Providencia spp., and Morganella spp.) extended-spectrum-β-lactamase-producing (carbapenem-resistant), followed by Enterococcus faecium (vancomycin-resistant), Staphylococcus aureus (methicillin, vancomycin resistant), Helicobacter pylori (clarithromycin-resistant), Campylobacter spp. and Salmonellae (fluoroquinolone-resistant), Neisseria gonorrhoeae (cephalosporin, fluoroquinolone-resistant), Streptococcus pneumoniae (penicillin-non-susceptible), Haemophilus influenzae (ampicillin-resistant) and Shigella spp. (fluoroquinolone-resistant) as priority targets. Mycobacteria species were not included in this review to narrow down the scope of the present work and focus entirely on the main species listed by the WHO [1,2]. In the last years, the interest in nanotechnology has become increasingly important for global industries. Applications in medicine extend from the use of nanomaterials for medical devices to the use of nanoparticles (NPs) as therapeutic agents, drug delivery systems, or diagnostic imaging systems. NPs are engineered structures defined as particles with a diameter of 1–100 nm [5,6], though some of the reported NPs exhibit a size >100 nm [7]. According to Mitchell et al. [8], NPs can be classified as lipid-based, polymeric, and inorganic NPs; to which carbon-based NPs can be added [9]. Lipid-based NPs are usually spherical platforms comprising at least one lipid bilayer surrounding one or more internal aqueous compartments and are used as a delivery system. They include liposomes, lipid NPs, and oil/water emulsions. Their advantages include high bioavailability, formulation simplicity, or self-assembly, making them very useful in nanomedicine. Polymeric NPs include polymersomes, dendrimers, polymer micelles, and nanospheres. They are good delivery vehicles because of their biocompatibility and simple formulation parameters; they are hydrosoluble, stable, and well suited for the delivery of drugs encompassing different sizes, structures, and polarities [8]. Polymer NPs allow for the encapsulation of molecules that can be released at targeted sites [10,11,12]. Among the inorganic NPs, there are silica NPs (crystalline or amorphous), metal NPs (such as copper, titanium, nickel, selenium, gold, silver), metal oxides (titanium dioxide, iron oxide, zinc oxide, magnesium oxide, etc.) or quantum dots (typically made of semiconducting materials, such as Si). Inorganic NPs have been used to synthesize nanostructured materials for various drug delivery and imaging applications and can have a variety of sizes, structures, and geometries [8,13].
Metal NP nanostructure may be diverse (nanotube, nanorod, nanowire, nanocrystal, spherical, and dendritic aggregated nanomaterial, quantum dots, etc.) among others [8]. Metal NPs can be covered with organic compounds such as polymers to give nanocomposites. The characteristics of nanomaterials (type, shape, size, electrical charge, surface coating, concentration, etc.) are responsible for their effectiveness [14,15]. Characterization of metal NPs is essential to know their mechanism of action and toxicity. For this purpose, different techniques such as UV-visible spectroscopy, X-ray diffraction analysis, scanning electron microscopy (SEM), transmission electron microscopy (TEM), Fourier transform infrared spectroscopy (FTIR), atomic force microscopy, Zeta potential measurement, dynamic light scattering (DLS), or single-particle inductively coupled plasma mass-spectrometry analysis can be used [16,17].
NPs are applied, mainly as drug delivery systems, in different therapeutic areas, such as CNS diseases [18], cardiovascular diseases [19], ocular pathologies [20], Alzheimer’s disease [21], diabetes treatment [22], or immunotherapies [23]. However, oncology is the main area of NP applications [8,24,25], and the second area corresponds to infectious diseases [26,27,28]. An advantage of NP formulations vs conventional systems is their multivalency, such that the presence of various functional groups from a NP permits a higher cell recognition and a higher target binding ability than those of linear polymers [12].
The use of metal NPs to treat infections is particularly interesting against multi-drug resistant (MDR) pathogens [29,30]. Numerous studies have tried to elucidate the mechanisms by which NPs inhibit bacterial growth [29,30,31,32], but a clear and complete understanding has not yet been achieved. The nature of the interaction between different functional groups of bacterial surface and NP surface has been studied [13,33]. Gram-positive and Gram-negative bacterial cell walls have a net negative charge. In Gram-positive, the negative charge is provided by teichoic acids, which are linked to the peptidoglycan or to the underlying plasma membrane. Teichoic acids are anionic owing to the presence of phosphates within their structure. Gram-negative bacteria have in their wall an outer membrane with phospholipids and lipopolysaccharides. Lipopolysaccharides confer a strong negative charge on their surface. Positively charged metal NPs have higher bactericidal activity than negatively charged or neutral metal NPs [34,35]. Thus, many studies have focused on NPs with positive surface charge, particularly AgNPs, which are considered the next generation of antimicrobials for the treatment and prevention of MDR microbes [36,37,38,39]. In addition, they experience slow oxidation and release cations.
Silver nanoparticles (AgNPs) have antibacterial, antiviral, and antifungal properties [25]. The NP nanosize enables interaction with biomolecules on the cell wall and membrane and further infiltration into microbial cells. It is necessary to distinguish between (a) the intrinsic antibacterial properties of NPs, such as the ability to damage bacterial membranes, and cause further damage to molecules or structures (DNA, enzymes, ribosomes, lysosomes, etc.) by diverse mechanisms, or inhibit bacterial biofilm formation, and (b) their properties as drug delivery systems such as the capacity to alter cell wall permeability, permitting the entry of antibiotics, even those attached to the coating shell, inside the cell. In this way, the released drugs can act against the bacteria at the target site of action [40].
The mechanisms of action of NPs are not fully understood due to the multi-factorial nature of the activity, making it difficult to decouple the individual mechanisms [30,32]. The main effects of NPs and cations are interactions with the cell membrane of bacteria, which leads to disruption (depolarization of membrane potential), changes its permeability, and allows the intracellular content to be released [30,41]; NPs accumulate in the cell wall, forming “pits” and pores, leading to cell death [32,42]. Released ions inhibit the site between cytochrome α2 and b-cytochromes in the respiratory chain and the cellular respiration process is interrupted in the electron transport chain [30,32,34]. There is inhibition of ribosomal subunits expression, which prevents translation and protein synthesis, inactivation of some cellular proteins and enzymes necessary for adenosine triphosphate (ATP) biosynthesis [30,43] alteration of the normal function of membrane-bound respiratory enzymes, inhibition of thiol group-containing enzymes, such as nicotinamide adenine dinucleotide dehydrogenase II (NADH-dehydrogenase II) in the respiratory system by silver ions [44,45] and generation of reactive oxygen species (ROS), which include hydrogen peroxide (H2O2), singlet oxygen (O2), hydroxyl radical (˙OH), and superoxide radical (O2) with oxidative deterioration to cell content [31]. In high levels, these species can damage the DNA [39], and cell membrane, lead to lipid peroxidation and protein oxidation, initiate lethal stress response cascades, and eventually cell death. [13,30,32,39,46] (Figure 1).
There are different approaches regarding the methods of synthesis of metal NPs, overall, there are top-bottom and bottom-up methodologies [25,47]. The former methods start from bulk quantities of materials that are reduced in size and are mixed with clusters of atoms or ions. Some top–bottom methods use physical technologies, such as thermal/laser ablation, mechanical milling, or sputtering. In the bottom-up approach, nanostructures are built atom by atom or particle by particle. This can be attained by a high degree of super saturation followed by nuclei growth [25]. Within the bottom-up approach, there are physical, chemical/electrochemical, and biological methods. Physical methods include condensation, vapor deposition, sol/gel processes, or pyrolysis [47]. In the case of AgNPs, some authors report on chemical synthesis using chemicals, such as NaBH4, sodium citrate, chitosan, polyvinylpyrrolidone (PVP), sodium dodecyl sulfate (SDS), etc. that act as reductants and/or capping agents, which are mixed with aqueous solutions of Ag+ from soluble silver salts (AgNO3) [48,49] under variable conditions of pH/temperature/reaction time to produce NPs. However, most researchers opt for biological synthesis from a variety of natural sources, such as bacteria, fungi, algae, parts of terrestrial plants (leaves, fruits, rhizomes, whole plant, …), sugars, etc. to reduce Ag+ and form AgNPs. These biosynthetic methods are considered green or environmentally friendly [25,47,49,50] because hazardous chemicals are not used or residues from the synthetic processes are not released into nature. Moreover, they are not expensive. In particular, the utilization of plant extracts for manufacturing NPs is affordable, readily scaled up, and environment-friendly. Plant extracts have the ability to generate NPs with a specified size, shape, and content. Plant produced NPs have the potential to be extensively employed in current medical processes [47].
The possible potential of the NPs, especially AgNPs, in the control of MDR pathogens has been studied [13,26,32,36,37,40]. However, an accurate assessment of the NP potential in the treatment of infections caused by MDR bacteria, and a critical comparative analysis between reports is not possible, as the methodologies used by researchers are different. No in-depth interlaboratory study using the same methodology, NPs, pathogen, and experimental conditions has been performed, which hinders the correct evaluation of the NP potential to combat these infections. Moreover, despite the advantages and potential applicability of NPs in combating MDR pathogens, their possible toxicity and safety issues have limited their general, efficient, and safe use [51].
Based on the above-cited characteristics of NPs, this review focuses on the current status and future prospects of AgNPs as a possible tool to treat infections caused by some MDR bacteria or bacterial families considered by the WHO [1,2] as priority targets to combat. AgNPs are described as the most efficient antibacterial agents based on their action against the pathogen and their toxicity to the host. The use of nanomaterials as potential antimicrobial agents might be considered a post-antibiotic era, which has the ability to overcome the problem of multi-drug resistance [29]; moreover, to avoid the issue of bacterial resistance to NPs, an understanding of the adaptive mechanisms of microbes to resist the action of NPs should be an objective in future studies [29].

2. Survey Methodology

The methodology used to survey the bibliography on the topic reviewed was:
(a) A systematic search on the internet using bibliographic databases (Google Scholar, MedLine, PubMed, Web of Science, Scopus, Springer-link, ScienceDirect, etc.) with the keywords “Nanoparticles”, “nanomaterials”, “Silver nanoparticles” combined with the keywords “multi-drug resistant bacteria”, “Acinetobacter baumannii”, “Pseudomonas aeruginosa”, “Staphylococcus aureus”, “Enterobacteriaceae”, “E. coli”, “Klebsiella”, “Salmonella”, “cytotoxicity”, “toxicity”.
(b) A subsequent search was made using the reference lists on the found articles or reviews. The references were selected on the basis of their relevance to the target microorganisms.
(c) The search had a first introductory part dedicated to performing a general introduction of the usage of nanoparticles in clinical applications, the techniques of characterization of metal NPs, the mechanisms of action, and the methods of synthesis.
(d) The focus was on the utilization of AgNPs to combat the four kinds of microorganisms cited and the toxicological issues (in vitro and in vivo) related to these NPs.

3. Acinetobacter baumannii

Acinetobacter baumannii is a Gram-negative, aerobic, non-motile bacterium that causes nosocomial infections, most notably ventilator-associated pneumonia, and bacteremia, and less frequently meningitis, skin and soft tissue infections, urinary tract infections, and endocarditis. A. baumannii pneumonia and bacteremia are typically acquired in the hospital and mainly affect critically ill patients. These are severe infections for which almost no treatments exist and are associated with high mortality [52,53]. Particularly concerning are pan-drug resistant strains of A. baumannii, with resistance to all clinically used antibiotics [4,54,55,56]. Consequently, novel strategies for managing these infections are required. The mechanisms of resistance, virulence, and pathogenicity of MDR A. baumannii have been recently reviewed [53].
The use of AgNPs has been explored as one of such strategies. Table 1 lists some of the research activities achieved on the use of AgNPs against A. baumannii. Fourteen antibiotics (amikacin, gentamicin, kanamycin, amoxicillin, ampicillin, ceftriaxone, vancomycin, ciprofloxacin, doxycycline, tetracycline, chloramphenicol, trimethoprim, ceftazidime, and penicillin) were mixed with AgNPs (8–12 nm) and assayed against A. baumannii in vitro. The bacterium resulted in highly sensitized through a synergistic effect becoming susceptible to antibiotics except for cephalosporins [57]. Biosynthesized AgNPs demonstrated good activity against A. baumannii both in vitro and in vivo by impregnation of cotton fabric and further application to wounds of Sprague–Dawley infected female rats [58]. Biosynthesized spherical AgNPs (27 nm) inhibited the growth of A. baumannii more than vancomycin used as a control [59]. AgNPs capped with different compounds have been studied regarding their activity against A. baumannii strains. While chitosan and SDS-capped AgNPs were ineffective, citrate and PVP-capped AgNPs showed a good inhibitory effect on such strains. PVP capped AgNPs (PVP-AgNPs) proved very effective against carbapenem-resistant strains of A. baumannii [60], being suggested as an alternative to carbapenem, and were successfully assayed in human pulmonary host cells (A-549) at doses that were not toxic to these cells. Thus, they might be used alone or along with carbapenem to cure infections caused by carbapenem-resistant strains of A. baumannii [61,62]. A nanocomposite made of AgNPs coated with SH-PEG-NOTA (thiol-containing polyethylene glycol linked to 1,4,7-triazacyclononane-1,4,7-triacetic acid), and imipenem, noted as IPM@AgNPs-PEG-NOTA, was assayed and shown to be a promising antibacterial agent of security, pH sensitivity, and high efficiency in reversing resistance and synergistically combatting carbapenem-resistant A. baumannii [63]. Single AgNPs completely inhibited carbapenem-resistant A. baumannii growth at 2.5 μg/mL and AgNP treatment showed synergistic effects with polymyxin B and rifampicin, and an additive effect with tigecycline [64]. Minimal inhibitory concentrations (MIC) of AgNPs against various A. baumannii strains were found in the 0.39–0.78 µg/mL range, and the more resistant strains were generally less susceptible to antibiotics [65]. AgNPs play an important role in the prevention of burn infections frequently caused by MDR A. baumannii because they are more active than AgNO3 or sulfadiazine and are less chemically contaminated than ion forms [65]. A xanthan gum polymer containing spherical AgNPs (diameter < 10 nm) was tested in vitro on MDR strains of A. baumannii which proved sensitive to the Ag nanocomposite [66]. AgNPs show synergism with imipenem and other antibiotics against planktonic cells and biofilms of A. baumannii [56,64,65,66,67]. Biofilms are surface-associated bacterial communities that are found embedded in a self-produced exopolysaccharide matrix that attaches to surfaces or living tissues [68]. Trimethyl chitosan-capped AgNPs with positive surface charge inhibited MDR A. baumannii strains and other pathogenic bacteria and MIC obtained by the microdilution method were ≤12.25 µg/mL for all the tested strains [48]. An extensive study on the antibacterial activity of AgNPs against A. baumannii AIIMS 7 in planktonic and biofilm mode demonstrated that AgNPs inhibited planktonic bacteria at a concentration of 16 μg/mL and exhibited a synergistic interaction with doxycycline, tetracycline, and erythromycin [69] in agreement with other studies [65].
The importance of visible light to induce the bactericidal mechanism of AgNPs has been reported by Shi et al. [70], who claimed that this is a key factor to catalyze the massive aggregation of cellular proteins in bacteria without the need for silver ion release or formation of ROS. AgNPs affected bacterial growth, distorted the cellular morphology, and induced intracellular oxidative stress thus rendering these bacteria susceptible to NPs. AgNPs interact with thiol-groups, which indicates their potential to inactivate cellular proteins [71]. AgNPs synthesized and functionalized with two capping agents (3-mercapto-1-propane sulfonate and 1-thio-D-glucose) were active against A. baumannii ATCC19606T and other bacteria belonging to the ESKAPE (Enterococcus faecium, S. aureus, Klebsiella pneumoniae, A. baumannii, P. aeruginosa, and Enterobacter species) group [71]. In vitro studies using AgNPs incubated with different antibiotics showed synergistic antibacterial activities against ESKAPE [72]. AgNPs can simultaneously induce apoptosis and inhibit new DNA synthesis in MDR A. baumannii in a dose-dependent manner [73]. Hetta et al. [74] assayed the activity of AgNPs against MDR A. baumannii in vitro. AgNPs produced marked inhibition zones in all tested bacterial strains (mean = 16 mm and range = 6–27 mm) at a level of 50 μg/mL and even on biofilms; the inhibitory activity was more pronounced on weak biofilm producers and this activity was due to a decrease in the expression of some genes related to the formation of biofilms. Thus, we believe that AgNPs can be useful to combat MDR A. baumannii, especially by exploiting their synergy with antibiotics; however, their application in human patients needs more research to resolve the associated toxicological issues, which are reviewed in Section 7.
Table 1. Research on the activity of AgNPs against A. baumannii in vitro.
Table 1. Research on the activity of AgNPs against A. baumannii in vitro.
Synthetic Method of the AgNPsAgNP Size (nm)Particle ShapeCappingAntibiotic AddedMIC/MBC (µg/mL)Proposed Mechanism of ActionRef.
From Acinetobacter calcoaceticus8–12SphericalNoAMI, AMP, AMX, CAZ, CHL, CIP CRO, DOX, GEN, KAN, PEN, TET, TMP, VANMIC (antibiotics + AgNPs) from <0.015 to 2048, depending on the antibiotic; the lowest MIC for DOX, TET and TMP; CIP (0.125), AMI, GEN, KAN (2); CAZ (512, CRO (2048)Synergy between AgNPs and antibiotics, except for cephalosporins. No MIC data for CHL and VAN[57]
From Cassia fistula fruit50–150Triangular, hexahedral, amorphousNoNo62 µl/mL (in vitro assay)Mechanism of action was not suggested[58]
From Salvia leriifolia leaf27 (avg.)SphericalNoNo(101.4 ± 2.4)% inhibition vs controlMechanism of action was not suggested[59]
From PVP6–10 (TEM)SphericalPVPNo. AgNPs were compared with CAR and other antibiotics but not mixedMIC for IPM: 64 (vs a highly resistant strain), 32, 8, 8. Four resistant strains were assayed. No MIC for PVP-AgNPs was provided, but they were active against 3 of the 4 strainsMechanism of action was not suggested. Reference to previous work[60]
Reduction with PVP, or Na citrate, or SDS, or chitosan (Chit)6–10 (PVP);
Not indicated (others)
SphericalPVP,
Citrate (CIT),
SDS, Chit
AMP, DOR, IPM64 (highly resistant strain)Synergy with IPM and DOR, (CIT-AGNPs: IPM (Chit- and SDS-AgNPs).
Synergy with DOR and AMP (PVP-AgNPs)
[61]
Reduction with Na citrate; then capping with SH-PEG2000-NOTA30 (avg.)SphericalSH-PEG-NOTA + IPMIPM64 (at conc. of 60–100 µg/mL)Mechanism of action not suggested.
Sinergy between AgNPs and IPM
[62]
Reduction with NaBH4 + Na citrate5–12
8.4 (avg.)
Spherical mainlyCitratePMB, RIF, TGCMIC: 2.5 (AgNPs alone), FIC index: 0.19 (PMB), 0.38 (RIF), 0.75 (TGC)Sinergy with PMB and RIF; additive effect with TIG[63]
Axonnite® prepared by micro-explosion2–5 (70–75%); 5–100 (30–25%)Not indicatedNoNo0.39–0.78Mechanism of action not indicated[64]
From Xanthomonas spp.<10SphericalXanthan gumNoNot indicatedMechanism of action not indicated[65]
From Dioscorea bulbifera8–20Mostly spherical some nanorod,
triangle
NoAminoglycosides, β-lactams, cephalosporins, CAR, PMB, VAN, and othersNo MIC was given. Only inhibition diameters on solid phase cultures were providedSinergy with β-lactams (mainly PIP) and ERY[66]
From bacteria 8-12VariableNADOX, ERY, TETMIC: 16 (against planktonic cells);
MBEC: 2000 (against biofilms)
Synergy with DOX, TET and ERY. Intracellular oxidative stress; interaction with thiol-groups[69]
Commercial 11.12 ± 0.07SphericalPVPNoMIC: 0.9 (MDRAB)
MIC: 2.1 (against a sensitive strain of A. baumannii ATCC 19606)
Photocatalytic induction of massive aggregation of cellular proteins under visible light. This process is not dependent on the bacterial species[70]
Reduction with NaBH4. Then mix with 3MPS and TG (variable ratios) or only with 3MPS 3 ± 1,
6 ± 2 or 10 ± 2 (by DLS)/
15–20 (by TEM)
Spherical3MPS-TG (two patterns) and 3MPSNoIC90 > 128 for A. baumannii ATCC19606Not reported[71]
Commercial 5–10Not indicatedNo15 antibiotics were used for assessing bacterial resistance. They were not mixed with AgNPs.CFU results showed that 38 MDRAB clinical isolates from hospital patients were sensitive to the AgNPs. MIC and MBC were not givenAgNPs induced apoptosis in MDRAB clinical isolates. This activity increases with increasing AgNP conc.
Bacterial DNA synthesis decreases with increasing AgNP level
[73]
Reduction with PVP 10–50SphericalNot indicatedNoMIC: 4–25 depending on the ability to produce biofilms more or less strongAgNPs significantly interrupted bacterial growth and multiplication[74]
Abbreviations: AMI: amikacin; AMO: amoxicillin; AMP: ampicillin; avg.: average; CAR: carbapenems; CAZ: ceftazidime; CFU: colony forming units; CHL: cholesterol; CIP: ciprofloxacin; CRO: ceftriaxone; DLS: dynamic light scattering; DOR: doripenem; DOX: doxycycline; ERY: erythromycin; FIX: fractional inhibitory concentration; GEN: gentamicin; IC90: 90% inhibitory concentration; IPM: imipenem; KAN: kanamycin; MBC: minimum bactericidal concentration; MBEC: minimum biofilm eradication concentration; MDRAB: multiple drug-resistant Acinetobacter baumannii; MIC: minimum inhibitory concentration; 3MPS: 3-mercapto-1-propanesulfonate; NA: Not available; NOTA: 1,4,7-triazacyclononane-1,4,7-triacetic acid; PEG: polyethylene glycol; PEN: penicillin; PIP: piperacillin; PMB: polymyxin B; PVP: polyvinylpyrrolidone; RIF: rifampicin; ROS: reactive oxygen species; SDS: sodium dodecyl sulfate; TEM: transmission electron microscopy; TET: tetracycline; TG: 1-thioglucose; TGC: tigecycline; TMP: trimethoprim; VAN: vancomycin.

