Next Article in Journal
Assessing Drug Administration Techniques in Zebrafish Models of Neurological Disease
Previous Article in Journal
Exploring the Cellular and Molecular Mechanism of Discoidin Domain Receptors (DDR1 and DDR2) in Bone Formation, Regeneration, and Its Associated Disease Conditions
Previous Article in Special Issue
Metallic Nanosystems in the Development of Antimicrobial Strategies with High Antimicrobial Activity and High Biocompatibility
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Metallic Nanoparticles: A Promising Arsenal against Antimicrobial Resistance—Unraveling Mechanisms and Enhancing Medication Efficacy

1
Department of Food Science and Technology, College of Agriculture and Life Sciences, Jeonbuk National University, Jeonju 54896, Republic of Korea
2
Department of Agricultural Convergence Technology, College of Agriculture and Life Science, Jeonbuk National University, Jeonju 54896, Republic of Korea
3
Nanobiotechnology Laboratory, Department of Biotechnology University of Malakand, Dir Lower, Chakdara 18800, Khyber Pakhtunkhwa, Pakistan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(19), 14897; https://doi.org/10.3390/ijms241914897
Submission received: 17 August 2023 / Revised: 25 September 2023 / Accepted: 3 October 2023 / Published: 4 October 2023
(This article belongs to the Special Issue Nanoparticles: New Antimicrobial Agents)

Abstract

:
The misuse of antibiotics and antimycotics accelerates the emergence of antimicrobial resistance, prompting the need for novel strategies to combat this global issue. Metallic nanoparticles have emerged as effective tools for combating various resistant microbes. Numerous studies have highlighted their potential in addressing antibiotic-resistant fungi and bacterial strains. Understanding the mechanisms of action of these nanoparticles, including iron-oxide, gold, zinc oxide, and silver is a central focus of research within the life science community. Various hypotheses have been proposed regarding how nanoparticles exert their effects. Some suggest direct targeting of microbial cell membranes, while others emphasize the release of ions from nanoparticles. The most compelling proposed antimicrobial mechanism of nanoparticles involves oxidative damage caused by nanoparticles-generated reactive oxygen species. This review aims to consolidate knowledge, discuss the properties and mechanisms of action of metallic nanoparticles, and underscore their potential as alternatives to enhance the efficacy of existing medications against infections caused by antimicrobial-resistant pathogens.

1. Introduction

Infectious diseases caused by microbes are a significant global health concern, with an escalating economic burden. The continuous increase in antibiotic resistance rates, which continues to accelerate annually [1], highlights the urgent need for innovative solutions. This concerning trend has resulted in higher mortality and morbidity rates among patients, leading the World Health Organization (WHO) to designate antimicrobial resistance as one of the top three global health challenges [2]. The consequences of antimicrobial resistance extend across various aspects of healthcare, including cancer treatment, premature infant care, transplantation, and surgical procedures. All of these are susceptible to severe disruptions without effective strategies to combat drug-resistant bacteria [3]. The challenges of rising treatment costs, prolonged treatment durations, and increased mortality rates due to antimicrobial resistance require the exploration of effective alternatives [4]. Efforts to address these challenges involve the promotion of diverse novel approaches to regulate antimicrobial medications [5]. These alternatives encompass a range of interventions, including vaccine development, phage therapy, immune stimulants, adjuvants, anti-treatment agents, probiotics, and their various derivatives [6]. Pre-disease vaccination, which aims to prevent diseases by strengthening the human immunological response to bacterial infections, is also a viable strategy. However, widespread implementation faces obstacles related to cost and accessibility [7]. Similarly, strategies like probiotics, prebiotics, symbiotics, and competitive exclusion have been established to deter pathogenic colonization but are constrained by variable efficacy and regulatory processes, such as those imposed by the Food and Drug Administration (FDA) [8]. Furthermore, bacterial cells employ primary defense mechanisms, including enzyme function inhibition and efflux pumps, to reduce susceptibility to antibiotics [9]. Consequently, the era of antibiotics teeters on the brink of extinction, demanding the innovation of new approaches to combat multidrug-resistant strains. Researchers are actively seeking alternative methods to address this pressing issue. In this contemporary context, nano-sized materials emerge as a promising alternative to traditional antimicrobial agents [10]. Nanoparticles (NPs) exhibit distinctive physico-chemical properties, including their controllable small size (ranging from 1 to 100 nm), high reactivity, functionalized structure, and large surface area-to-mass ratio [11]. Leveraging NPs as delivery vehicles for antimicrobial drugs has proven highly effective, mitigating numerous limitations associated with conventional antimicrobial therapies [12]. Notably, microorganisms find it challenging to develop resistance to NPs because these versatile entities can concurrently target multiple cellular pathways [13]. Consequently, NPs have the potential to replace traditional antibiotics and antifungal agents in the treatment of microorganism infections that have become resistant to standard therapies [14]. The synthesis of NPs can be achieved through various methods, each with its own set of advantages and disadvantages. These methods encompass biological, physical, and chemical approaches [15]. Physical methods often yield high quantities of NPs, making them attractive for large-scale production; however, they tend to be energy-intensive and cost-inefficient, which may limit their practicality for some applications [16]. On the other hand, chemical synthesis methods are generally cost-effective and efficient in producing NPs, but they often involve the use of hazardous and volatile chemicals, which can pose environmental risks and safety concerns [17]. In recent years, there has been a growing emphasis on environmentally friendly approaches to NPs synthesis. One such approach is green synthesis, which has garnered significant attention due to its cost-effectiveness, environmental sustainability, and relative simplicity [18]. Green synthesis methods utilize plant extracts, bacteria, and fungi to produce NPs. Among these, plant extracts are particularly desirable as they eliminate the need for complex cell culture maintenance and downstream processing [19]. Harnessing their exceptional chemical properties and physical characteristics, NPs offer a promising avenue for addressing the challenge posed by multidrug-resistant bacteria. In today’s challenging landscape of antibiotic resistance, the need for effective strategies to combat multidrug-resistant strains has reached a critical juncture. This pressing issue necessitates a comprehensive approach that spans research, development, and implementation. Throughout this manuscript, we will delve into the multifaceted nature of effective strategies, exploring their significance in addressing the threat posed by multidrug-resistant strains. From innovative methodologies to strategic interventions, our discussion will emphasize the importance of these strategies as a central theme in our quest to combat this global health crisis. This study will cover various types of metallic NPs, delving into their unique mechanisms of action. It will particularly focus on investigating the antimicrobial properties of different metallic NPs, evaluating how NPs can disrupt multidrug resistance in bacteria and fungi, and assessing their potential as a viable solution for combating bacterial infections.

2. Antibiotic Resistance

Antibiotics are used to combat microbial infections, utilizing a range of mechanisms such as inhibiting enzymes, altering membrane structure, and disrupting transcription and translation processes [15]. However, some microbes have evolved to become resistant to antibiotics, posing a significant challenge to the efficacy of antimicrobial agents [20]. Antibiotic resistance is a primary factor contributing to increased drug dosages, extended hospital stays, heightened toxicity, and elevated mortality rates [21]. Multiple factors, such as the over-prescription, misuse, and excessive use of antibiotics, including their extensive use in agriculture, and the scarcity of new antibiotics, all contribute to the occurrence of antibiotic resistance [22].
Antibiotic resistance is a formidable challenge in combating microbial infections caused by bacteria. However, it is important to note that antibiotic resistance is not confined solely to bacteria. Fungi, another category of microorganisms, can also develop mechanisms of resistance to antimicrobial agents [23]. While the primary focus of this section remains on bacterial resistance, the emergence of antifungal resistance is a significant concern in medical and agricultural settings. Fungi, like bacteria, can adapt and develop various strategies to resist the effects of antifungal drugs, posing additional complexities in the fight against microbial infections. Although this review predominantly addresses bacterial antibiotic resistance, it is vital to recognize the broader landscape of antimicrobial challenges, which includes both bacterial and fungal aspects [24].
Bacteria employ multiple mechanisms for developing resistance to antibiotics, including intrinsic mechanisms that are typically genetically transmitted. Unlike eukaryotic cells, bacteria are prokaryotes and lack a nuclear membrane to protect their genetic material [25]. As a result, random or spontaneous mutations can occur frequently in the bacterial genome due to the exposed nature of their genetic material. Excessive exposure of a bacterial strain to a particular antibiotic can lead to genetic mutations that give rise to a novel protein that helps the bacterium fight against the antibiotic [26]. The bacterium employs various methods to develop intrinsic resistance to antibiotics, and one of these methods involves utilizing newly produced proteins as tools for survival [27]. Specifically, the protein functions by obstructing the intake of the antibiotic into a bacterial cell, thereby conferring resistance [28]. This involves the production of a mutated protein that triggers the efflux pump present in the bacterial cellular wall. As a result, after antibiotics enter a cell, they are recognized by the altered protein and subsequently pumped out from the cell through an efflux pump [29]. Additionally, bacteria can utilize an enzymatic reaction to inactivate antibiotics as another intrinsic resistance strategy [30]. Bacteria can acquire antibiotic resistance through various mechanisms, such as modifying the antibiotic target, evading the antibiotic target, and vertical gene transfer [31]. The adaptive mechanism employed by bacteria during an infection involves the development of biofilms, which play a crucial role in their survival [32]. Biofilms, complex communities of microorganisms, also interact with their environment. In the case of metalloids and heavy metals, these biofilms can act as both a protective barrier and a source of resistance. They have been found to sequester and immobilize metalloids and heavy metals, reducing their toxicity to the bacteria within the biofilm. These arrangements eventually lead to reduced buildup of antibiotics within the bacteria cells, leading to suboptimal therapeutic levels of the medication [33]. Consequently, higher and more frequent doses of antibiotics may be necessary, which can have dangerous effects on humans and animals. Figure 1 illustrates the mechanism of antibiotic resistance.