4. Pseudomonas aeruginosa

Pseudomonas aeruginosa is a Gram-negative bacterium highly susceptible to genetic changes leading to resistance to antimicrobials and the consequent complications in impaired or immunocompromised patients. It can form biofilms [68]. Due to its ability to survive in harsh environments, P. aeruginosa is one of the most important agents in nosocomial infections [75] and its disease spectrum continues to expand from urinary tract infection to septicemia, osteomyelitis, and endocarditis, posing new challenges because resistance to current therapy limits the available treatment options [76]. Among the different nanosized antibacterial agents, silver is the most effective because of its broad-spectrum activity against bacteria, viruses, and eukaryotic microorganisms [77]. A summary of the research performed on the usage of AgNPs against P. aeruginosa is shown in Table 2. Morones et al. [77] proposed that AgNPs act in three ways against Gram-negative bacteria: (1) AgNPs mainly in the range of 1–10 nm attach to the surface of the cell membrane and drastically disturb its proper functioning, such as permeability and respiration; (2) they can penetrate inside the bacteria and cause further damage by possibly interacting with sulfur- and phosphorus-containing compounds, such as DNA; and (3) AgNPs release Ag+ ions, which additionally contribute to the bactericidal effect of the AgNPs. Biogenic AgNPs (60–80 nm size range) showed activity against an antibiotic-resistant strain of P. aeruginosa and enhanced the antimicrobial activity of ampicillin, gentamicin, vancomycin, and streptomycin when combined with them [78]. Biosynthesized AgNPs in the 20–50 nm size range also exhibited high antimicrobial activity against this bacterium [79,80]. The MIC and the minimal bactericidal concentration (MBC) for P. aeruginosa (6.4 pM) were higher than for other bacterial strains, which could be due to their biofilm-forming ability [80]. The MBC of AgNPs synthesized from P. putida (15–40 nm) against P. aeruginosa was 1 µg/mL [81]. Spherical silver nanocomposites (5–50 nm) biosynthesized from Lactococcus lactis exerted antimicrobial activities against P. aeruginosa and the MIC90 against this bacterium was 6.25 μg/mL [82]. AgNPs showed synergistic activity against MDR P. aeruginosa when combined with antibiotics. AgNP functionalization with ampicillin (AMP-AgNP) showed advantage over non-functionalized AgNPs as they killed ampicillin-resistant strains of P. aeruginosa (MBC = 1 µg/mL) [83].
AgNPs can inhibit biofilm formation by P. aeruginosa in ocular-related infectious disease microbial keratitis by more than 95% by arresting the synthesis of the exopolysaccharide matrix [84]. Such effectiveness in inhibiting biofilm formation was found at bacterial levels of 104–105 CFU/mL with inhibition rates of 56–67% [85]. The activity of 10-nm size commercial AgNPs against P. aeruginosa strains with resistance to some antibiotics was evaluated using concentrations of 0.156–5.0 µg/mL. After 12 h, a dose of 5.0 µg/mL proved very effective (approximately 99.9% bacterial death), even when tested against hospital MDR strains [75]. Habash et al. [86] assessed the efficacy of citrate-capped AgNPs alone and combined with the antibiotic aztreonam against P. aeruginosa biofilms in vitro. The effects of aztreonam alone were limited or even enhanced biofilm biomass since doses as high as 512 µg/mL had no lasting effect on cell viability within the biofilm. AgNPs (10–20 nm) evaluated individually showed more efficacy than aztreonam in preventing biofilm and planktonic cell recovery. The 10 nm AgNPs alone presented only minor, but significant defects in biofilm architecture, even when alterations in cellular morphology or ultrastructure were not evident. In contrast, AgNPs (40–60 nm) demonstrated limited inhibition of biofilm biomass and viability, while the 100 nm AgNPs showed no major inhibitory effects [86]. Citrate-capped AgNPs proved more effective against P. aeruginosa biofilms than applying ionic Ag, which indicates that AgNPs can release small amounts of Ag+, thus resulting in higher activity. AgNPs of 8 nm were more efficient in detaching P. aeruginosa biofilms than particles of 20 and 35 nm [87]. The biofilm removal effect of AgNPs was size-dependent, as the smaller nanoparticles showed higher effectiveness. Citrate-capped AgNPs (10–20 nm) synergistically potentiated tobramycin activity to inhibit these biofilms, which may be due to the disruption of cellular membranes [88]. Minimum biofilm eradication concentration assays using clinical P. aeruginosa strains showed that small AgNPs inhibited biofilms better than larger AgNPs, although the synergy effect is likely strain-dependent. The MIC of AgNPs against this bacterium is in the order of 1–2 μg/mL [46,89]. Thus, the AgNP antibacterial effects against this bacterium are dose- and time-dependent [90,91]. The antibacterial activity of AgNPs is due to the generation of ROS, malondialdehyde, and leakage of proteins and sugars in bacterial cells. Moreover, AgNP-treated bacteria had significantly lower lactate dehydrogenase (LDH) activity and lower adenosine triphosphate levels than the control. Furthermore, AgNP-treated bacteria showed downregulated expression of glutathione, upregulation of glutathione S-transferase, and downregulation of both superoxide dismutase and catalase. These physiological and biochemical measurements observed in AgNP-treated bacteria suggest that AgNPs can induce bacterial cell death. The antibacterial activity of the biosynthesized AgNPs was evaluated against Gram-negative bacteria, such as E. coli and P. aeruginosa [46] and it was found that AgNPs showed bactericidal rather than bacteriostatic effect; a bactericidal agent is preferred clinically because bacterial killing should produce a faster resolution of the infection, improves clinical outcome, and reduces the likelihood of the emergence of resistance and the spread of infection [90]. Biosynthesized AgNPs (7.1 nm) exhibited antibacterial activity and killed all the cells of this bacterium when treated with 2.7 μg/mL for 4 h [92]. The intracellular ROS production suppressed the antioxidant defense and exerted mechanical damage to the membrane. AgNPs inhibit the catalase and peroxidase activity so that the excessive ROS is not eliminated, which may result in impaired DNA and ribosome and declined synthesis of the macromolecules [91,93]. AgNPs also induce surface charge neutralization and alter the cell membrane permeability, causing non-viability of the cells. The synergistic effect of AgNPs combined with antibiotics against both susceptible and resistant P. aeruginosa was evaluated and was shown only against susceptible P. aeruginosa [94]. This synergistic effect has been reported not only for P. aeruginosa but also for other bacteria [95,96]. However, strains of P. aeruginosa resistant to streptomycin and rifampicin retained their resistance when these antibiotics were combined with AgNPs [94]. The activity of biogenic AgNPs alone and combined with antibiotics was evaluated against P. aeruginosa, and the lowest MIC found was 16 μg AgNPs/mL [97], which is higher than other reported MIC values [46,95] although differences may be due to differences in size. Electrochemically synthesized AgNPs were able to reduce the biofilm’s viability of P. aeruginosa achieving biofilm suppression at a level of 17 μg/mL (4 × MIC) [98]. By using proteomic analysis, it was demonstrated that the mechanisms of the AgNP antibiofilm activity involve interferences with multiple processes of the P. aeruginosa biofilm formation, such as bacterial motility, oxidative stress response, iron homeostasis, respiration, and quorum sensing systems [93]. Therefore, AgNPs exhibit a remarkable antibacterial activity against MDR P. aeruginosa, representing a possible alternative for antibiotics and they can also be promising antibiofilm agents. The systemic administration of these particles seems at this time difficult to implement owing to the possible accumulation and damage of tissues and organs (toxicity); however, either coating of prosthetic devices/catheters or their topical application for the treatment of skin infections and the prevention of disease in burnt patients may be a future application of AgNPs [99,100]. Therefore, the utility of AgNPs against this bacterium has been evidenced. As in the case of A. baumannii, their usefulness in treating human infections requires more research.

5. Enterobacteriaceae

The Enterobacteriaceae family is ubiquitous and its members are found worldwide in different ecological sources. Some species are part of the normal flora of animals, including humans, although many are frequently associated with diarrheal disease and extraintestinal infections. This family includes more than 210 species and 53 genera, and these numbers continue to increase. Some of the most important pathogens in human history, such as Yersinia pestis, belong to this family. Other pathogens of huge public health concern are Salmonella enterica serotype or serovar Typhi, Shigella spp., Klebsiella pneumoniae and E. coli. Other Enterobacteriaceae causing infections in humans include Citrobacter spp., Enterobacter spp., Serratia marcescens, or Proteus spp.
A recent problem in the medical field is the increasing number of bacterial strains within the Enterobacteriaceae family able to produce extended-spectrum β-lactamases (ESBL) [101]. The carbapenems are the mainstay of therapy for treating serious and life-threatening infections caused by Enterobacteriaceae producing ESBL, but the emergence of resistance to carbapenems has led to limited therapeutic options. Mechanisms of resistance to carbapenems include the production of β-lactamases, efflux pumps, and mutations that alter the expression and/or function of porins [102].
Using E. coli as a model for Gram-negative bacteria, it was proved that negatively charged AgNPs may be used as antimicrobial agents. Table 3 lists some relevant results attained in the research on the application of AgNPs against enterobacteria. Aggregates composed of AgNPs and dead bacterial cells were observed by SEM. The AgNPs interact with elements of the bacterial membrane and damage the cell. TEM analysis and Energy Dispersive Analysis X-ray confirmed the incorporation of AgNPs into the membrane and the formation of pits on the cell surface [103].
The antibacterial efficiency of AgNPs was tested against E. coli in solution and solid medium. The AgNPs exhibited antibacterial activity at low levels and they were cytotoxic at a concentration of 8 μg/cm2 of solid culture medium [104]. The mechanism behind the antibacterial activity of AgNPs was related to their high surface area/volume ratio. The effectiveness of penicillin G, amoxicillin, erythromycin, clindamycin, and vancomycin against E. coli increased in the presence of AgNPs [105]. Highly monodispersed AgNPs (average size 13.5 nm) inhibited E. coli growth, with MIC > 3.3 nM, and the inhibition effect was dose-dependent [106]. Biosynthesized AgNPs having an average size of 30.5 nm in a protein matrix were very effective to control the growth of K. pneumoniae [107]. Spherical AgNPs (5–40 nm) biosynthesized from Fusarium acuminatum showed an efficient antibacterial effect against E. coli, S. Typhi, and Staphylococcus epidermidis, which was 1.4–1.9 fold stronger than that of Ag+ [108]. AgNPs biosynthesized from E. coli proved very efficient against E. coli giving an MBC of 8 µg/mL [81].
Chitosan-capped silver nanocomposites had activity against E. coli higher than that of chitosan alone [109,110]. Large AgNPs (160–180 nm diameter) showed only moderate activity against S. Typhi and K. pneumonia and were inactive against Vibrio cholera [111]. Clinical strains of E. coli, K. pneumoniae, Enterobacter sp., and Proteus morganii isolated from patients affected by urinary tract infections were treated in vitro with AgNPs. Enterobacter sp. proved very susceptible but the E. coli strain was not affected. This was attributed to the MDR nature of E. coli developed by point mutations [112]. However, there is evidence that small AgNPs (5 nm) can inhibit growth and even kill the cells of E. coli by destroying the bacterial membranous structure and altering its permeability [113], which agrees with Devina Merin et al. [114] who found that AgNPs provided good results in terms of inhibition of strains of Klebsiella sp., Proteus vulgaris, and E. coli. Salmonella Typhimurium was less sensitive to AgNPs (20–30 nm size range) than E. coli [115]. Concerning capping agents, PVP-AgNPs showed better antibacterial properties both in vitro and in vivo than citrate-capped AgNPs [116]. This may be due to the better stability and higher uptake by the cells of the PVP-AgNPs. Cell uptake of capped AgNPs was significantly higher than for uncapped AgNPs in the presence of serum. The antibacterial activity of N-stearoyl ethanolamine (NSEA) capped AgNPs was tested against pathogenic bacteria, such as E. coli, S. Typhi, Shigella spp., and K. pneumoniae, and their estimated MIC values ranged from 6 to 12 μg/mL [117]. The formation of ROS in E. coli after damage to the cell surface has been suggested as a cause of bacterial toxicity due to the activity of AgNPs [118].
As indicated in previous sections, synergistic effects between AgNPs and some antibiotics also occur with enterobacteria [78,96,119,120,121]. AgNPs showed antibacterial activity against MDR strains that produce a broad spectrum of β-lactamases or carbapenemase (ESBL-positive E. coli, ESBL-positive K. pneumoniae, AmpC-positive E. coli, and K. pneumoniae-carbapenemase (KPC)-positive K. pneumoniae) when combined with cefotaxime, ceftazidime, meropenem, ciprofloxacin, and gentamicin, as reflected by low MIC values [96]. The strongest activity was demonstrated against ESBL-positive E. coli, as only 0.8 mg AgNPs/L killed the bacteria while the highest MIC (6.75 mg/L) was required to kill ESBL-positive K. pneumoniae. This study of the antibacterial activity of AgNPs combined with antibiotics confirmed the existence of a synergistic effect resulting from the combination of these two antimicrobial agents. A lack of β-lactamase production in bacteria when antibiotics were combined with AgNPs confirmed the restoration of the antibacterial effect of antibiotics in the presence of Ag [96].
As found for P. aeruginosa, AMP-AgNPs have also an advantage over enterobacteria because they kill ampicillin-resistant strains of E. coli. MBCs were 1 µg/mL against E. coli, 2 µg/mL against V. cholera, and 4 µg/mL against E. aerogenes [83]. Regarding the mechanism of action of AgNPs against E. coli, the stabilities of the antibacterial activity under various pH-values and temperature conditions, the protein leakage caused by increased membrane permeability, and the inactivation of LDH due to the nanoparticle-induced formation of ROS were demonstrated [122]. The antibacterial activities of ampicillin, kanamycin, erythromycin, and chloramphenicol increased in the presence of AgNPs against test strains [120]. AgNP-loaded TiO2 nanotube arrays were fabricated on titanium implants for a customized release of Ag+. The antibacterial properties of silver nanotubular structures combined with vancomycin, rifampicin, gentamicin, and levofloxacin were tested in vitro. Improved effectiveness of the combined therapy was observed for all tested bacterial strains, including E. coli. After the treatment, experiments further proved the synergistic antibacterial effect both in vitro and in vivo [121]. Although bacterial resistance to antibiotics is well-known, and bacterial resistance to Ag has been reported [122], the possible development of resistance to AgNPs has not been fully explored. Susceptible strains of E. coli and other bacteria were converted into AgNP-resistant strains by culturing them in agar media containing AgNPs until a concentration near or over the MIC was reached on which the bacteria could grow [90]. AgNPs (10–25 nm) mostly spherical were assayed against enterobacteria (Salmonella spp. and Shigella spp.) isolated from poultry feces; at a concentration of 16 μg/mL AgNPs showed bacteriostatic and bactericidal effects against S. Montevideo, Shigella sonnei, and S. enteritidis but at a concentration of 8 μg/mL the nanoparticles had both bacteriostatic and bactericidal effects in the case of S. Poona, Shigella boydii, and S. Typhimurium [123]. Lower MIC and MBC values (0.085 ± 0.126 μg/mL and 0.508 ± 0.315 μg/mL, respectively) were obtained for 20 MDR Salmonella spp. strains recovered from feces of diarrheal sheep and goats when treated with PVP-capped chemically synthesized AgNPs (2.95–12.2 nm size range). Interestingly, an in vivo assay on a mouse model showed that AgNPs had no toxic or pathologic effects [124]. In a recent study, biosynthesized AgNPs from Massilia sp. showed strong antimicrobial activity against K. pneumoniae and S. aureus enteritidis. The MICs of biosynthesized AgNPs against K. pneumoniae and S. enteritidis were 12.5 and 25.0 μg/mL, respectively while the MBC against both pathogens was 50.0 μg/mL [125]. A study on the susceptibility of three different Salmonella serovars (Enteritidis, Hadar, and Senftenberg) to AgNPs showed an immediate, time-limited, and serovar-dependent reduction of bacterial viability [126]. For S. Senftenberg, the reduction was observed for up to 4 h of incubation in the presence of 200 µg AgNPs/mL; on the contrary, S. Enteritidis and S. Hadar were inhibited for up to 48 h. Thus, success is strongly Salmonella strain-dependent, since great differences in terms of effective dose and time of action were observed for the examined serovars. Reverse transcription and PCR experiments demonstrated the constitutive expression of the plasmidic silver resistance determinant (SilB) by S. Senftenberg, thus suggesting the importance of a cautious use of AgNPs [126]. The use of AgNPs conjugated with antimicrobial peptides such as andersonin Y1 (AY1) and two AY1-cysteine derivatives has been recently explored and found useful as a new strategy to combat MDR bacteria (P. aeruginosa, E. coli, K. pneumoniae, and S. Typhi) [127].
The mechanism of action of Ag/AgCl NPs is similar for different enterobacteria (Serratia marcescens, a strain of K. pneumoniae carbapenemase-producer, a strain of ESBL K. pneumoniae and E. coli ATCC 25922) and independent of the presence of enzymatic mechanisms of resistance to β-lactamase. This finding was confirmed by the MIC determination since all bacterial strains showed the same sensitivity profile (MIC = 10.52 μg of Ag/AgCl NPs/mL) [128]. Two E. coli strains developed resistance to AgNPs after repeated exposure due to the production of the adhesive flagellum protein flagellin, which triggers nanoparticle aggregation. Resistance evolved without any genetic changes; only phenotypic change was needed to reduce the NP colloidal stability and eliminate their antibacterial activity. The resistance mechanism was not overcome by additional stabilization of AgNPs using surfactants or polymers but it was strongly suppressed by inhibiting flagellin production with pomegranate rind extract [129]. The antibacterial usefulness of AgNPs against some of the most important species of the Enterobacteriaceae family has been put in evidence but its clinical application requires further research as previously indicated for other bacteria.
Table 2. Research on the activity of AgNPs against Pseudomonas aeruginosa in vitro.
Table 2. Research on the activity of AgNPs against Pseudomonas aeruginosa in vitro.
Synthetic Method of the AgNPsAgNP Size (nm)Particle ShapeCappingAntibiotic AddedMIC/MBC (µg/mL)Proposed Mechanism of ActionRef.
Commercial in a carbon matrix16 ± 8Cub-octahedral, multiple-twinned icosahedral and decahedralNoNoNot providedDisruption of bacterial membrane altering permeability/respiration; damage of S and P containing compounds (DNA); AgNPs release Ag+[77]
From Phoma glomerata60–80SphericalBio-moleculesAMP, GEN, KAN, STR, VANNot providedSynergy with AMP, GEN, VAN, and STR[78]
From Pseudomonas aeruginosa20–50SphericalNot indicatedNo50Reference to the mechanisms of action given in ref. [77][79]
From Pseudomonas aeruginosa25–45SphericalNot indicatedNoMEC: 6.4 pM
MBC: 6.4 pM
Reference to previously reported mechanisms. Effect on the membrane: release of Ag+[80]
From Pseudomonas putida15–40Spherical, truncated triangle, triangle, and hexagonalBiological coronaNoMIC: 1Reference to the same mechanisms given in ref. [77][81]
From Lactococcus lactis 565–50;
avg. 19 ± 2
SphericalOrganic material from L. lactisNo6.25No mechanism of action was proposed[82]
Reduction with NaBH4 + Na citrate4Not indicatedCitrate/AMPAMP linked to AgNP, (AMP-AgNP)MBC: 1 for AMP-AgNPs, against all the tested bacteria, four times lower than AgNPs alone (MBC: 4)No mechanism of action was proposed[83]
From Bacillus licheniformis50 (avg.)SphericalNot indicatedNoA concentration of 100 nM inhibits 95–98% biofilmBiofilm inhibition by arresting the synthesis of the exopolysaccharide matrix[84]
Commercial20–30Not indicatedNot indicatedNo20 µg/mL inhibits about 56% biofilm of MDR P. aeruginosa and 67% of other strainBiofilm inhibition. Mechanism of action was not provided[85]
Commercial10, 20, 40, 60, and 100Not providedCitrateATM MIC against planktonic cells of AgNPs alone: from 0.234
(10 nm particles) to 7.50 (100 nm particles)
Synergistic effects of AgNP/ATM against biofilms are size-dependent.
Optimal size: 10 nm, followed by 20 nm; worst size: 100 nm
[86]
Commercial10Not providedCitrateNo5, approx. 99.9% MDR P. aeruginosa deathSuggested mechanisms of action are those proposed in refs. [77,103,113][75]
(1) Reduction with gallic acid;
(2) growth of small AgNP nuclei by citrate/ascorbic acid addition at various Ag+/Ag0 ratios
8 (avg.)
before seeding; then up to 66 nm. Sizes used in experiments were 8, 20 and 35 nm
Spherical; pseudo-spherical, cylindrical.Other shapes after growthCitrateNo600 µg AgNPs/mL produced biofilm detachment in 90% (8-nm AgNPs), and lower % (20- and 35-nm AgNps) depending on the media usedEffectiveness is size related. Low sizes are more effective than high sizes against biofilms. The low efficacy of AgNPs in this study may be due to citrate capping. AgNPs are more efficient than silver ions. Attachment of the NPs onto the microbial cell membrane leads to increased permeability, inhibition of cell wall synthesis, plasmolysis, and cell death[87]
Commercial10, 20, 40, 60, and 100Not providedCitrateTOBMIC against planktonic bacteria: 0.156–0.625 (10 nm); 0.312–2.5 (20 nm); 2.50–10 (100 nm). MEBC against biofilms: 1.25–5.0 (10 and 20 nm); 2.5–>10 (40 nm); 5.0–>10 (60 nm)Synergistic effect (10, 20, 40, 60 nm). Additive effect: 100 nm. The efficacy to inhibit biofilms and planktonic cells is dependent on strain and it is higher for smaller AgNPs either alone or combined with TOB[88]
Quercetin11SphericalQuercetinNoMIC: 1Antibacterial activity due to membrane disruption, generation of malondialdehyde and ROS, and leakage of proteins and sugars in cells. Found in treated cells: downregulated expression of glutathione, upregulation of glutathione S-transferase, downregulation of superoxide dismutase and catalase; inactivation of respiratory chain; low lactate dehydrogenase activity, and low adenosine triphosphate[46]
From A. baumannii37–168SphericalNot indicatedNoMIC: 1.56Not provided[89]
Commercial100Not indicatedNoNoMIC: 83.3 (±16.7) mM
MBC: 83.3–100 mM
Anti-biofilm activity not reported
It is suggested a mode of action of AgNPs previously reported and similar to that of Ag+, which complex groups containing S, O, or N atoms that are present as thiols or phosphates on amino acids and nucleic acids, ROS production, membrane destabilization, etc.[90]
Commercial from cyclodextrin5–20, mostly 5–10 (chosen for the experiment)Near sphericalNo indicatedNoMIC: 1.406–5.625; MBC: 2.813–5.625The cell wall becomes thin; the cell membrane shrivels and fractures. Production of excessive ROS (oxidative stress); destruction of the redox homeostasis; alteration of the activity of the redox relevant enzymes; apoptosis-like effect. Activity is dose- and time-dependent[91]
From protein cell-free extract of Rhizopus oryzae9.2 (avg.)SphericalProtein coronaNoMIC: 2.25 ± 0.2
MBC: 2.7 ± 0.2
Excessive ROS production. Cell membrane permeability is affected. Membrane destabilization by ROS can be
responsible for surface charge neutralization leading to cellular material leakage and cell death. Damages due to AgNP interaction with intracellular proteins and nucleic acids
[92]
Commercial from cyclodextrin5–20, mostly 5–10 (chosen for the experiment)Near sphericalNo indicatedNo6.25 prevents biofilm formationBiofilm damage is dose-dependent. AgNPs may induce downregulation of flagellins, fimbrillins, and other proteins of biofilms. Bacterial adhesion and motility are inhibited. The iron homeostasis is disturbed. Excessive ROS can cause lipid peroxidation, impairment of DNA and ribosomes, reduction in synthesis of macromolecules, and bacterial death. Respiratory enzymes are affected, which conducts to hypoxia. ROS production may influence the QS system and inhibit the expression of the virulent factors[93]
High-voltage method2–35Not indicatedNoAMP, CIP, CTZ, MEM, OXA, RIF, STR, TETMIC: 1;
MBIC: 4 (for AgNPs alone)
Synergistic interaction with AMP, STR, RIF, and TET. No interaction with the remaining antibiotics (planktonic cells). The synergistic interactions depend on the doses. No interaction concerning biofilm formation was observed. AgNPs induced synthesis of bacterial DnaK chaperone, but HtpG chaperone synthesis was unaffected[94]
Reduction of [Ag(NH3)2]+ by D-maltose
(modified Tollens process)
26Not indicatedNoAMI, ATM, CFP, CIP, CST, CTZ, FEP, GEN, MEM, OFX, PIP, TZPMIC: 7.5 (AgNPs alone)Synergistic effect of antibiotics combined with AgNPs[95]
From Streptomyces xinghaiensis5–20 (TEM)
64 (avg.) (nano tracking analysis)
SphericalNot providedAMP, KAN, TETMIC: 16 and
MBC: 32 (AgNPs alone)
No interaction between AgNPs and the tested antibiotics is reported[97]
Electrochemical process55.6 ± 2.9Quasi-sphericalNoTOBMIC: 1.07–4.25 and
MBC: 2.125–4.25 (for AgNPs alone)
AgNPs exhibited a comparable or higher antibacterial activity compared to TOB including anti-biofilm activity. AgNPs showed a dose-dependent effect and caused biofilm eradication at a concentration of 4 × MIC. They deconstructed the exopolysaccharide matrix and produced cell lysis[98]
From (a) leaf extract of Citrus latifolia; or (b) from Aspergillus flavus5–70, mostly in the range 20–30SphericalNoAMI, CAZ, CIP, KAN, LVX, MEM, TZPAgNPs alone, MIC: 4–128; for AgNPs from A. flavus: or 8–>128 for AgNPs from C. latifoliaDamage to the cell wall, membrane, and DNA, induction of ROS production. AgNPs derived from A. flavus showed synergistic effects with MEM and LVX[99]
Abbreviations: AMI: amikacin; AMP: ampicillin; ATM: aztreonam; avg.: average; CAZ: ceftazidime; CFP: cefoperazone; CIP: ciprofloxacin; CST: colistin; FEP: cefepime; GEN: gentamicin; KAN: kanamycin; LVX: levofloxacin; MBC: minimum bactericidal concentration; MBIC: minimum biofilm inhibitory concentration; MEBC: minimum eradication biofilm concentration; MDR: multi-drug resistant; MEM: meropenem; MIC: minimum inhibitory concentration; mM: millimolar; OFX: ofloxacin; OXA: oxacillin; PIP: piperacillin; pM: picomolar; QS: quorum sensing; RIF: rifampicin; ROS: reactive oxygen species; STR: streptomycin; TEM: transmission electron microscopy; TET: tetracycline; TOB: tobramycin; TZP: piperacillin/tazobactam; VAN: vancomycin.
Table 3. Research on the activity of AgNPs against Enterobacteriaceae in vitro.
Table 3. Research on the activity of AgNPs against Enterobacteriaceae in vitro.
Synthetic Method of the AgNPsAgNP Size (nm)Particle ShapeCappingAntibiotic AddedMIC/MEB (µg/mL)Proposed Mechanism of ActionRef.
Ascorbic acid + Daxad® 1912 (mode)Not indicatedNot indicatedNoMIC: 50–60 (against E. coli)Formation of ‘pits’ in the bacterial cell wall. AgNPs accumulate on the cell wall/membrane and in the cells. Leaking of intracellular substances[103]
From K. pneumoniae5–32, 22.5 (avg.)Not indicatedNot indicatedYesNot determined. Assays on solid media. Inhibition diameters measuredNo mechanism of action was proposed. AgNPs + antibiotics increase the inhibition zone more than some antibiotics alone[105]
Reduction with NaBH44–20, 13.4 (avg.)Not indicatedNot indicatedNoMIC: 3.3–6.6 nM (against E. coli)Formation of free radicals on the AgNP surface and free radical-induced membrane damage[106]
From spent mushroom substrate30.5 ± 4.0SphericalProteinsNoNot determinedNo mechanism was proposed. Antibacterial properties against K. pneumoniae increased with AgNP concentration[107]
From Fusarium acuminatum5–40; 13 (avg.)SphericalNot indicatedNoNot determined. E. coli and S. Typhi were inhibited, but the efficacy was lowReference to mechanisms proposed in ref. [103]. The reaction of silver with SH groups of proteins in the cell inactivates proteins[108]
From chitosan solution + NaOH producing chitosan-AgNP composites2–4 (90%)SphericalNot indicatedNoAgNP-chitosan composite: MIC: 100;
MBC: 120 (against E. coli)
Destabilization of the bacterial cell wall by the composite. The binding of AgNPs to thiol-containing proteins present in the cell wall leads to penetration. The composite was more efficient than AgNPs or chitosan alone for inactivating bacteria, possibly due to a synergistic effect[109]
From chitosan solution4–18, 6–8 (50%)Not indicatedNot indicatedNoAgNP-chitosan;
MIC: 10;
MEB: 10 (against E. coli)
Chitosan-based AgNPs have a dual mechanism of action for antibacterial activity, the bactericidal effect of AgNPs, and the cationic effects of chitosan[110]
From S. aureus160–180Not indicatedNot indicatedNoMIC/MBC were not given. S. Typhi and K. pneumoniae showed low susceptibility; V. cholera was not susceptibleNo mechanism of action was proposed[111]
Commercial5Not indicatedNot indicatedNoMIC: 10 (against E. coli)AgNPs accelerate the reducing sugars/protein leakage from the cytoplasm in E. coli. The activity of respiratory chain dehydrogenases decreases with time. Cell membranes were severely damaged[113]
From various marine microalgaeNot indicatedNot indicatedProbably proteins from the sourceNoMIC was not given. Inhibition against Klebsiella spp., Proteus vulgaris and E. coli.No mechanism of action was proposed[114]
From Streptomyces hygroscopicus20–30SphericalNot indicatedNoMIC/MBC not given. The highest antimicrobial activity was against E. coli. Lower activity was found against S. TyphimuriumNo mechanism of action was proposed[115]
Reduction with:
(a) NaBH4, (b) Na citrate, (c) Ethylene glycol + PVP
(a) 75 ± 4.5
(b) 82 ± 5.2
(c) 86 ± 6.7
Not indicated(a) Uncapped
(b) Citrate (CIT)
(c) PVP
NoMIC: 6–6.33 (against S. Typhimurium);
6.33–6.83 (against S. Typhi);
6.67–7 (against Shigella flexneri).
No mechanism was proposed. The antibacterial activity was in the order PVP-AgNPs > CIT-AgNPs >> uncapped AgNPs[116]
NaBH4 + N-acylethanolamineNot indicatedSphericalN-acylethanol-amineNoMIC: 6.67 (against K. pneumoniae);
7.22 (against Shigella sp.); 7.22 (against S. Typhi); 9.06 (against E. coli)
No antibacterial mechanism was proposed[117]
From dried powder of Ocimum gratissimum leaf
extract
16 ± 2 (TEM)TriangularProteins from the sourceNoMIC: 4;
MBC: 8 (against MDR E. coli)
Intracellular ROS generation; membrane was fragmentary.
Inhibition of biofilm formation
[118]
From Phoma glomerata60–80SphericalBiomoleculesAMP, GEN, KAN, STR, VANNot providedAgNPs enhanced the antimicrobial activity of antibiotics against E. coli. Synergy with AMP, GEN, KAN, VAN, and STR[78]
Reduction of [Ag(NH3)2]+ by D-maltose
(modified Tollens process)
26Not indicatedNoCIP, CTX, CTZ, GEN MERMIC AgNPs alone: 0.8 (ESBL-positive E. coli); 3.4 (AmpC-positive E. coli and KPC-positive K. pneumoniae); 6.8 (ESBL-positive K. pneumoniae)Synergistic effects of antibiotics combined with AgNPs (< 1 µg/mL) against multi-resistant enterobacteria that produce broad-spectrum -lactamases or carbapenemase[96]
Ascorbic acid + Daxad® 1920CubicAMX linked to AgNPs (supposedly)AMXMIC AgNPs alone: 40 (against E. coli)Synergistic effects that may be caused by (a) bonding between AMX and AgNPs or (b) AgNPs can act as carriers of AMX[119]
From Trichoderma viride20–40Variable, spherical and other shapesNot indicatedAMP, CHL, ERY, KANMIC (AgNPs alone): 30 (against E. coli); 35 (against S. Typhi)Synergistic effects with all the assayed antibiotics. The effectivity order was AMP > KAN > ERY > CHL[120]
CommercialNot indicatedNot indicatedNoNoMIC: 100 (against E. coli)Bacterial protein leakage by increasing the membrane permeability. Formation of ROS that inactivate LDH[122]
Reduction with Na citrate + PVP10–25Most spherical, some prismaticNot indicated (probably citrate/PVP)NoMIC: 8–16; MBC: 8–16 (against Salmonella spp. and Shigella spp.)No antibacterial mechanism was proposed[123]
Reduction with: NaBH4, addition of Na citrate + PVP6.8 ± 2.28SphericalPVPNoMIC ≤ 0.002–0.313 (MBC: 0.078–1.250 (against Salmonella spp. 7 species, 20 strains)No antibacterial mechanism was proposed[124]
From bacteria Massilia sp.15–55, 23.2 (avg.) by TEM;
109.3 (avg.) by DLS
Most sphericalConjugated molecules not specifiedNoMIC: 12.5 (against K. pneumoniae) and 25.0 (against S. Enteritidis);
MBC: 50 (against both bacteria)
Morphological damage and distortion of the cell wall of both species. It can be attributed to oxidative stress due to the formation of ROS causing membrane detachment[125]
Commercial5 to > 500, mostly 6–20. Two populations within the rangeMost spherical, but also polygonalNoNoNot determined against 3 Salmonella serovars: Senftenberg, Hadar, and Enteritidis. Best conc. 200 µg/mL. AgNPs were most effective against S. Enteritidis, and not effective against S. SenftenbergThe surface area of AgNPs is important for their activity, as Ag+ release, the determining factor for antimicrobial activity, might be dependent on the surface area (importance of the AgNP shape)[126]
Reduction with NaBH4 and Na citrate. Further conjugation with peptides10 (avg.)Not indicatedCitrate. Then conjugation with peptides AY1 and two AY1 cysteine derivatives at the two terminal positions (C and N) of AY1: AY1C and CAY1NoMIC80%: ~50 µM (AY1-NP), 12 µM (AY1C-NP), 10 µM (CAY1-NP) against E. coli; 10 µM (AY1C-NP), 5 µM (CAY1-NP) against K. pneumoniae; 15 µM (AY1C-NP and CAY1-NP) against S. TyphiCell membrane rupture by nano-conjugates. It is suggested that there exists an interaction of peptides with negatively charged phosphate head groups of lipid moieties as well as with water molecules. Interaction with the hydrophobic tails of the membrane produces pores. Then, AgNPs attach the DNA[127]
From Fusarium oxysporum. Ag/AgCl-NP produced55 ± 18 (TEM);
89 (DLS)
Pseudo-sphericalProteinsIPMMIC of Ag/AgCl-NPs: 10.52 for all the bacteria tested (ESBL and K. pneumoniae carbapenemase-KPCAg/AgCl-NPs + IPM were more active than IPM alone, but no synergistic effect is deduced from the inhibition diameters[128]
According to refs. [95,96]26Not indicatedNoNoMIC increased for E. coli CCM 3954 from 3.38 to > 54 after 9 successive cultures and for E. coli 013 from 13.5 to > 54 after 14 successive culturesThe increasing MIC values show the gradual development of bacterial resistance against AgNPs, not against Ag+. Bacteria repeatedly exposed to sub-inhibitory concentrations of AgNPs can rapidly develop resistance to their antibiotic activity. Resistance is due production of flagellin, a protein of the bacterial flagellum, which causes AgNP aggregation[129]
Abbreviations: AMX: amoxicillin; avg.: average; AY1: andersonin-Y1; AY1C: andersonin-Y1-cysteine derivative at C-terminus; CTX: cefotaxime; CHL: chloramphenicol; CAY1: andersonin-Y1-cysteine derivative at N-terminus; DLS: dynamic light scattering; ERY: erythromycin; ESBL: Extended-spectrum beta-lactamase; IPM: imipenem; KAN: kanamycin; KPC: Klebsiella pneumoniae carbapenemase; LDH: lactate dehydrogenase; MBC: minimum bactericidal concentration; MIC: minimum inhibitory concentration; MIC80%: minimal concentration that kills 80% of bacteria; µM: micromolar; nM: nanomolar; PVP: polyvinylpyrrolidone; ROS: reactive oxygen species; TEM: transmission electron microscopy.