3. Mechanistic Action of MNPs against Bacteria

NPs exhibit unique properties that set them apart from their macroscopic counterparts, making them highly effective in various applications. These properties include crystallinity, exceptional stability, reduced size, the surface plasmon resonance effect, unique shapes, and a higher surface-to-volume ratio [34]. These attributes bestow upon NPs exceptional antibacterial, antifungal, larvicidal, and antiprotozoal capabilities. Specifically, NPs’ distinct size, crystal structure, and reduced dimensions position them as superior alternatives to existing antibiotics, offering potential relief from the burden of antibiotic resistance [35]. Additionally, NPs demonstrate manageable morphologies and excellent size uniformity. Anisotropy, a crucial property of NPs, implies that different facets of their crystal structure possess distinct reactivity. The optical properties of metallic salts transform when converted into nanoforms, enabling significant customization of these characteristics. These remarkable property alterations, combined with the ability to tailor them to specific requirements, establish NPs as a highly promising avenue for addressing antibiotic resistance. Moreover, these changes in properties, coupled with the potential for tailoring them to specific needs, position NPs as a highly promising avenue for addressing antibiotic resistance. Several factors contribute to the antibacterial activity of NPs against bacteria. These include their large surface area that comes into contact with bacteria, electrostatic interactions, or hydrophobic interactions [36]. NPs that interfere with bacteria typically cause oxidative stress, enzyme inhibition, protein degradation, and changes in gene expression [37]. Nevertheless, oxidative stress, metal release, and non-oxidative pathways are the most common antibacterial mechanisms [38]. Among these mechanisms, Reactive Oxygen Species (ROS)-induced oxidative stress stands out as one of the main mechanisms supporting NPs in their antibacterial operation. In bacteria, ROS is produced primarily through aerobic respiration and is balanced by antioxidant cell machinery; however, an excess of ROS can lead to more significant insults, biomolecular oxidation, and cellular damage [39]. Further, when investigating the generation of ROS by NPs, it is crucial to consider the environmental conditions under which this phenomenon occurs. ROS generation by NPs can vary significantly depending on whether it takes place under light or dark conditions [40]. NPs can be viewed as reactive centers, particularly when exposed to electromagnetic activation, such as light conditions, which can greatly influence their intensity and kinetics. Metal ions are slowly released by metal oxides that are carried by the cell to the inner cell, where they interact with protein and nucleic acid functional groups [41]. This interface alters the composition of the holding cell, blocks the enzyme, and intervenes with normal bacteria in the cellular processes [42]. NPs that interfere with the bacteria cell wall create a focal stream of ions that continually emit ions and cause more toxicity to the cells [43]. The surface morphological features of NPs have fundamental effects on their behavior, and where the surface of the NPs is rougher; dissolution happens more rapidly [44]. The association of NPs with the cell wall is implicated in non-oxidative processes. In bacteria, the protective barriers to environmental defense are the cell membrane and cell wall. Different adsorbent ways for NPs are given by the components of the cell membrane and component [45]. The Gram-negative bacterial cell wall comprises phospholipids, lipopolysaccharides, and lipoproteins, forming a barrier only for certain macro molecules to enter [46]. The cell wall consists of a thin layer of peptidoglycans and abundant pores for the Gram-positive bacteria, which allow foreign molecules to penetrate contributing to the binding to proteins and other cellular components covalently that disrupt bacterial cell activity [47]. Lipid polysaccharides on the Gram-negative bacterial cell wall create regions that are negatively charged and attract NPs [48], and because teichoic acid is expressed only in Gram-positive bacterial strains, NPs are spread throughout the phosphate chain. The result is, therefore, more pronounced in Gram-positive bacteria more than that of Gram-negative bacteria [37]. As an example, Zinc oxide NPs (ZnONPs) were synthesized, and their antimicrobial activity against various bacteria was assessed. The results revealed a reliance on the structure and composition of the bacterial cell wall for the antibacterial effects. These NPs exhibited more potent antibacterial activity against Gram-positive bacteria, while specific components appeared to resist the adhesion of ZnONP to bacterial membranes [49]. Figure 2 show the mechanistic approach of NPs against bacteria.

3.1. Effect of Silver NPs against Bacteria

Among different types of NPs, silver NPs (AgNPs) stand out as one of the most potent antimicrobial agents [50]. When considering the antimicrobial properties of AgNPs, it is vital to delve into the diverse mechanisms underlying their effectiveness. These mechanisms are influenced not only by NPs size but also by a spectrum of experimental conditions, including dissolution kinetics, pH levels in the medium, solution ionic strength, media composition, and the specific microorganisms or biofilms encountered. One fundamental aspect of AgNPs’ antimicrobial action is their interaction with the microbial cell membrane. AgNPs possess a remarkable ability to disrupt membrane permeability and interfere with cellular respiration [51]. This disruption stems from AgNPs binding to and penetrating the cell membrane, ultimately affecting its integrity. Additionally, AgNPs can engage with thiol groups within microbial cells. This interaction has the potential to trigger the production of ROS, which can disrupt respiratory enzymes, ultimately leading to microbial cell death [51]. For example, the bactericidal effect of AgNPs against Escherichia coli has been observed, potentially attributed to the disruption of the proton motive force [52]. The mode of action of AgNPs, although explored over the last few decades, remains a topic of ongoing research and discussion. Some studies suggest that AgNPs kill or inhibit the growth of resistant bacteria by producing Ag+ ions. In this scenario, a redox reaction occurs within the bacterial cell when exposed to AgNPs, resulting in the production of silver ions due to their oxidation [53]. These silver ions then alter the macromolecules which lead to the growth inhibition of bacteria. AgNPs tend to affect bacteria directly rather than producing silver ions. It is proposed that AgNPs degrade the cell membrane of bacteria as they are positively charged and have a higher affinity towards the negatively charged peptidoglycan layer of the bacterial cell membrane [54]. It has also been concluded from some research studies that AgNPs react with the DNA of bacteria and control its replication [51]. When AgNPs penetrate through the cell membrane of bacteria, they release silver ions, and these silver ions turn the DNA into its condensed form and alter its replication process [55]. This alteration results in no replication, and cell death occurs as an endpoint of this reaction.

3.2. Effect of Zinc Oxide NPs against Bacteria

To combat the problem of antibiotic resistance, ZnONPs have shown promise in different applications, particularly in combating antibiotic resistance [56]. Among the properties of ZnONP, chemical sensing, semiconducting, electric conductivity, and piezoelectric are included [57]. ZnONP exhibits biocompatibility with human cells and demonstrates efficacy as an antimicrobial agent. For instance, Bacillus subtilis and Staphylococcus aureus, which are Gram-positive bacteria, werefound to be sensitive to ZnONP [58]. Escherichia coli, Pseudomonas aeruginosa, and Campylobacter jejuni are some Gram-negative bacteria found to be influenced by ZnONP [59]. The antibacterial potential of ZnONP depends on their particle size, morphology, and concentration [60]. Various ZnONP structures have significantly different antibacterial activities. ZnONP of rods and wires shape can easily discern bacterial cell walls as compared to spherical shapes [61]. Flower-shaped particles have been proven to be more efficient than both rod and spherical shaped when observed against Staphylococcus aureus and Escherichia coli [60]. Enhanced antibacterial activity of ZnONP with a larger surface area and increased concentration was reported [62]. ZnONP of smaller sizes have been proven to be comparatively more effective in their antibacterial activities because of their enhanced potential for penetration [63]. The ZnONP activity was also checked against Staphylococcus aureus and Escherichia coli and effective results were concluded due to their smaller particle size [64]. These size-dependent activities of ZnONP were the same for both Gram-positive and Gram-negative strains [65]. The concentration of ZnONP has noticeable effects on antibacterial activity [66]. A higher concentration of ZnONP can have enhanced antibacterial activity to increase cell death [67]. In a research study, 80 nm-sized ZnONP was employed to combat Escherichia coli. The findings revealed that the suppression of bacterial growth was more pronounced at higher concentrations of ZnONP compared to lower concentrations [68].
Studies about mechanisms through which ZnONP kills bacteria and acts as an antibacterial agents are very limited. Various modes of action of ZnONP are concluded from different research studies. Like AgNPs, ZnONP also inhibits the growth of bacteria by generating ROS such as hydrogen peroxides and hydroxyl radicals’ induction. ROS itself is the cause of different antibacterial mechanisms. For example, ZnO localized interaction causes cell wall damage [69]. ZnONPs exhibit another mode of action by altering the cell membrane and permeability of bacteria. Additionally, they function as carriers of zinc ions, releasing these ions through an oxidation reaction once they enter the bacterial cell. These released zinc ions are highly toxic to bacteria, particularly by weakening the mitochondria [70]. ROS alters several macromolecules in the cell. As a result, they express the oxidative stress gene causing growth inhibition and cell death [71]. ZnONPs are used in the food packaging industries for their potential of incorporation into packaging materials as they release NPs against bacteria and prevent foodborne diseases [72].