6. Staphylococcus aureus

Staphylococcus aureus is a Gram-positive coccus whose cells tend to occur either singly or forming pairs, tetrads, and distinctive irregular “grape-like” structures. Humans are usually colonized by S. aureus on external skin surfaces and the upper respiratory tract, particularly the nasal passages. Resistant strains typically produced β-lactamase, which inactivated the β-lactam antibiotics. Efforts were made to synthesize penicillin derivatives that were resistant to β-lactamase hydrolysis. This was achieved in 1959 with the synthesis of methicillin, which had the phenol group of benzylpenicillin disubstituted with methoxy groups [130]. In 1961 there were reports from the United Kingdom of S. aureus isolates that had acquired resistance to methicillin (methicillin-resistant S. aureus, MRSA), and MRSA isolates were soon recovered from other countries. MRSA is now a problem in hospitals worldwide and is increasingly recovered from nursing homes and the community [131]. S. aureus has also become resistant to other last-resort antibiotics, such as vancomycin, daptomycin, and linezolid [132].
Table 4 lists the most relevant results obtained by the application of AgNPs against S. aureus in vitro.
AgNPs can inhibit S. aureus growth at concentrations > 33 nM, around ten times higher than those required by E. coli [106]. Biosynthesized AgNPs performed successful antimicrobial testing against S. aureus [107,133]. AgNPs may be applied to MRSA, the main cause of nosocomial infections worldwide [90,132,134,135,136,137]. AgNPs at sublethal doses together with ampicillin act synergistically against MRSA, with the effect being more pronounced when a lower concentration of ampicillin is present [136]. Synergy effects on MRSA between streptomycin and AgNPs and other nanoparticles have been reported recently [138].
AgNPs act as potential antimicrobial agents and help to inhibit biofilm formation by MRSA and vancomycin-resistant S. aureus (VRSA) [139]. According to Ayala-Núñez et al. [134], AgNPs inhibit bacterial growth of both MRSA and non-MR S. aureus in a bactericidal rather than a bacteriostatic manner (MBC/MIC ratio  ≤  4). Nanosilver size mediates MRSA inhibition and the cytotoxicity to human cells, being smaller NPs the ones with better antibacterial activity and nontoxic effect on human cells in vitro. As in the case of E. coli, chitosan-based AgNPs harbored high antibacterial activity against S. aureus [110]. The bactericidal effects of AgNPs are not affected by drug-resistant mechanisms of MRSA. AgNPs generate oxidative stress in S. aureus mediated by an increase of ROS, which can cause high levels of oxidized proteins and lipids, DNA fragmentation, and modification in membrane potential [122]. It has been found that a strong interaction between AgNPs and the peptidoglycan layer exists and that AgNPs interact with bacterial cell walls individually or via Ag+ release generating “pits”. Thereafter, AgNPs accumulate and connect more strongly with underlying layers, also releasing Ag+. These phenomena influence the destruction of Gram-positive bacteria more than the damage of Gram-negative bacteria because of the thicker peptidoglycan layer [140].
A MBC of 20 µg/mL was determined for AgNP against S. aureus. When S. aureus cells were exposed to 50 µg AgNPs/mL for 6 h, the cell DNA was condensed to a tension state and could have lost its replicating abilities and when cells were exposed for 12 h, the cell wall was broken, and the cellular contents were released into the environment. The protein content was highly altered as well [141]. Biosynthesized AgNPs showed enhanced quenching activity against S. aureus biofilm and prevented biofilm formation while a synergistic effect of AgNPs with antibiotics (gentamicin, chloramphenicol) in biofilm quenching was effective [142]. The antibacterial activities of penicillin G, amoxicillin, erythromycin, clindamycin, and vancomycin against a test strain of S. aureus increased in the presence of AgNPs and the highest enhancing effects were observed for vancomycin, amoxicillin, and penicillin G [105]. The synergistic activity of some of these antibiotics and AgNPs against MRSA was relatively lower than against Gram-negative bacteria [78]. AMP-AgNPs destroy MRSA isolates [83], and to increase knowledge on this topic, AgNPs combined with gentamicin and oxacillin were tested against an MRSA isolate. The activity of these antibiotics increased in the presence of AgNPs, which has been attributed to the interaction of the AgNPs with hydroxyl and amide groups present in the antibiotic molecules [143]. Considering the clinical importance of S. aureus and the global emergence of MRSA, the inhibitory effects of AgNPs on growth and capsule formation as a virulence factor of this microorganism were investigated. The AgNP-mediated formation of ROS has been detected in S. aureus cells [122]. Consequently, bacterial cell membrane, protein structure, and the intracellular system can be damaged, which could enhance protein leakage by increasing the membrane permeability and decreasing the activity of LDH. The growth and reproduction of AgNP-treated bacteria were quickly inhibited and the pH and temperature conditions did not affect the growth of the treated bacteria [122]. After comparing the activity of AgNPs against S. aureus and E. coli, the lower efficacy of the AgNPs against S. aureus was attributed to differences in the membrane structure. Abbaszadegan et al. [35] found that positively charged NPs were more effective against all tested bacterial species (S. aureus and other Gram-positive and Gram-negative bacteria) than neutral or negatively charged NPs. Stable, well-defined AgNPs, mostly spherical in shape (15 nm) were very active against MRSA isolates and coagulase-negative staphylococci in HIV patients while the in vitro toxicity was scarce and growth inhibition was dose-dependent as usual [144]. According to current research regarding the synergistic effect of AgNPs and antibiotics, it may be expected that combinations of AgNPs and antibiotics (mainly AMP) including the addition of carrier/polymer for a more effective delivery system to the target site of MRSA, can enhance antimicrobial activity and decrease the toxicity of the separate components [137].
Table 4. Research on the activity of AgNPs against S. aureus in vitro.
Table 4. Research on the activity of AgNPs against S. aureus in vitro.
Synthetic Method of the AgNPsAgNP Size (nm)Particle ShapeCappingAntibiotic AddedMIC/MBC (µg/mL)Proposed Mechanism of ActionRef.
Microwave AgNO3 solution, 1000 W, 15 s0.5–24, 1 (avg.)Not indicatedNot indicatedNoMIC: 12.5 (against MSSA and MRSA)Unequal AgNP distribution on the exterior (9.5–33 nm) and interior (5–9 nm) of the bacteria. Reduction of the
PG layer generates destabilization and permeabilization of the bacterial cell membrane and causes osmotic rupture and lysis
[132]
Reduction with NaBH44–20,
13.4 (avg.)
Not indicatedNot indicatedNoMIC > 33 nMFormation of free radicals on the AgNP surface and free radical-induced membrane damage[106]
From plant Gynura procumbens aqueous extract. Then, AgNPs were mixed with fungal chitosan (FCS)10–100Spherical, triangle, and hexagonalChitosan encapsulating AgNPNoMIC: 4.08 ± 0.47No mechanism of action was suggested[133]
Commercial (two sources, a and b)(a) ~100;
(b) 10, 30–40
Not indicatedNoNo10 nm, MIC99: 1800; MBC: 2700; 30–40 nm, MIC99: 10790; MBC: 10790; ~100 nm,
MIC99: 2250; MBC: 8990 (against MRSA)
No mechanism of action was suggested[134]
Reduction with NaBH4, + polyvinyl alcohol (PVA)9Not indicatedNot indicatedNoMIC: 1.95;
MBC: 3.91 (against MRSA and S. aureus)
No mechanism of action was suggested[135]
From leaves of Ricinus communis7.25Spherical or ovalConjugated with STRSTRMIC of STR-AgNPs: 3.12 ± 0.9 (S. aureus)Synergistic effect of AgNPs and STR[138]
Reduction with Na citrate, + polyvinyl alcohol (PVA)17 (avg.)SphericalNot indicatedNoMIC: 2; MBC: 4AgNPs changed the secondary structure (a-helix) of the bacterial cell wall and destroyed its primary structure with the formation of pits, the release of Ag+, PG fragmentation with the release of muramic acid in the medium[140]
Commercial5Not indicatedNot indicatedNoMIC: 5;
MBC: 20
AgNPs over pass cell wall and act on the cell membrane to damage the relative enzymes and interfere with cell metabolism. AgNPs enter bacteria cells and condensed DNA to prevent DNA from replicating and cells from reproducing. Simultaneously, AgNPs continuously act on the cell wall and cell membrane to destroy them[141]
From B. cereus and glucose32SphericalNot indicatedGEN, CHLMIC and MEB were not indicatedSynergistic effect of AgNPs along with antibiotics in biofilm quenching, but the mechanism of action was not suggested[142]
From Streptomyces coelicolor pigments by photo-irradiation within 20 min28–50IrregularNot indicatedGEN, OXAOnly inhibition zones on solid cultures were measured.
GEN: 14 mm; AgNPs + GEN: 22 mm; OXA: 10 mm; AgNPs + OXA: 20 mm
The synergistic activity of AgNPs with both antibiotics was attributed to the interaction of the AgNPs with hydroxyl and amide groups in the antibiotics[143]
From Alysicarpus monilifer leaf extract5–45
15 ± 2 (avg.)
Spherical deriving in nanoprismsNo indicatedNoMIC: 60; MBC: 80 (against coagulase-negative staphylococci):
MIC: 80; MBC: 100 (against MRSA)
AgNPs are capable of affecting the integrity of cell membranes and interacting with disulfide bonds of intracellular enzymes, disturbing metabolic processes and inhibiting the major functions of bacterial cells, including cellular uptake and respiration[144]
From chitosan solution4–18, 6–8 (50% of the AgNPs)Not indicatedNot indicatedNoAgNP-chitosan MIC: 10;
MBC: 10 (S. aureus)
Chitosan-based AgNPs have a dual mechanism of action for antibacterial activity, the bactericidal effect of AgNPs, and the cationic effects of chitosan[110]
From Pseudomonas aeruginosa25–45SphericalNot indicatedNoMIC (pM): 0.4–3.2 (against MSSA –MRSA clinical strains).
3.2 (against S. epidermidis)
MBC (pM): 0.8–3.2 (against MSSA -MRSA clinical strains)
6.2 (S. epidermidis)
Reference to previously reported mechanisms. Effect on the membrane: release of Ag+[80]
CommercialNot indicatedNot indicatedNoNoMIC: 100 (S. aureus)Bacterial protein leakage by increasing the membrane permeability. Formation of ROS that inactivate LDH[122]
Reduction with NaBH4 + Na citrate4Not indicatedCitrate/AMPAMP linked to AgNP (AMP-AgNP)MBC: 1 for AMP-AgNPs, MBC: 4 for AgNPs alone No mechanism of action was proposed[83]
Abbreviations: AMP: ampicillin; avg.: average; GEN: gentamicin; CHL: chloramphenicol; MBC: minimum bactericidal concentration; MIC: minimum inhibitory concentration; MIC99: minimum concentration that inhibits 99% of bacteria; MRSA: methicillin-resistant Staphylococcus aureus; MSSA: methicillin-sensitive Staphylococcus aureus; nM: nanomolar; OXA: oxacillin; PG: peptidoglycan.; pM: picomolar; STR: streptomycin.