3.3. Effect of Gold NPs against Bacteria

Gold NPs (AuNPs) are a comparatively a more unique type of NPs. AuNPs have various clinical applications in the field of nanomedicine against different strains of bacteria [73]. AuNPs synthesis can be achieved through various techniques, including physical methods such as laser ablation, chemical methods involving chloroauric acid, and biological approaches utilizing plants and their extracts. Among these methods, biological approaches, also known as green synthesis, are favored for their reduced environmental impact, cost-effectiveness, and ease of handling [74]. AuNPs hold the potential for conjugation with other antibiotics, antibacterial peptides, and specific antigens [75]. While AuNPs do exhibit antibacterial properties, they may not be as effective against bacteria as AgNPs [76], However, it is worth noting that they can be cytotoxic and genotoxic to animal cells [77]. AuNPs exhibit low toxicity and greater cytocompatibility when compared to AgNPs. Antibacterial activities of AuNPs against both Gram-negative and Gram-positive strains of bacteria have been investigated [78]. For example, a study assessed the effects of these NPs on both Mycobacterium tuberculosis and Escherichia coli, revealing significant antibacterial activity against Gram-positive Mycobacterium tuberculosis and Gram-negative Escherichia coli [79]. Research into the antibacterial mechanisms of AuNPs is currently a topic of active investigation. A study reported that AuNPs can interact with the cell membranes of Escherichia coli, leading to membrane degradation and growth inhibition. They also observed that the inhibitory effects varied with changes in surface modification agents [80]. AuNPs with similar size and shape caused cell lysis when PAH (Poly-allylamine hydrochloride) was capped, but not in the case of citrate-capped particles [81]. One of the modes of action of AuNPs involves interaction with cellular barriers, such as cell walls and cell membranes [82]. Another study highlighted distinct interactions between AuNPs and the protective barriers of Gram-negative strains when compared to Gram-positive bacteria. In this investigation, it was observed that Au-DAPT-coated gold nanoparticles significantly increased the permeability of E. coli membranes by 70%, leading to nucleic acid leakage. In contrast, the impact on P. aeruginosa was measured at 42% [83]. AuNPs can also affect bacteria by interacting with various biological substances like DNA and proteins, inhibiting enzyme activity and neutralizing plasmid movement within bacteria [84]. For instance, it was proposed that AuNPs induce DNA fragmentation, ultimately resulting in bacterial cell death [85]. Another study suggested that the photothermal effect of AuNPs is one of their convincing mechanisms, converting infrared light energy into local heat [86]. This heat causes protein denaturation, cell fluid evaporation, and structural breakdown, resulting in bacterial growth inhibition or death. Furthermore, AuNPs can cause a redox imbalance, similar to AgNPs, by generating ROS that induce cell apoptosis and necrosis. They have the potential to disrupt the macromolecules of bacteria, leading to oxidative stress [87]. Despite these findings, further research is needed to fully uncover the antibacterial mode of action of AuNPs.