7. Toxicity of AgNPs

A growing concern has emerged regarding the biological impacts of NP usage and possible risks to the environment and human health. NPs exhibit an exceptionally increased surface-to-volume ratio due to their ultra-small size. This provides reactivity and, hence, toxicity to these particles. Penetration of NPs across cell barriers is mostly size-dependent. Decreased size exponentially increases surface area resulting in higher levels of oxidation and DNA damaging capabilities [145]. Thus, this passes into living organism cells and can cause several cell lesions [146,147]. The toxicity of AgNPs has been reviewed [148] and it was concluded that cytotoxicity of AgNPs can be considered as dependent on different properties such as size, shape, dose, agglomeration, or aggregation; however, there are not presently adequate studies to obtain a concrete idea of the cytotoxicity of AgNPs or the mechanism behind the toxicity. Binding AgNPs with a coating layer of peptides or other suitable biocompatible molecules can lower toxicity at the time the antibacterial effectivity increases [137,149,150].
Reasonably, in vitro studies were performed before in vivo experiments. Although in vitro data is not a substitute for whole-animal studies, in vitro models can reveal toxicity mechanisms that can serve as a basis for further assessing the potential risk of NP exposure [151].
The mitochondrial function decreased significantly when the immortalized in vitro rat-liver derived cell line (BRL 3A) was used to evaluate the acute toxic effects of AgNPs (15–100 nm) at AgNP doses of 5–50 µg/mL [151]. Exposure of HT 1080 (human fibrosarcoma) and A431 cells (human skin/carcinoma) cells to AgNPs at doses up to 6.25 μg/mL caused apoptosis, oxidative stress, and morphology changes [152]. Concerns about the potential NP cytotoxicity and genotoxicity have increased in the last years leading to intensive studies [153,154,155]. In vitro experiments have evidenced that AgNPs are not only transported into cells and internalized, but also target endosomes and lysosomes [156,157], so that brain astrocytes accumulate AgNPs in a time-, dose-, and temperature-dependent way, likely involving endocytic pathways. AgNPs induce lung fibroblasts, impair the cellular membrane, and cause DNA damage and genotoxicity, chromosome aberration, and apoptosis [148,158,159,160]. Non-agglomerated AgNPs from the cell culture medium were detected as agglomerates within the human mesenchymal stem cells (hMSC) [146]. The silver agglomerates were located in the perinuclear region and the 80 nm AgNPs occurred mainly within endo-lysosomal structures, not in the cell nucleus, endoplasmic reticulum, or Golgi complex. Damage to hMSC vitality at concentrations of 10 µg/mL was evident when working on hMSC cultures doped with AgNPs (47 nm) [161]. ROS generation and oxidative stress play a crucial role in this context. AgNPs may induce geno- and cytotoxic effects in hMSC at high exposure concentrations although subtoxic levels may activate hMSC [161]. Exposure to AgNPs of human alveolar basal epithelial cells (A549) produced ROS generation and reductions in cell viability and mitochondrial membrane potential [153,162]. AgNPs showed more toxicity in A549 cells than in L132 normal human lung cells, which had no significant membrane leakage. Toxicity is dose-dependent and AgNPs target cancer cells rather than normal cells [162]. PVP-AgNPs do not have toxic effects on human pulmonary host cells (A549) at levels therapeutic to A. baumannii infection, being the IC50 = 130 µM [62]. Cytotoxicity was assessed according to ISO 10993-533 by monitoring the neutral red uptake assay using mouse fibroblasts NCTC 929, and tumor cells HeLa and HepG2 (100 μL; 1 × 105 cells/mL) seeded into 96-well plates and left to adhere during 24 h [75]. Cells were exposed to 10-nm AgNPs previously dispersed and serially diluted at concentrations from 10.0 to 0.156 μg/mL. The bactericidal AgNP levels were non-cytotoxic in NCTC 929, HepG2, and HeLa tumor cells. Low toxicity to the cell lines HepG2 and HeLa was observed at 5.0 μg/mL while cytotoxicity was evident at 10 μg/mL [75]. AgNPs (20 nm) killed S. aureus but were non-toxic to HeLa cells [134]. Wypij et al. [97] have reported a high cytotoxic effect (IC50 about 4 μg/mL) of biogenic AgNPs (5–20 nm) in vitro using the mouse 3T3 fibroblasts and HeLa cell line. Cytotoxicity inconsistencies with results from previous research were possibly due to experimental differences [97]. The combined use of AgNPs and antibiotics has made it possible to reduce the dosage of both antimicrobials and their toxicity toward mouse fibroblasts and HeLa cells [97]. On the other hand, the AgNP concentration required for 50% reduction in viability in HeLa cells was 200 μg/mL using NPs of 45 nm size [163] or 100 μg/mL using NPs with a size of 62 nm [164]. Składanowski et al. [155] also reported AgNPs low cytotoxicity (IC50 = 64.5 μg/mL) against mouse fibroblasts (L929 cell line). The combination of some antibiotics with AgNPs at their MIC values decreased cell viability in comparison with untreated cells. The highest cytotoxic effect was detected for ampicillin/sulbactam, cefazolin, meropenem, and chloramphenicol combined with AgNPs, which was attributed to the additive cytotoxicity of the antibiotics and AgNPs [95].
Barbasz et al. [165] studied the toxicity of three AgNP types [uncapped (negatively charged), citrate-capped (negatively charged) and cysteamine-capped (positively charged)] with sizes of 11–14 nm towards histiocytic lymphoma (U-937) and human promyelocytic cells (HL-60) and found that uncapped negatively charged AgNPs exhibited the highest toxicity towards the tumor cell line. They concluded that the AgNP cytotoxicity mechanism is a combination of effects coming from the NP surface charge, released silver ions and the biological activity of stabilizing agent molecules, and that their results confirmed that disruption in mitochondrial functions and generation of oxidative stress are the main reasons of cell death.
Low doses of antibiotics and AgNPs slightly decrease cell viability (to 90–95%) in comparison with the control cells, depending on the antibiotics used [97], which suggests that AgNPs in combination with antibiotics have much potential for application as antimicrobial agents. AgNPs at concentrations ≤ 30 µg/mL did not display cytotoxic effects to human cells, blood, or to environmentally important organisms [166,167,168,169]. The results of extensive hemocompatibility tests based on plasma concentrations of activation markers, cell surface markers, and blood cell alterations reflected good hemocompatibility of AgNPs (12 nm) at concentrations up to 3 µg/mL. No biological relevant alterations during blood contact were observed but 30 µg AgNP/mL induced activation of various hematologic parameters and this dose should not be used in vivo [169].
An in vivo study on 60 healthy volunteers orally exposed to commercial AgNPs in a prospective, placebo-controlled, single-blind, dose-monitored, and cross-over design did not show clinically important changes in metabolic, urine, hematologic, physical findings, or imaging morphology after 14 days of exposure to 10 µg/mL (5–10 nm size) and 32 µg/mL (25–40 nm size) of AgNPs. Thus, exposure to low AgNP doses has no adverse or toxic effects on humans according to that study [170]. However, the results from different studies related to the toxic effects of AgNPs using experimental animals disagree and conclusions are controversial [124,171,172,173,174,175,176,177,178,179]. Generally, oral exposures to AgNP caused weight loss, inflammatory and immune responses, hepatic alterations, increased levels of neurotransmitters, and changed blood values in animal model experiments at concentrations of units or tens of mg/kg [171,172,173,174,176]. Kim et al. [106] found that 28 days of repeated oral doses of commercial AgNPs (60 nm) to Sprague–Dawley rats induced liver toxicity, affected coagulation of peripheral blood, and had a dose-dependent deposition of AgNPs in the blood, stomach, brain, liver, kidneys, lungs, and testes of the rats indicating that the AgNPs were systemically distributed in the tissues. A conclusion was that exposure to > 300 mg of AgNPs may result in only slight liver damage. A study on the oral toxicity of AgNPs (56 nm) over a 90-day period in F344 rats concluded that the target organ for the AgNPs was the liver in male and female rats [173]. A NOAEL (no observable adverse effect level) of 30 mg/kg of body weight/day and a LOAEL (lowest observable adverse effect level) of 125 mg/kg of body weight/day was suggested. Nevertheless, the LOAEL expressed by increased cytokine concentration was 0.5 mg/kg of body weight/day in mice following 28-day oral AgNP exposure [176]. Dermal toxicity studies showed that exposure to > 0.1 mg AgNPs/kg results in slight spleen, liver, and skin damage in guinea pigs, thus denoting that this administration way supposes more toxicity than oral or inhalation ways [180]. After subcutaneous injections of AgNPs at 62.8 mg/kg in rats, the NPs were translocated to the blood circulation and distributed to the kidney, liver, spleen, brain, and lung. Moreover, AgNPs caused blood–brain barrier destruction and neuronal degeneration [181,182]. However, doses of 62.8 mg/kg in the case of subcutaneous injection or 300 mg/kg in the case of oral administration are very high and they do not need to be administrated to treat bacterial infections, especially when AgNPs are combined with antibiotics [95].
To clarify the toxic effects of AgNPs, a genotoxicity test, oral and dermal toxicity test, skin toxicity test, and eye toxicity test was conducted according to the OECD test guidelines and GLP [178]. AgNPs (10 nm) did not show a severe toxic effect on microorganisms, mammalian cell lines, or target animal organs. Notwithstanding, certain concentrations of AgNPs induced cytotoxicity in microorganisms and mammalian cell lines. Abnormal signs or mortality following the acute oral or dermal exposure of rats at a dose of 2000 mg/kg were not found, making the LD50 of AgNPs for Sprague Dawley rats above 2000 mg/kg. Other authors agree with these results [183]. In the dermal irritation and corrosion test, the AgNPs did not generate abnormal clinical signs or mortality in New Zealand White Rabbits and did not induce any erythema, eschar, or edema formation during the experimental period. In the skin sensitization test, a weak skin sensitization effect was found in one guinea pig (5%), which showed discrete or patchy erythema induced by AgNPs [178].
Changes in the acute toxicity of intraperitoneally administered AgNPs (10, 60 and 100 nm) in BALB/c mice (0.2 mg/mouse) have been observed [184]. The smaller AgNPs exhibited more toxicity than the larger ones. After 6 h of administration congestion, vacuolation, single cell necrosis, focal necrosis in the liver, congestion in the spleen and apoptosis in the thymus cortex was observed. These results agree with other studies on the acute toxicity of citrate-capped and PVP-capped AgNPs in mice but using intravenous injection (a single dose of 10 mg/kg). After 24 h, the highest silver concentrations occurred in the spleen and liver, followed by the lung, kidney, and brain [185].
Unfortunately, the issue of the relevant AgNP dose required for system or local elimination of infection is not addressed yet. Similarly, pharmacological and pharmacokinetic data on AgNPs have not been described so far. Therefore, prediction of the therapeutic AgNP doses and their adverse effects is very difficult at this time. This is still an open field, which requires further exploration to determine if AgNPs combined with antibiotics may be effective for the local and systematic therapy of infectious diseases without showing adverse effects.
Concerning toxicity in humans, there is little information on assays performed in volunteer patients [170,186]. NPs with size ≤ 35 nm can penetrate and cross the blood–brain barrier, particles with size ≤ 40 nm may enter nuclei of cells, and those with size 1–100 nm can cross the biological membrane and can be translocated inside cell organelles or entities such as the mitochondria, lysosome, nucleus, and others [187]. AgNPs can induce size-dependent cytotoxicity in human lung cells due to the substantial release of Ag in the cell [25]. Silver-coated wound dressing tested in human burns patients produced reversible hepatotoxicity and argyria-like discoloration of the treated area of skin, elevated plasma and urine silver concentrations, and increased liver enzymes. AgNP doses in the 5–10 μg/mL range proved toxic in eukaryotic cells. If effective antimicrobial AgNP doses were higher than cytotoxic levels, its practical use in humans would be problematic [188]. The pharmacokinetic and pharmacodynamic characteristics of the NPs, including AgNPs, have been evaluated but in-depth knowledge is needed [189]. The key issue to overcoming toxicity problems in AgNP treatments is finding silver nanocomposites capped with antibiotics that can act as efficient antimicrobial systems against antibiotic-resistant bacteria without toxicity to human tissues. The objective must be optimizing the ratio of maximal antibacterial activity/null or minimal toxicity to human organs. Thus, studies aimed at standardizing the optimal size, shape, purity, stability, capping agents, antibiotic combination, and doses of AgNPs to permit clinical usage on critically infected patients with minimal or null side effects are needed.

8. Conclusions and Future Perspectives

The abuse of antibiotics worldwide has contributed to the development of MDR infectious bacteria. This is a real problem for public health because nosocomial infections are very difficult to cure. There is an urgent demand for new treatments to counteract the increased morbidity/mortality rates and treatment costs. Nanotechnology has become a new tool to fight against MDR microorganisms. Metal NPs may be used in medicine to combat the infections caused by these bacteria. Particularly, AgNPs have been applied to this objective and have been shown to be very effective against A. baumannii, P. aeruginosa, Enterobacteriaceae, and MDR S. aureus. The mechanism of action of AgNPs involves interactions at various levels. They change the plasma membrane permeability causing the release of the intracellular content and leading to cell death. They stop DNA replication, inhibit the expression of ribosomal subunits and inactivate proteins/enzymes. They change the normal function of membrane-bound respiratory enzymes and lead to the formation of ROS with oxidative deterioration of cell content. AgNP toxicity is dependent on dose, size, shape, and capping/coating agents among other factors. The addition of suitable antibiotics increases the antibacterial activity and decreases toxicity due to synergistic effects. Low sizes (5–30 nm) are more effective against bacteria than larger diameters. The shape and the electrical charge influence its activity. AgNP toxicity to animal cells has been studied mainly in vitro on cell cultures but also in vivo. The smallest AgNPs display the highest toxicity to animal cells while large sizes are less toxic. Therefore, the use of small AgNPs must be exerted with great care. Different researchers do not always agree on toxicological aspects. The administration way influences the toxicity in animals. The doses should be as low as possible but the relevant dose required for the elimination of infection is not addressed. Thus, more studies are needed to establish the best compromise between toxicity and therapeutic effects by finding the best balance encompassing dose/size/shape/charge and coating with substances such as suitable antibiotics and/or capping compounds that can improve AgNP effectiveness. Detoxification of affected organs is a very important issue to be addressed. Another perspective to be considered in future is the study of possible bacterial mechanisms of resistance against NPs. Owing to the urgent action demanded by the WHO against MDR bacteria such research is of paramount importance. The answer to the research question of this review is yes, AgNPs can be useful in treatments but more research (mainly in toxicology) is needed before AgNP-based therapies may be approved for use in clinical medicine by the Food and Drug Administration or the European Medicine Agency.

Author Contributions

Conceptualization, E.M.M.; methodology, E.M.M.; validation, M.J.; formal analysis, E.M.M.; investigation, E.M.M.; resources, E.M.M. and M.J.; writing—original draft preparation, E.M.M.; writing—review and editing, E.M.M. and M.J.; supervision, E.M.M. and M.J.; project administration, M.J.; funding acquisition, M.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the European Regional Development Fund (ERDF) and the Ministry of Economy and Competitiveness, the Ministry of Science, Innovation and Universities (Spanish Government) through Project number RTI2018-097593-B-C22.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. WHO. Antimicrobial Resistance in the European Union and the World. 2012. Available online: https://www.who.int/director-general/speeches/detail/antimicrobial-resistance-in-the-european-union-and-the-world (accessed on 15 June 2022).
  2. WHO. Antimicrobial Resistance Global Report on Surveillance; World Health Organization: Geneva, Switzerland, 2014; p. 2014257. Available online: https://apps.who.int/iris/bitstream/handle/10665/112642/9789241564748_eng.pdf?sequence=1&isAllowed=y (accessed on 16 June 2022).
  3. Prestinaci, F.; Pezzotti, P.; Pantosti, A. Antimicrobial resistance: A global multifaceted phenomenon. Pathog. Glob. Health. 2015, 109, 309–318. [Google Scholar] [CrossRef] [PubMed]
  4. Willyard, C. The drug-resistant bacteria that pose the greatest health threats. Nature 2017, 543, 15. [Google Scholar] [CrossRef] [PubMed]
  5. Horikoshi, S.; Serpone, N. Introduction to nanoparticles. In Microwaves in Nanoparticle Synthesis: Fundamentals and Applications; Horikoshi, S., Serpone, N., Eds.; Wiley-VCH Verlag GmbH and Co. KGaA: Weinheim, Germany, 2013; pp. 1–24. [Google Scholar]
  6. Singh, L.P.; Bhattacharyya, S.K.; Kumar, R.; Mishra, G.; Sharma, U.; Singh, G.; Ahalawat, S. Sol-gel processing of silica nanoparticles and their applications. Adv. Colloid Interface Sci. 2014, 214, 17–37. [Google Scholar] [CrossRef] [PubMed]
  7. Vichery, C.; Nedelec, J.M. Bioactive glass nanoparticles: From synthesis to materials design for biomedical applications. Materials 2016, 9, 288. [Google Scholar] [CrossRef]
  8. Mitchell, M.J.; Billingsley, M.M.; Haley, R.M.; Wechsler, M.E.; Peppas, N.M.; Langer, R. Engineering precision nanoparticles for drug delivery. Nat. Rev. Drug Discov. 2021, 20, 101–124. [Google Scholar] [CrossRef]
  9. Zain, M.; Yasmeen, H.; Yadav, S.S.; Amir, S.; Bilal, M.; Shahid, A.; Khurshid, M. Applications of nanotechnology in biological systems and medicine. In Micro and Nano Technologies, Nanotechnology for Hematology, Blood Transfusion, and Artificial Blood; Chapter 10; Denizli, A., Nguyen, T.A., Rajan, M., Alam, M.F., Rahman, K., Eds.; Elsevier: Amsterdam, The Netherlands, 2022; pp. 215–235. [Google Scholar] [CrossRef]
  10. Palao-Suay, R.; Gómez-Mascaraque, L.G.; Aguilar, M.R.; Vázquez-Lasa, B.; Román, J.S. Self-assembling polymer systems for advanced treatment of cancer and inflammation. Prog. Polym. Sci. 2016, 53, 207–248. [Google Scholar] [CrossRef]
  11. Gao, W.; Chen, Y.; Zhang, Y.; Zhang, Q.; Zhang, L. Nanoparticle-based local antimicrobial drug delivery. Adv. Drug Deliv. Rev. 2018, 127, 45–57. [Google Scholar] [CrossRef]
  12. Lam, S.J.; Wong, E.H.H.; Boyer, C.; Qiao, G.G. Antimicrobial polymeric nanoparticles. Progr. Polym. Sci. 2018, 76, 40–64. [Google Scholar] [CrossRef]
  13. Rai, M.K.; Ingle, A.P.; Pandit, R.; Paralikar, P.; Gupta, I.; Chaud, M.V.; Dos Santos, C.A. Broadening the spectrum of small-molecule antibacterials by metallic nanoparticles to overcome microbial resistance. Int. J. Pharm. 2017, 532, 139–148. [Google Scholar] [CrossRef]
  14. Hoseinzadeh, E.; Makhdoumi, P.; Taha, P.; Hossini, H.; Stelling, J.; Kamal, M.A.; Ashraf, G.M. A review on nano-antimicrobials: Metal nanoparticles, methods, and mechanisms. Curr. Drug Metab. 2016, 18, 120–128. [Google Scholar] [CrossRef]
  15. Tripathi, N.; Goshisht, M.K. Recent advances and mechanistic insights into antibacterial activity, antibiofilm activity, and cytotoxicity of silver nanoparticles. ACS Appl. Bio Mater. 2022, 5, 1391–1463. [Google Scholar] [CrossRef]
  16. Merrifield, R.C.; Stephan, C.; Lead, J.R. Single-particle inductively coupled plasma mass spectroscopy analysis of size and number concentration in mixtures of monometallic and bimetallic (core-shell) nanoparticles. Talanta 2017, 162, 130–134. [Google Scholar] [CrossRef]
  17. 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]
  18. Kaushik, A.; Jayant, R.D.; Bhardwaj, V.; Nair, M. Personalized nanomedicine for CNS diseases. Drug Discov. Today 2018, 23, 1007–1015. [Google Scholar] [CrossRef]
  19. Jiang, W.; Rutherford, D.; Vuong, T.; Liu, H. Nanomaterials for treating cardiovascular diseases: A review. Bioact. Mat. 2017, 2, 185–198. [Google Scholar] [CrossRef]
  20. Reimondez-Troitiño, S.; Csaba, N.; Alonso, M.J.; de la Fuente, M. Nanotherapies for the treatment of ocular diseases. Eur. J. Pharm. Biopharm. 2015, 95, 279–293. [Google Scholar] [CrossRef]
  21. Zheng, X.; Zhang, C.; Guo, Q.; Wan, X.; Zhang, Q. Dual-functional nanoparticles for precise drug delivery to Alzheimer’s disease lesions: Targeting mechanisms, pharmacodynamics and safety. Int. J. Pharm. 2017, 525, 237–248. [Google Scholar] [CrossRef]
  22. Wong, C.Y.; Al-Salami, H.; Dass, C.R. Potential of insulin nanoparticle formulations for oral delivery and diabetes treatment. J. Control. Release 2017, 264, 247–275. [Google Scholar] [CrossRef]
  23. Zang, X.; Zhao, X.; Hu, H.; Qiao, M.; Chen, D. Nanoparticles for tumor immunotherapy. Review. Eur. J. Pharm. Biopharm. 2017, 115, 243–256. [Google Scholar] [CrossRef]
  24. Li, R.; Liu, B.; Gao, J. The application of nanoparticles in diagnosis and theranostics of gastric cancer. Cancer Lett. 2017, 386, 123–130. [Google Scholar] [CrossRef]
  25. Chandrakala, V.; Aruna, V.; Angajala, G. Review on metal nanoparticles as nanocarriers: Current challenges and perspectives in drug delivery systems. Emergent Mater. 2022, 1–23. [Google Scholar] [CrossRef]
  26. Singh, R.; Smitha, M.S.; Singh, S.P. The role of nanotechnology in combating multidrug resistant bacteria. J. Nanosci. Nanotechnol. 2014, 14, 4745–4756. [Google Scholar] [CrossRef]
  27. Zazo, H.; Colino, C.I.; Lanao, J.M. Current applications of nanoparticles in infectious diseases. J. Control. Release 2016, 224, 86–102. [Google Scholar] [CrossRef]
  28. Ma, Z.; Garrido-Maestu, A.; Jeong, K.C. Application, mode of action, and in vivo activity of chitosan and its micro and nanoparticles as antimicrobial agents: A review. Carbohydr. Polym. 2017, 176, 257–265. [Google Scholar] [CrossRef]
  29. Mba, E.I.; Nweze, E.I. Nanoparticles as therapeutic options for treating multidrug-resistant bacteria: Research progress, challenges, and prospects. World J. Microbiol. Biotechnol. 2021, 37, 108. [Google Scholar] [CrossRef]
  30. Cheeseman, S.; Christofferson, A.J.; Kariuki, R.; Cozzolino, D.; Daeneke, T.; Crawford, R.J.; Truong, V.K.; Chapman, J.; Elbourne, A. Antimicrobial metal nanomaterials: From passive to stimuli-activated applications. Adv. Sci. 2020, 7, 1902913. [Google Scholar] [CrossRef]
  31. Slavin, Y.N.; Asnis, J.; Häfeli, U.O.; Bach, H. Metal nanoparticles: Understanding the mechanisms behind antibacterial activity. J. Nanobiotechnol. 2017, 15, 65. [Google Scholar] [CrossRef]
  32. Hochvaldová, L.; Večeřová, R.; Kolář, M.; Prucek, R.; Kvítek, L.; Lapčík, L.; Panáček, A. Antibacterial nanomaterials: Upcoming hope to overcome antibiotic resistance crisis. Nanotechnol. Rev. 2022, 11, 1115–1142. [Google Scholar] [CrossRef]
  33. Darabdhara, G.; Boruah, P.K.; Hussain, N.; Borthakur, P.; Sharma, B.; Sengupta, P.; Das, M.R. Magnetic nanoparticles towards efficient adsorption of gram positive and gram negative bacteria: An investigation of adsorption parameters and interaction mechanism. Colloids Surf. A Physicochem. Eng. Asp. 2017, 516, 161–170. [Google Scholar] [CrossRef]
  34. Lee, K.J.; Browning, L.M.; Nallathamby, P.D.; Xu, X.H.N. Study of charge-dependent transport and toxicity of peptide-functionalized silver nanoparticles using zebrafish embryos and single nanoparticle plasmonic spectroscopy. Chem. Res. Toxicol. 2013, 26, 904–917. [Google Scholar] [CrossRef]
  35. Abbaszadegan, A.; Ghahramani, Y.; Gholami, B.; Hemmateenejad, A.; Dorostkar, S.; Nabavizadeh, M.; Sharghi, H. The effect of charge at the surface of silver nanoparticles on antimicrobial activity against gram-positive and gram-negative bacteria: A preliminary study. J. Nanomater. 2015, 2015, 720654. [Google Scholar] [CrossRef]
  36. Kawish, M.; Ullah, F.; Ali, H.S.; Saifullah, S.; Ali, I.; Rehman, J.U.; Imran, M. Bactericidal potentials of silver nanoparticles: Novel aspects against multidrug resistance bacteria Chapter 10. In Micro and Nano Technologies, Metal Nanoparticles for Drug Delivery and Diagnostic Applications; Shah, M.R., Imran, M., Ullah, S., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 175–188. [Google Scholar] [CrossRef]
  37. Rai, M.K.; Deshmukh, S.D.; Ingle, A.P.; Gade, A.K. Silver nanoparticles: The powerful nanoweapon against multidrug-resistant bacteria. J. Appl. Microbiol. 2012, 112, 841–852. [Google Scholar] [CrossRef] [PubMed]
  38. Rai, M.; Kon, K.; Ingle, A.; Duran, N.; Galdiero, S.; Galdiero, M. Broad-spectrum bioactivities of silver nanoparticles: The emerging trends and future prospects. Appl. Microbiol. Biotechnol. 2014, 98, 1951–1961. [Google Scholar] [CrossRef] [PubMed]
  39. Ahmed, K.B.A.; Raman, T.; Veerappan, A. Future prospects of antibacterial metal nanoparticles as enzyme inhibitor. Mater. Sci. Eng. C 2016, 68, 939–947. [Google Scholar] [CrossRef]
  40. Wang, L.; Hu, C.; Shao, L. The antimicrobial activity of nanoparticles: Present situation and prospects for the future. Int. J. Nanomed. 2017, 12, 1227–1249. [Google Scholar] [CrossRef]
  41. Häffner, S.M.; Malmsten, M. Membrane interactions and antimicrobial effects of inorganic nanoparticles. Adv. Colloid Interface Sci. 2017, 248, 105–128. [Google Scholar] [CrossRef]
  42. Prabhu, S.; Poulose, E.K. Silver nanoparticles: Mechanism of antimicrobial action, synthesis, medical applications and toxicity effects. Int. Nano Lett. 2012, 2, 32. [Google Scholar] [CrossRef]
  43. Hsueh, Y.H.; Lin, K.S.; Ke, W.J.; Hsieh, C.T.; Chiang, C.L.; Tzou, D.Y.; Liu, S.T. The antimicrobial properties of silver nanoparticles in Bacillus subtilis are mediated by released Ag+ ions. PLoS ONE 2015, 10, e0144306. [Google Scholar] [CrossRef] [Green Version]
  44. Tamayo, L.A.; Zapata, P.A.; Vejar, N.D.; Azocar, M.I.; Gulppi, M.A.; Zhou, X.; Thompson, G.E.; Rabagliati, F.M.; Paez, M.A. Release of silver and copper nanoparticles from polyethylene nanocomposites and their penetration into Listeria Monocytogenes. Mater. Sci. Eng. 2014, 40, 24–31. [Google Scholar] [CrossRef]
  45. Dayma, P.B.; Mangrola, A.V.; Suriyaraj, S.P.; Dudhagara, P.; Patel, R.K. Synthesis of bio-silver nanoparticles using desert isolated Streptomyces intermedius and its antimicrobial activity. J. Pharm. Chem. Biol. Sci. 2019, 7, 94–101. [Google Scholar]
  46. Yuan, Y.G.; Peng, Q.L.; Gurunathan, S. Effects of silver nanoparticles on multiple drug-resistant strains of Staphylococcus aureus and Pseudomonas aeruginosa from mastitis-infected goats: An alternative approach for antimicrobial therapy. Int. J. Mol. Sci. 2017, 18, 569. [Google Scholar] [CrossRef]
  47. Ahmed, S.; Ahmad, M.; Swami, B.L.; Ikram, S. A review on plants extract mediated synthesis of silver nanoparticles for antimicrobial applications: A green expertise. J. Adv. Res. 2016, 7, 17–28. [Google Scholar] [CrossRef]
  48. Chang, T.Y.; Chen, C.C.; Cheng, K.M.; Chin, C.Y.; Chen, Y.H.; Chen, X.A.; Sun, J.R.; Young, J.J.; Chiueh, T.S. Trimethyl chitosan-capped silver nanoparticles with positive surface charge: Their catalytic activity and antibacterial spectrum including multidrug-resistant strains of Acinetobacter baumannii. Colloids Surf. B Biointerfaces 2017, 155, 61–70. [Google Scholar] [CrossRef]
  49. Siddiqi, K.S.; Hussein, A.; Rao, R.A.K. A review on biosynthesis of silver nanoparticles and their biocidal properties. J. Nanobiotechnol. 2018, 16, 14. [Google Scholar] [CrossRef]
  50. Chopra, H.; Bibi, S.; Singh, I.; Hasan, M.M.; Khan, M.S.; Yousafi, Q.; Baig, A.A.; Rahman, M.M.; Islam, F.; Emran, T.B.; et al. Green metallic nanoparticles: Biosynthesis to applications. Front. Bioeng. Biotechnol. 2022, 10, 874742. [Google Scholar] [CrossRef]
  51. Wacker, M.G.; Proykova, A.; Mendes, G.; Santos, L. Dealing with nanosafety around the globe regulation vs. innovation. Int. J. Pharma. 2016, 509, 95–106. [Google Scholar] [CrossRef]
  52. García-Quintanilla, M.; Pulido, M.R.; López-Rojas, R.; Pachón, J.; McConnell, M.J. Emerging therapies for multidrug resistant Acinetobacter baumannii. Trends Microbiol. 2013, 21, 157–163. [Google Scholar] [CrossRef]
  53. Srikanth, D.; Joshi, S.V.; Shaik, M.G.; Pawar, G.; Bujji, S.; Kanchupalli, V.; Chopra, S.; Nanduri, S. A comprehensive review on potential therapeutic inhibitors of nosocomial Acinetobacter baumannii superbugs. Bioorg. Chem. 2022, 124, 105849. [Google Scholar] [CrossRef]
  54. Taccone, F.S.; Rodríguez-Villalobos, H.; De Backer, D.; De Moor, V.; Deviere, J.; Vincent, J.L.; Jacobs, F. Successful treatment of septic shock due to pan-resistant Acinetobacter baumannii using combined antimicrobial therapy including tigecycline. Eur. J. Clin. Microbiol. Infect. Dis. 2006, 25, 257–260. [Google Scholar] [CrossRef]
  55. Maragakis, L.L.; Perl, T.M. Acinetobacter baumannii: Epidemiology, antimicrobial resistance, and treatment options. Clin. Infect. Dis. 2008, 46, 1254–1263. [Google Scholar] [CrossRef]
  56. Valencia, R.; Arroyo, L.A.; Conde, M.; Aldana, J.M.; Torres, M.J.; Fernández-Cuenca, F.; Garnacho-Montero, J.; Cisneros, J.M.; Ortíz, C.; Pachón, J.; et al. Nosocomial outbreak of infection with pandrug-resistant Acinetobacter baumannii in a tertiary care university hospital. Infect. Control Hosp. Epidemiol. 2009, 30, 57–263. [Google Scholar] [CrossRef]
  57. Singh, R.; Wagh, P.; Wadhwani, S.; Gaidhani, S.; Kumbhar, A.; Bellare, J.; Chopade, B.A. Synthesis, optimization, and characterization of silver nanoparticles from Acinetobacter calcoaceticus and their enhanced antibacterial activity when combined with antibiotics. Int. J. Nanomed. 2013, 8, 4277–4290. [Google Scholar]
  58. Ghafoori, S.M.; Entezari, M.; Taghva, M.E.A.; Tayebi, Z. Biosynthesis and evaluation of the characteristics of silver nanoparticles using Cassia fistula fruit aqueous extract and its antibacterial activity. Adv. Nat. Sci. Nanosci. Nanotechnol. 2017, 8, 045019. [Google Scholar] [CrossRef]
  59. Baghayeri, M.; Mahdavi, B.; Hosseinpor-Mohsen Abadi, Z.; Farhadi, S. Green synthesis of silver nanoparticles using water extract of Salvia leriifolia: Antibacterial studies and applications as catalysts in the electrochemical detection of nitrite. Appl. Organomet. Chem. 2018, 32, e4057. [Google Scholar] [CrossRef]
  60. Tiwari, V.; Khokar, M.K.; Tiwari, M.; Barala, S.; Kumar, M. Anti-bacterial activity of polyvinyl pyrrolidone capped silver nanoparticles on the carbapenem resistant strain of Acinetobacter baumannii. J. Nanomed. Nanotechnol. 2014, 5, 1000246. [Google Scholar] [CrossRef]
  61. Tiwari, M.; Raghav, R.; Tiwari, V. Comparative anti-bacterial activity of differently capped silver nanomaterial on the carbapenem sensitive and resistant strains of Acinetobacter baumannii. J. Nanomed. Nanotechnol. 2015, 6, 1000314. [Google Scholar] [CrossRef]
  62. Tiwari, V.; Tiwari, M.; Solanki, M. Polyvinylpyrrolidone-capped silver nanoparticle inhibits infection of carbapenem-resistant strain of Acinetobacter baumannii in the human pulmonary epithelial cell. Front. Immunol. 2017, 8, 973. [Google Scholar] [CrossRef]
  63. Li, X.; Gui, R.; Li, J.; Huang, R.; Shang, Y.; Zhao, Q.; Liu, H.; Jiang, H.; Shang, X.; Wu, X.; et al. Novel multifunctional silver nanocomposite serves as a resistance-reversal agent to synergistically combat carbapenem-resistant Acinetobacter baumannii. ACS Appl. Mater. Interfaces 2021, 13, 30434–30457. [Google Scholar] [CrossRef]
  64. Wan, G.; Ruan, L.; Yin, Y.; Yang, T.; Ge, M.; Cheng, X. Effects of silver nanoparticles in combination with antibiotics on the resistant bacteria Acinetobacter baumannii. Int. J. Nanomed. 2016, 11, 3789–3800. [Google Scholar] [CrossRef]
  65. Łysakowska, M.E.; Ciebiada-Adamiec, A.; Klimek, L.; Sienkiewicz, M. The activity of silver nanoparticles (Axonnite) on clinical and environmental strains of Acinetobacter spp. Burns 2015, 41, 364–371. [Google Scholar] [CrossRef]
  66. Silva Santos, K.; Mendonça Barbosa, A.; da Costa, L.P.; Santos Pinheiro, M.; Prior Pinto Oliveira, M.B.; Pad, F.F. Silver nanocomposite biosynthesis: Antibacterial activity against multidrug-resistant strains of Pseudomonas aeruginosa and Acinetobacter baumannii. Molecules 2016, 21, 1255. [Google Scholar] [CrossRef]
  67. Ghosh, S.; Patil, S.; Ahire, M.; Kitture, R.; Kale, S.; Pardesi Swaranjit, K.; Cameotra, S.; Bellare, J.; Dhavale, D.D.; Jabgunde, A.; et al. Synthesis of silver nanoparticles using Dioscorea bulbifera tuber extract and evaluation of its synergistic potential in combination with antimicrobial agents. Int. J. Nanomed. 2012, 7, 483–496. [Google Scholar]
  68. Maunders, E.; Welch, M. Matrix exopolysaccharides; the sticky side of biofilm formation. FEMS Microbiol. Lett. 2017, 364, fnx120. [Google Scholar] [CrossRef] [PubMed]
  69. Singh, R.; Vora, J.; Nadhe, S.; Wadhwani, S.A.; Shedbalkar, U.U.; Chopade, B.A. Antibacterial activities of bacteriagenic silver nanoparticles against nosocomial Acinetobacter baumannii. J. Nanosci. Nanotechnol. 2018, 18, 3806–3815. [Google Scholar] [CrossRef] [PubMed]
  70. Shi, T.; Wei, Q.; Wang, Z.; Zhang, G.; Sun, X.; He, Q.Y. Photocatalytic protein damage by silver nanoparticles circumvents bacterial stress response and multidrug resistance. mSphere 2019, 4, e00175-19. [Google Scholar] [CrossRef] [PubMed]
  71. Porcaro, F.; Carlini, L.; Ugolini, A.; Visaggio, D.; Visca, P.; Fratoddi, I.; Venditti, I.; Meneghini, C.; Simonelli, L.; Marini, C.; et al. Synthesis and structural characterization of silver nanoparticles stabilized with 3-mercapto-1-propansulfonate and 1-thioglucose mixed thiols for antibacterial applications. Materials 2016, 9, 1028. [Google Scholar] [CrossRef] [PubMed]
  72. Mulani, M.S.; Kamble, E.E.; Kumkar, S.N.; Tawre, M.S.; Pardesi, K.R. Emerging strategies to combat ESKAPE pathogens in the era of antimicrobial resistance: A review. Front. Microbiol. 2019, 10, 539. [Google Scholar] [CrossRef] [PubMed]
  73. Chen, M.; Yu, X.; Huo, Q.; Yuan, Q.; Li, X.; Xu, C.; Bao, H. Biomedical potentialities of silver nanoparticles for clinical multiple drug-resistant Acinetobacter baumannii. J. Nanomater. 2019, 2019, 3754018. [Google Scholar] [CrossRef]
  74. Hetta, H.F.; Al-Kadmy, I.M.S.; Khazaal, S.S.; Abbas, S.; Suhail, A.; El-Mokhtar, M.A.; Abd Ellah, N.H.; Ahmed, E.A.; Abd-ellatief, R.B.; El-Masry, E.A.; et al. Antibiofilm and antivirulence potential of silver nanoparticles against multidrug-resistant Acinetobacter baumannii. Sci. Rep. 2021, 11, 10751. [Google Scholar] [CrossRef]
  75. 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]
  76. Kerr, K.G.; Snelling, A.M. Pseudomonas aeruginosa: A formidable and ever-present adversary. J. Hospital Infect. 2009, 73, 338–344. [Google Scholar] [CrossRef]
  77. Morones, J.R.; Elechiguerra, J.L.; Camacho, A.; Holt, K.; Kouri, J.B.; Ramírez, J.T.; Yacaman, M.J. The bactericidal effect of silver nanoparticles. Nanotechnology 2005, 16, 2346–2353. [Google Scholar] [CrossRef]
  78. Birla, S.S.; Tiwari, V.V.; Gade, A.K.; Ingle, A.P.; Yadav, A.P.; Rai, M.K. Fabrication of silver nanoparticles by Phoma glomerata and its combined effect against Escherichia coli, Pseudomonas aeruginosa and Staphylococcus aureus. Lett. Appl. Microbiol. 2009, 48, 173–179. [Google Scholar] [CrossRef]
  79. Oza, G.; Pandey, S.; Shah, R.; Sharon, M. Extracellular fabrication of silver nanoparticles using Pseudomonas aeruginosa and its antimicrobial assay. Adv. Appl. Sci. Res. 2012, 3, 1776–1783. [Google Scholar]
  80. Quinteros, M.A.; Aiassa Martínez, I.M.; Dalmasso, P.R.; Páez, P.L. Silver nanoparticles: Biosynthesis using an ATCC reference strain of Pseudomonas aeruginosa and activity as broad spectrum clinical antibacterial agents. Int. J. Biomater. 2016, 2016, 5971047. [Google Scholar] [CrossRef]
  81. Singh, P.; Mijakovic, I. Antibacterial effect of silver nanoparticles is stronger if the production host and the targeted pathogen are closely related. Biomedicines 2022, 10, 628. [Google Scholar] [CrossRef]
  82. Viorica, R.P.; Pawel, P.; Kinga, M.; Michal, Z.; Katarzyna, R.; Boguslaw, B. Lactococcus lactis as a safe and inexpensive source of bioactive silver composites. Appl. Microbiol. Biotechnol. 2017, 101, 7141–7153. [Google Scholar] [CrossRef]
  83. Brown, A.N.; Smith, K.; Samuels, T.A.; Lu, J.; Obare, S.O.; Scotta, M.E. Nanoparticles functionalized with ampicillin destroy multiple-antibiotic-resistant isolates of Pseudomonas aeruginosa and Enterobacter aerogenes and methicillin-resistant Staphylococcus aureus. Appl. Environ. Microbiol. 2012, 78, 2768–2774. [Google Scholar] [CrossRef]
  84. Kalishwaralal, K.; BarathManiKanth, S.; Pandian, S.R.K.; Deepak, V.; Gurunathan, S. Silver nanoparticles impede the biofilm formation by Pseudomonas aeruginosa and Staphylococcus epidermidis. Colloids Surf. B Biointerfaces 2010, 79, 340–344. [Google Scholar] [CrossRef]
  85. Palanisamy, N.K.; Ferina, N.; Amirulhusni, A.N.; Mohd-Zain, Z.; Hussaini, J.; Ping, L.J.; Durairaj, R. Antibiofilm properties of chemically synthesized silver nanoparticles found against Pseudomonas aeruginosa. J. Nanobiotechnol. 2014, 12, 2. [Google Scholar] [CrossRef]
  86. Habash, M.B.; Park, A.J.; Vis, E.C.; Harris, R.J.; Khursigara, C.M. Synergy of silver nanoparticles and aztreonam against Pseudomonas aeruginosa AO1 Biofilms. Antimicrob. Agents Chemother. 2014, 58, 5818–5830. [Google Scholar] [CrossRef]
  87. Loo, C.Y.; Young, P.M.; Cavaliere, R.; Whitchurch, C.B.; Lee, W.H.; Rohanizadeh, R. Silver nanoparticles enhance Pseudomonas aeruginosa PAO1 biofilm detachment. Drug Dev. Ind. Pharm. 2014, 40, 719–729. [Google Scholar] [CrossRef]
  88. Habash, M.B.; Goodyear, M.C.; Park, A.J.; Surette, M.D.; Vis, E.C.; Harris, R.J.; Khursigara, C.M. Potentiation of tobramycin by silver nanoparticles against Pseudomonas aeruginosa biofilms. Antimicrob. Agents Chemother. 2017, 61, e00415-17. [Google Scholar] [CrossRef]
  89. Shaker, M.A.; Shaaban, M.I. Synthesis of silver nanoparticles with antimicrobial and anti-adherence activities against multidrug-resistant isolates from Acinetobacter baumannii. J. Taibah Univ. Sci. 2017, 12, 291–297. [Google Scholar] [CrossRef]
  90. Lara, H.H.; Ayala-Núñez, N.V.; del Carmen Ixtepan Turrent, L.; Rodríguez Padilla, C. Bactericidal effect of silver nanoparticles against multidrug-resistant bacteria. World J. Microbiol. Biotechnol. 2010, 26, 615–621. [Google Scholar] [CrossRef]
  91. Liao, S.; Zhang, Y.; Pan, X.; Zhu, F.; Jiang, C.; Liu, Q.; Cheng, Z.; Dai, G.; Wu, G.; Wang, L.; et al. Antibacterial activity and mechanism of silver nanoparticles against multidrug-resistant Pseudomonas aeruginosa. Int. J. Nanomed. 2019, 14, 1469–1487. [Google Scholar] [CrossRef]
  92. Ramalingam, B.; Parandhaman, T.; Das, S.K. Antibacterial effects of biosynthesized silver nanoparticles on surface ultrastructure and nanomechanical properties of Gram-negative bacteria viz. Escherichia coli and Pseudomonas aeruginosa. ACS Appl. Mater. Interfaces 2016, 8, 4963–4976. [Google Scholar] [CrossRef]
  93. Zhang, Y.; Pan, X.; Liao, S.; Jiang, C.; Wang, L.; Tang, Y.; Wu, G.; Dai, G.; Chen, L. Quantitative proteomics reveals the mechanism of silver nanoparticles against multidrug-resistant Pseudomonas aeruginosa biofilms. J. Proteome Res. 2020, 19, 3109–3312. [Google Scholar] [CrossRef]
  94. Markowska, K.; Grudniak, A.M.; Krawczyk, K.; Wrobel, I.; Wolska, K.I. Modulation of antibiotic resistance and induction of a stress response in Pseudomonas aeruginosa by silver nanoparticles. J. Med. Microbiol. 2014, 63, 849–854. [Google Scholar] [CrossRef]
  95. Panáček, A.; Smékalová, M.; Kilianová, M.; Prucek, R.; Bogdanová, K.; Večeřová, R.; Kolář, M.; Havrdová, M.; Płaza, G.A.; Chojniak, J.; et al. Strong and nonspecific synergistic antibacterial efficiency of antibiotics combined with silver nanoparticles at very low concentrations showing no cytotoxic effect. Molecules 2016, 21, 26. [Google Scholar] [CrossRef]
  96. Panáček, A.; Smékalová, M.; Večeřová, R.; Bogdanová, K.; Röderová, M.; Kolář, M.; Kilianová, M.; Hradilová, S.; Froninga, J.P.; Havrdová, M.; et al. Silver nanoparticles strongly enhance and restore bactericidal activity of inactive antibiotics against multiresistant Enterobacteriaceae. Colloids Surf. B Biointerfaces 2016, 142, 392–399. [Google Scholar] [CrossRef] [PubMed]
  97. Wypij, M.; Czarnecka, J.; Świecimska, M.; Dahm, H.; Rai, M.; Golinska, P. Synthesis, characterization and evaluation of antimicrobial and cytotoxic activities of biogenic silver nanoparticles synthesized from Streptomyces xinghaiensis OF1 strain. World J. Microbiol. Biotechnol. 2018, 34, 23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Pompilio, A.; Germiniani, C.; Bosco, D.; Rana, R.; Aceto, A.; Bucciarelli, T.; Scotti, L.; Bonaventura, G.D. Electrochemically synthesized silver nanoparticles are active against planktonic and biofilm cells of Pseudomonas aeruginosa and other cystic fibrosis-associated bacterial pathogens. Front. Microbiol. 2018, 9, 1349. [Google Scholar] [CrossRef] [PubMed]
  99. Campos-Beleño, C.; Villamizar-Gallardo, R.A.; López-Jacome, L.E.; González, E.E.; Muñoz-Carranza, S.; Franco, B.; Morales-Espinosa, R.; Coria-Jiménez, R.; Franco-Cendejas, R.; Hernández-Durán, M.; et al. Biologically synthesized silver nanoparticles as potent antibacterial effective against multidrug-resistant Pseudomonas aeruginosa. Lett. Appl. Microbiol. 2022. [Google Scholar] [CrossRef]
  100. Huang, Y.; Bai, L.; Yang, Y.; Yin, Z.; Guo, B. Biodegradable gelatin/silver nanoparticle composite cryogel with excellent antibacterial and antibiofilm activity and hemostasis for Pseudomonas aeruginosa-infected burn wound healing. J. Colloid Interface Sci. 2022, 608, 2278–2289. [Google Scholar] [CrossRef]
  101. Jenkins, C.; Rentenaar, R.; Landraud, L.; Brisse, S. Enterobacteriaceae. In Infectious Diseases, 4th ed.; Cohen, J., Powderly, W.G., Opal, S.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; Volume 2, Section 8, Clinical Microbiology: Bacteria; pp. 1565–1578.e2. [Google Scholar] [CrossRef]
  102. Papp-Wallace, K.M.; Endimiani, A.; Taracila, M.A.; Bonomo, R.A. Carbapenems: Past, present, and future. Antimicrob. Agents Chemother. 2011, 55, 4943–4960. [Google Scholar] [CrossRef]
  103. 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]
  104. Baker, C.; Pradhan, A.; Pakstis, L.; Pochan, D.J.; Shah, S.I. Synthesis and antibacterial properties of silver nanoparticles. J. Nanosci. Nanotechnol. 2005, 5, 244–249. [Google Scholar] [CrossRef]
  105. Shahverdi, A.R.; Fakhimi, A.; Shahverdi, H.R.; Minaian, S. Synthesis and effect of silver nanoparticles on the antibacterial activity of different antibiotics against Staphylococcus aureus and Escherichia coli. Nanomed. Nanotechnol. Biol. Med. 2007, 3, 168–171. [Google Scholar] [CrossRef]
  106. Kim, J.S.; Kuk, E.; Yu, K.; Kim, J.H.; Park, S.J.; Lee, H.J.; Kim, S.H.; Park, Y.K.; Hwang, Y.H.; Lee, Y.S.; et al. Antimicrobial effects of silver nanoparticles. Nanomedicine 2007, 3, 95–101. [Google Scholar] [CrossRef]
  107. Vigneshwaran, N.; Kathe, A.A.; Varadarajan, P.V.; Nachane, R.P.; Balasubramanya, R.H. Silver−protein (core−shell) nanoparticle production using spent mushroom substrate. Langmuir 2007, 23, 7113–7117. [Google Scholar] [CrossRef]
  108. Ingle, A.; Gade, A.; Pierrat, S.; Sonnichsen, C.; Rai, M. Mycosynthesis of silver nanoparticles using the fungus Fusarium acuminatum and its activity against some human pathogenic bacteria. Curr. Nanosci. 2008, 4, 141–144. [Google Scholar] [CrossRef]
  109. Sanpui, P.; Murugadoss, A.; Durga Prasad, P.V.; Sankar Ghosh, S.; Chattopadhyay, A. The antibacterial properties of a novel chitosan–Ag-nanoparticle composite. Int. J. Food Microbiol. 2008, 124, 142–146. [Google Scholar] [CrossRef]
  110. Wei, D.; Sun, W.; Qian, W.; Ye, Y.; Ma, X. The synthesis of chitosan-based silver nanoparticles and their antibacterial activity. Carbohydr. Res. 2009, 344, 2375–2382. [Google Scholar] [CrossRef]
  111. Nanda, A.; Saravanan, M. Biosynthesis of silver nanoparticles from Staphylococcus aureus and its antimicrobial activity against MRSA and MRSE. Nanomed. Nanotechnol. Biol. Med. 2009, 5, 452–456. [Google Scholar] [CrossRef]
  112. Inbaneson, S.J.; Ravikumar, S.; Manikandan, N. Antibacterial potential of silver nanoparticles against isolated urinary tract infectious bacterial pathogens. Appl. Nanosci. 2011, 1, 231–236. [Google Scholar] [CrossRef]
  113. Li, W.R.; Xie, X.B.; Shi, Q.S.; Zeng, H.Y.; Ou-Yang, Y.S.; Chen, Y.B. Antibacterial activity and mechanism of silver nanoparticles on Escherichia coli. Appl. Microbiol. Biotechnol. 2010, 85, 1115–1122. [Google Scholar] [CrossRef]
  114. Devina Merin, D.; Prakash, S.; Valentine Bhimba, B. Antibacterial screening of silver nanoparticles synthesized by marine micro algae. Asian Pac. J. Trop. Med. 2010, 3, 797–799. [Google Scholar] [CrossRef]
  115. Sadhasivam, S.; Shanmugam, P.; Yun, K. Biosynthesis of silver nanoparticles by Streptomyces hygroscopicus and antimicrobial activity against medically important pathogenic microorganisms. Colloids Surf. B Biointerfaces 2010, 81, 358–362. [Google Scholar] [CrossRef]
  116. Gnanadhas, D.P.; Thomas, M.B.; Thomas, R.; Raichur, A.M.; Chakravortty, D. Interaction of silver nanoparticles with serum proteins affects their antimicrobial activity in vivo. Antimicrob. Agents Chemother. 2013, 57, 4945–4955. [Google Scholar] [CrossRef]
  117. Anbazhagan, V.; Ahmed, K.B.A.; Janani, S. Synthesis of catalytically active silver nanoparticles using lipid derived signaling molecule, N-steroylethanolamine: Promising antibacterial agent and selective colorimetric sensor for mercury ion. Sens. Actuators B Chem. 2014, 200, 92–100. [Google Scholar] [CrossRef]
  118. Das, B.; Dash, S.K.; Mandal, D.; Ghosh, T.; Chattopadhyay, S.; Tripathy, S.; Das, S.; Dey, S.K.; Das, D.; Somenath, R. Green synthesized silver nanoparticles destroy multidrug resistant bacteria via reactive oxygen species mediated membrane damage. Arabian J. Chem. 2017, 10, 862–876. [Google Scholar] [CrossRef]
  119. Li, P.; Li, J.; Wu, C.Z.; Wu, Q.S.; Li, J. Synergistic antibacterial effects of beta-lactam antibiotic combined with silver nanoparticles. Nanotechnology 2005, 16, 1912–1917. [Google Scholar] [CrossRef]
  120. Fayaz, A.M.; Balaji, K.; Girilal, M.; Yadav, R.; Kalaichelvan, P.T.; Venketesan, R. Biogenic synthesis of silver nanoparticles and their synergistic effect with antibiotics: A study against gram-positive and gram-negative bacteria. Nanomedicine 2010, 6, 103–109. [Google Scholar] [CrossRef] [PubMed]
  121. Xu, N.; Cheng, H.; Xu, J.W.; Li, F.; Gao, B.; Li, Z.; Gao, C.; Huo, K.F.; Fu, J.J.; Xiong, W. Silver-loaded nanotubular structures enhanced bactericidal efficiency of antibiotics with synergistic effect in vitro and in vivo. Int. J. Nanomed. 2017, 12, 731–743. [Google Scholar] [CrossRef] [PubMed]
  122. Kim, S.H.; Lee, H.S.; Ryu, D.S.; Choi, S.J.; Lee, D.S. Antibacterial activity of silver-nanoparticles against Staphylococcus aureus and Escherichia coli. Korean J. Microbiol. Biotechnol. 2011, 39, 77–85. [Google Scholar]
  123. Omara, S.T.; Zawrah, M.F.; Samy, A.A. Minimum bactericidal concentration of chemically synthesized silver nanoparticles against pathogenic Salmonella and Shigella strains isolated from layer poultry farms. J. Appl. Pharm. Sci. 2017, 7, 214–221. [Google Scholar]
  124. Farouk, M.M.; El-Molla, A.; Salib, F.A.; Soliman, Y.A.; Shaalan, M. The role of silver nanoparticles in a treatment approach for multidrug-resistant Salmonella species isolates. Int. J. Nanomed. 2020, 15, 6993–7011. [Google Scholar] [CrossRef]
  125. Huq, M.A.; Akter, S. Biosynthesis, characterization and antibacterial application of novel silver nanoparticles against drug resistant pathogenic Klebsiella pneumoniae and Salmonella Enteritidis. Molecules 2021, 26, 5996. [Google Scholar] [CrossRef]
  126. Losasso, C.; Belluco, S.; Cibin, V.; Zavagnin, P.; Mičetić, I.; Gallocchio, F.; Zanella, M.; Bregoli, L.; Biancotto, G.; Ricci, A. Antibacterial activity of silver nanoparticles: Sensitivity of different Salmonella serovars. Front. Microbiol. 2014, 5, 227. [Google Scholar] [CrossRef]
  127. Pal, I.; Bhattacharyya, D.; Kumar Kar, R.; Zarena, D.; Bhunia, A.; Atreya, H.S. A peptide-nanoparticle system with improved efficacy against multidrug resistant bacteria. Sci. Rep. 2019, 9, 4485. [Google Scholar] [CrossRef]
  128. Picoli, S.U.; Durán, M.; Andrade, P.F.; Duran, N. Silver nanoparticles/silver chloride (Ag/AgCl) synthesized from Fusarium oxysporum acting against Klebsiella pneumoniae carbapenemase (KPC) and extended spectrum beta-lactamase (ESBL). Front. Nanosci. Nanotech. 2016, 2, 107–110. [Google Scholar] [CrossRef]
  129. Panáček, A.; Kvítek, L.; Smékalová, M.; Večeřová, R.; Kolář, M.; Röderová, M.; Dyčka, F.; Šebela, M.; Prucek, R.; Tomanec, O.; et al. Bacterial resistance to silver nanoparticles and how to overcome it. Nature Nanotechnol. 2018, 13, 65–71. [Google Scholar] [CrossRef]
  130. Stapleton, P.D.; Taylor, P.W. Methicillin resistance in Staphylococcus aureus: Mechanisms and modulation. Sci. Prog. 2002, 85 Pt 1, 57–72. [Google Scholar] [CrossRef]
  131. Enright, M.C.; Robinson, D.A.; Randle, G.; Feil, E.J.; Grundmann, H.; Spratt, B.G. The evolutionary history of methicillin-resistant Staphylococcus aureus (MRSA). Proc. Natl. Acad. Sci. USA 2002, 99, 7687–7692. [Google Scholar] [CrossRef]
  132. Romero-Urbina, D.G.; Lara, H.H.; Velázquez-Salazar, J.J.; Arellano-Jiménez, M.J.; Larios, E.; Srinivasan, A.; López-Ribot, J.L.; Yacamán, M.J. Ultrastructural changes in methicillin-resistant Staphylococcus aureus induced by positively charged silver nanoparticles. Beilstein J. Nanotechnol. 2015, 6, 2396–2405. [Google Scholar] [CrossRef]
  133. Sathiyaseelan, A.; Saravanakumar, K.; Mariadoss, A.V.A.; Wang, M.H. Biocompatible fungal chitosan encapsulated phytogenic silver nanoparticles enhanced antidiabetic, antioxidant and antibacterial activity. Int. J. Biol. Macromol. 2020, 153, 63–71. [Google Scholar] [CrossRef]
  134. Ayala-Núñez, N.V.; Lara Villegas, H.H.; Turrent, L.; Padilla, C.R. Silver nanoparticles toxicity and bactericidal effect against Methicillin-Resistant Staphylococcus aureus: Nanoscale does matter. Nanobiotechnology 2009, 5, 2–9. [Google Scholar] [CrossRef]
  135. Wady, A.F.; Machado, A.L.; Foggi, C.C.; Zamperini, C.A.; Zucolotto, V.; Moffa, E.B.; Vergani, C.E. Effect of a silver nanoparticles solution on Staphylococcus aureus and Candida spp. J. Nanomater. 2014, 2014, 545279. [Google Scholar] [CrossRef]
  136. 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. 2019, 72, 50–53. [Google Scholar] [CrossRef]
  137. Masimen, M.A.A.; Harun, N.A.; Maulidiani, M.; Ismail, W.I.W. Overcoming methicillin-resistance Staphylococcus aureus (MRSA) using antimicrobial peptides-silver nanoparticles. Antibiotics 2022, 11, 951. [Google Scholar] [CrossRef] [PubMed]
  138. Ghaffar, N.; Javad, S.; Farrukh, M.A.; Shah, A.A.; Gatasheh, M.K.; AL-Munqedhi, B.M.A.; Chaudhry, O. Metal nanoparticles assisted revival of streptomycin against MDRS Staphylococcus aureus. PLoS ONE 2022, 17, e0264588. [Google Scholar] [CrossRef] [PubMed]
  139. Percival, S.L.; Bowler, P.G.; Dolman, J. Antimicrobial activity of silver-containing dressings on wound microorganisms using an in vitro biofilm model. Int. Wound J. 2007, 4, 186–191. [Google Scholar] [CrossRef] [PubMed]
  140. Mirzajani, F.; Ghassempour, A.; Aliahmadi, A.; Esmaeili, M.A. Antibacterial effect of silver nanoparticles on Staphylococcus aureus. Res. Microbiol. 2011, 162, 542–549. [Google Scholar] [CrossRef]
  141. Li, W.R.; Xie, X.B.; Shi, Q.S.; Duan, S.S.; Ouyang, Y.S.; Chen, Y.B. Antibacterial effect of silver nanoparticles on Staphylococcus aureus. Biometals 2011, 24, 135–141. [Google Scholar] [CrossRef]
  142. Chaudhari, P.R.; Masurkar, S.A.; Shidore, V.B.; Kamble, S.P. Effect of biosynthesized silver nanoparticles on Staphylococcus aureus biofilm quenching and prevention of biofilm formation. Nano-Micro Lett. 2012, 4, 34–39. [Google Scholar] [CrossRef]
  143. Manikprabhu, D.; Lingappa, K. Antibacterial activity of silver nanoparticles against methicillin-resistant Staphylococcus aureus synthesized using model Streptomyces sp. pigment by photo-irradiation method. J. Pharm. Res. 2013, 6, 255–260. [Google Scholar] [CrossRef]
  144. Kasithevar, M.; Saravanan, M.; Prakash, P.; Kumar, H.; Ovais, M.; Barabadi, H.; Khan Shinwari, Z. Green synthesis of silver nanoparticles using Alysicarpus monilifer leaf extract and its antibacterial activity against MRSA and CoNS isolates in HIV patients. J. Interdiscip. Nanomed. 2017, 2, 131–141. [Google Scholar] [CrossRef]
  145. Eleftheriadou, M.; Pyrgiotakis, G.; Demokritou, P. Nanotechnology to the rescue: Using nano-enabled approaches in microbiological food safety and quality. Curr. Opin. Biotechnol. 2017, 44, 87–93. [Google Scholar] [CrossRef]
  146. Greulich, C.; Diendorf, J.; Simon, T.; Eggeler, G.; Epple, M.; Köller, M. Uptake and intracellular distribution of silver nanoparticles in human mesenchymal stem cells. Acta Biomater. 2011, 7, 347–354. [Google Scholar] [CrossRef]
  147. Liz, R.; Simard, J.C.; Leonardi, L.B.; Girard, D. Silver nanoparticles rapidly induce atypical human neutrophil cell death by a process involving inflammatory caspases and reactive oxygen species and induce neutrophil extracellular traps release upon cell adhesion. Int. Immunopharmacol. 2015, 28, 616–625. [Google Scholar] [CrossRef]
  148. Akter, M.; Sikder, M.T.; Rahmana, M.M.; Ullahe, A.K.M.A.; Hossain, K.F.B.; Banik, S.; Hosokawa, T.; Saito, T.; Kurasaki, M. A systematic review on silver nanoparticles-induced cytotoxicity: Physicochemical properties and perspectives. J. Adv. Res. 2018, 9, 1–16. [Google Scholar] [CrossRef]
  149. Li, W.; Li, Y.; Sund, P.; Zhang, N.; Zhaoa, Y.; Qin, S.; Zhao, Y. Antimicrobial peptide-modified silver nanoparticles for enhancing the antibacterial efficacy. RSC Adv. 2020, 10, 38746–38754. [Google Scholar] [CrossRef]
  150. Xu, J.; Li, Y.; Wang, H.; Zhu, M.; Feng, W.; Liang, G. Enhanced antibacterial and anti-biofilm activities of antimicrobial peptides modified silver nanoparticles. Int. J. Nanomed. 2021, 16, 4831–4846. [Google Scholar] [CrossRef]
  151. Hussain, S.M.; Hess, K.L.; Gearhart, J.M.; Geiss, K.T.; Schlager, J.J. In vitro toxicity of nanoparticles in BRL 3A rat liver cells. Toxicol. Vitr. 2005, 19, 975–983. [Google Scholar] [CrossRef]
  152. Arora, S.; Jain, J.; Rajwade, J.M.; Paknikar, K.M. Cellular responses induced by silver nanoparticles. Toxicol. Lett. 2008, 179, 93–100. [Google Scholar] [CrossRef]
  153. Chairuangkitti, P.; Lawanprasert, S.; Roytrakul, S.; Aueviriyavit, S.; Phummiratch, D.; Kulthong, K.; Chanvorachote, P.; Maniratanachote, R. Silver nanoparticles induce toxicity in A549 cells via ROS-dependent and ROS-independent pathways. Toxicol. Vitr. 2013, 27, 330–338. [Google Scholar] [CrossRef]
  154. Zhang, T.; Wang, L.; Chen, Q.; Chen, C. Cytotoxic potential of silver nanoparticles. Yonsei Med. J. 2014, 55, 283–291. [Google Scholar] [CrossRef]
  155. Składanowski, M.; Golinska, P.; Rudnicka, K.; Dahm, H.; Rai, M. Evaluation of cytotoxicity, immune compatibility and antibacterial activity of biogenic silver nanoparticles. Med. Microbiol. Immunol. 2016, 205, 603–613. [Google Scholar] [CrossRef]
  156. Asharani, P.V.; Hande, M.P.; Valiyaveettil, S. Anti-proliferative activity of silver nanoparticles. BMC Cell Biol. 2009, 10, 65. [Google Scholar] [CrossRef]
  157. Luther, E.M.; Koehler, Y.; Diendorf, J.; Epple, M.; Dringen, R. Accumulation of silver nanoparticles by cultured primary brain astrocytes. Nanotechnology 2011, 22, 375101. [Google Scholar] [CrossRef] [PubMed]
  158. Asharani, P.V.; Low Kah Mun, G.; Hande, M.P.; Valiyaveettil, S. Cytotoxicity and genotoxicity of silver nanoparticles in human cells. ACS Nano 2009, 3, 279–290. [Google Scholar] [CrossRef] [PubMed]
  159. Yang, E.J.; Kim, S.; Kim, J.S.; Choi, I.H. Inflammasome formation and IL-1b release by human blood monocytes in response to silver nanoparticles. Biomaterials 2012, 33, 6858–6867. [Google Scholar] [CrossRef] [PubMed]
  160. Jiang, X.; Foldbjerg, R.; Miclaus, T.; Wang, L.; Singh, R.; Hayashi, Y.; Sutherland, D.; Chen, C.; Autrup, H.; Beer, C. Multi-platform genotoxicity analysis of silver nanoparticles in the model cell line CHO-K1. Toxicol. Lett. 2013, 222, 55–63. [Google Scholar] [CrossRef]
  161. Hackenberg, S.; Scherzed, A.; Kessler, M.; Hummel, S.; Technau, A.; Froelich, K.; Ginzkey, C.; Koehler, C.; Hagen, R.; Kleinsasser, N. Silver nanoparticles: Evaluation of DNA damage, toxicity and functional impairment in human mesenchymal stem cells. Toxicol. Lett. 2011, 201, 27–33. [Google Scholar] [CrossRef]
  162. 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, 35. [Google Scholar] [CrossRef]
  163. Manivasagan, P.; Venkatesan, J.; Senthilkumar, K.; Sivakumar, K.; Kim, S.K. Biosynthesis, antimicrobial and cytotoxic effect of silver nanoparticles using a novel Nocardiopsis sp. MBRC-1. Bio. Med. Res. Int. 2013, 2013, 287638. [Google Scholar] [CrossRef]
  164. Rathod, D.; Golinska, P.; Wypij, M.; Dahm, H.; Rai, M. A new report of Nocardiopsis valliformisstrain OT1 from alkaline Lonar crater of India and its use in synthesis of silver nanoparticles with special reference to evaluation of antibacterial activity and cytotoxicity. Med. Microbiol. Immunol. 2016, 205, 435–447. [Google Scholar] [CrossRef]
  165. Barbasz, A.; O’cwieja, M.; Romanc, M. Toxicity of silver nanoparticles towards tumoral human cell lines U-937 and HL-60. Colloids Surf. B Biointerfaces 2017, 156, 397–404. [Google Scholar] [CrossRef]
  166. Kvítek, L.; Vanickova, M.; Panáček, A.; Soukupova, J.; Dittrich, M.; Valentova, E.; Prucek, R.; Bancirova, M.; Milde, D.; Zbořil, R. Initial study on the toxicity of silver nanoparticles (nps) against Paramecium caudatum. J. Phys. Chem. C 2009, 113, 4296–4300. [Google Scholar] [CrossRef]
  167. Panáček, A.; Kolar, M.; Večeřová, R.; Prucek, R.; Soukupova, J.; Krystof, V.; Hamal, P.; Zbořil, R.; Kvítek, L. Antifungal activity of silver nanoparticles against Candida spp. Biomaterials 2009, 30, 6333–6340. [Google Scholar] [CrossRef]
  168. Panáček, A.; Prucek, R.; Safarova, D.; Dittrich, M.; Richtrova, J.; Benickova, K.; Zbořil, R.; Kvítek, L. Acute and chronic toxicity effects of silver nanoparticles (nps) on Drosophila melanogaster. Environ. Sci. Technol. 2011, 45, 4974–4979. [Google Scholar] [CrossRef]
  169. Krajewski, S.; Prucek, R.; Panáček, A.; Avci-Adali, M.; Nolte, A.; Straub, A.; Zbořil, R.; Wendel, H.P.; Kvítek, L. Hemocompatibility evaluation of different silver nanoparticle concentrations employing a modified chandler-loop in vitro assay on human blood. Acta Biomater. 2013, 9, 7460–7468. [Google Scholar] [CrossRef]
  170. Munger, M.A.; Radwanski, P.; Hadlock, G.C.; Stoddard, G.; Shaaban, A.; Falconer, J.; Grainger, D.W.; Deering-Rice, C.E. In vivo human time-exposure study of orally dosed commercial silver nanoparticles. Nanomedicine 2014, 10, 1–9. [Google Scholar] [CrossRef] [PubMed]
  171. Cha, K.; Hong, H.W.; Choi, Y.G.; Lee, M.J.; Park, J.H.; Chae, H.K.; Ryu, G.; Myung, H. Comparison of acute responses of mice livers to short-term exposure to nano-sized or micro-sized silver particles. Biotechnol. Lett. 2008, 30, 1893–1899. [Google Scholar] [CrossRef]
  172. 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]
  173. Kim, Y.S.; Song, M.Y.; Park, J.D.; Song, K.S.; Ryu, H.R.; Chung, Y.H.; Chang, H.K.; Lee, J.H.; Oh, K.H.; Kelman, B.J.; et al. Subchronic oral toxicity of silver nanoparticles. Part. Fibre Toxicol. 2010, 7, 20. [Google Scholar] [CrossRef]
  174. Park, E.J.; Bae, E.; Yi, J.; Kim, Y.; Choi, K.; Lee, S.H.; Yoon, J.; Lee, B.C.; Park, K. Repeated-dose toxicity and inflammatory responses in mice by oral administration of silver nanoparticles. Environ. Toxicol. Pharmacol. 2010, 30, 162–168. [Google Scholar] [CrossRef]
  175. Loeschner, K.; Hadrup, N.; Qvortrup, K.; Larsen, A.; Gao, X.Y.; Vogel, U.; Mortensen, A.; Lam, H.R.; Larsen, E.H. Distribution of silver in rats following 28 days of repeated oral exposure to silver nanoparticles or silver acetate. Part. Fibre Toxicol. 2011, 8, 18. [Google Scholar] [CrossRef]
  176. Hadrup, N.; Loeschner, K.; Mortensen, A.; Sharma, A.K.; Qvortrup, K.; Larsen, E.H.; Lam, H.R. The similar neurotoxic effects of nanoparticulate and ionic silver in vivo and in vitro. Neurotoxicology 2012, 33, 416–423. [Google Scholar] [CrossRef]
  177. Van der Zande, M.; Vandebriel, R.J.; van Doren, E.; Kramer, E.; Rivera, Z.H.; Serrano-Rojero, C.S.; Gremmer, E.R.; Mast, J.; Peters, R.J.B.; Hollman, P.C.H.; et al. Distribution, elimination, and toxicity of silver nanoparticles and silver ions in rats after 28-day oral exposure. ACS Nano 2012, 6, 7427–7442. [Google Scholar] [CrossRef] [PubMed]
  178. Kim, J.S.; Song, K.S.; Sung, J.H.; Ryu, H.R.; Choi, B.G.; Cho, H.S.; Lee, J.K.; Yu, I.J. Genotoxicity, acute oral and dermal toxicity, eye and dermal irritation and corrosion and skin sensitisation evaluation of silver nanoparticles. Nanotoxicology 2013, 7, 953–960. [Google Scholar] [CrossRef]
  179. Park, K. Toxicokinetic differences and toxicities of silver nanoparticles and silver ions in rats after single oral administration. J. Toxicol. Environ. Health Part A 2013, 76, 1246–1260. [Google Scholar] [CrossRef] [PubMed]
  180. Korani, M.; Rezayat, S.M.; Gilani, K.; Bidgoli, S.A.; Adeli, S. Acute and subchronic dermal toxicity of nanosilver in guinea pig. Int. J. Nanomed. 2011, 6, 855–862. [Google Scholar] [CrossRef] [PubMed]
  181. Tang, J.L.; Xiong, L.; Wang, S.; Wang, J.Y.; Liu, L.; Li, J.A.; Wan, Z.Y.; Xi, T.F. Influence of silver nanoparticles on neurons and blood–brain barrier via subcutaneous injection in rats. Appl. Surf. Sci. 2008, 255, 502–504. [Google Scholar] [CrossRef]
  182. Tang, J.L.; Xiong, L.; Wang, S.; Wang, J.Y.; Liu, L.; Li, J.G.; Yuan, F.Q.; Xi, T.F. Distribution, translocation and accumulation of silver nanoparticles in rats. J. Nanosci. Nanotechnol. 2009, 9, 4924–4932. [Google Scholar] [CrossRef] [PubMed]
  183. Padilla-Camberos, E.; Juárez-Navarro, K.J.; Sánchez-Hernández, I.M.; Torres-González, O.R.; Flores-Fernández, J.M. Toxicological Evaluation of Silver Nanoparticles Synthesized with Peel Extract of Stenocereus queretaroensis. Materials 2022, 15, 5700. [Google Scholar] [CrossRef]
  184. Cho, Y.M.; Mizuta, Y.; Akagi, J.I.; Toyoda, T.; Sone, M.; Ogawa, K. Size-dependent acute toxicity of silver nanoparticles in mice. J. Toxicol. Pathol. 2018, 31, 73–80. [Google Scholar] [CrossRef]
  185. Recordati, C.; De Maglie, M.; Bianchessi, S.; Argentiere, S.; Silvana, C.; Mattiello, C.; Cubadda, F.; Aureli, F.; D’Amato, M.; Raggi, A.; et al. Tissue distribution and acute toxicity of silver after single intravenous administration in mice: Nano-specific and size-dependent effects. Part. Fibre Toxicol. 2016, 13, 12. [Google Scholar] [CrossRef]
  186. Yildirimer, L.; Thanh, N.T.K.; Loizidou, M.; Seifalian, A.M. Toxicological considerations of clinically applicable nanoparticles. Nano Today 2012, 6, 585–607. [Google Scholar] [CrossRef]
  187. Emond, C. Kinetic behaviour of nanoparticles across the biological physiology. J. Phys. Conf. Ser. 2011, 304, 12089. [Google Scholar] [CrossRef]
  188. Smock, K.J.; Schmidt, R.L.; Hadlock, G.; Stoddard, G.; Grainger, D.W.; Munger, M.A. Assessment of orally dosed commercial silver nanoparticles on human ex vivo platelet aggregation. Nanotoxicology 2014, 8, 328–333. [Google Scholar] [CrossRef]
  189. Li, N.Y.; Ko, W.C.; Hsueh, P.R. Nanoparticles in the treatment of infections caused by multidrug-resistant organisms. Front. Pharmacol. 2019, 10, 11533. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Possible mechanisms of action of nanoparticles in the bacteria.
Figure 1. Possible mechanisms of action of nanoparticles in the bacteria.
Antibiotics 11 01205 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mateo, E.M.; Jiménez, M. Silver Nanoparticle-Based Therapy: Can It Be Useful to Combat Multi-Drug Resistant Bacteria? Antibiotics 2022, 11, 1205. https://doi.org/10.3390/antibiotics11091205

AMA Style

Mateo EM, Jiménez M. Silver Nanoparticle-Based Therapy: Can It Be Useful to Combat Multi-Drug Resistant Bacteria? Antibiotics. 2022; 11(9):1205. https://doi.org/10.3390/antibiotics11091205

Chicago/Turabian Style

Mateo, Eva M., and Misericordia Jiménez. 2022. "Silver Nanoparticle-Based Therapy: Can It Be Useful to Combat Multi-Drug Resistant Bacteria?" Antibiotics 11, no. 9: 1205. https://doi.org/10.3390/antibiotics11091205

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