3.4. Effect of Iron NPs against Bacteria

  • Like other metallic NPs, iron NPs (FeNPs) have demonstrated their antimicrobial efficacy against numerous pathogenic bacteria, suggesting their potential for use in combating microbial infections [88]. FeNPs are significantly cost-effective compared to silver and gold NPs [89], they are also slightly preferred over silver and other NPs as they are less toxic to humans unlike other NPs, especially silver, which can cause cytotoxicity in various human cell lines [90]. FeNPs are considered less toxic and therefore, Ferumoxytol an intravenous Fe3O4 formulation, was also approved by the FDA as a treatment for iron-deficient patients. Apart from being less toxic, the byproduct of FeO-NPs, which is iron, can be stored by the body [91]. Antibacterial activities of NPs vary from 80–100 nm of semi-crystalline biogenic Fe3O4 that have been developed from leaf extract of T. procumbens and revealed to have bactericidal activity against Gram-negative bacterium P. aeruginosa [92]. Similarly, it was reported that Fe2O3−NPs produced from Skimmia laureola exhibit the highest antibacterial activity against Ralstonia solanacearum by degrading the cell wall [93]. In another study, it was suggested that rod-shaped FeNPs produced using Eichhornia crassipes leaf extract exhibited the highest inhibition against Pseudomonas fluorescens and Staphylococcus aureus [94]. FeNPs synthesized from Gardenia jasminoides and Lawsonia inermis, were tested against Staphylococcus aureus, and displayed a noticeable increase in the inhibition zone, going from 15 mm to 16 mm [95]. The antimicrobial activity of FeNPs depends on its surface coating. Different NPs have a different surface coating, and this property of NPs seems to play a huge role in their antibacterial efficiency [96]. FeNPs are used as antimicrobial agents and are usually capped with other compounds, such as alginate, for stability purposes because uncapped FeNPs are slightly unstable [97,98]. FeNPs prove their efficacy through the presence of iron, resulting in the production of ROS, such as H2O2, that can damage the cell membrane once they enter the intracellular space, ultimately leading to bacterial cell death. The bactericidal activity of FeNPs is a result of the oxidative stress caused by FeNPs-generated ROS [99]. Another possible mechanism of FeO-NPs is the damage of DNA through Fenton’s reaction [100]. In Fenton’s reaction, the superoxide anion O- is dismutased to hydrogen peroxide H2O2, which releases Fe ions (in the Fe2+ or Fe3+ oxidation state) that can cause direct damage to bacterial DNA, leading to its growth inhibition [101]. Table 1 describes the methods of synthesis and antibacterial effects of different types of metallic NPs.

4. Antifungal Activities of Metallic NPs

Fungal species are versatile in adaptation to any environmental conditions [126]. They have the capability of colonizing even in a medium having an extreme or precarious environment and this adaptability results in a variety of problems. Most agricultural practices are found to be affected by resistant fungal species [127]. All the important stages of crop cultivation, such as sowing, growth, production, and after-harvest stages, can be adversely affected. Nowadays chemical treatments are used to control the negative effects of pathogenic fungal species [128]. Despite the low cost and easy availability, the excessive use of chemical suppressors leads to a variety of problems, including environmental pollution, human disease, and ecological imbalance. Additionally, these can also cause resistance in fungal species, hence generating stronger fungal species that cannot be treated with chemical products [129]. Recently, environment-friendly alternatives have been introduced that are used to counter the negative effects of chemical remedies, but they still have considerable limitations. Plant extracts and essential oils can be used alternatively, but they can be processing intensive, and their high acquisition cost and persistent application make them a less attractive approach [130]. However, the incredible potential of NPs can be exploited against resistant fungal species [131] as they are proved effective and applicable as opposed to bacteria that are resistant as mentioned in the above section.

5. Mechanism of Action of Metallic NPs against Fungal Species

Metallic nanoparticles [MNPs] have various interactions with fungi that can result in advantageous or detrimental outcomes, depending on factors such as the NPs type, fungal species, and environmental conditions [132]. Many NPs have been reported to act in several ways against resistant fungal species. Exposure to NPs induces changes in the fungal cell wall, including surface alterations, cell aggregation, pit and pore formation, and overall deformation [133]. Studies have revealed that NPs may directly interact with and embed themselves within fungal cell walls during adsorption, resulting in morphological changes [134]. The inner membranes also undergo distortion, with altered organelle arrangement, such as an increased count of intracellular vesicles and vacuoles, and reduced cytoplasmic content, causing the release of cellular contents [135]. Smaller NPs may facilitate fluid-phase endocytosis, bypassing the need for significant cell wall damage. Exposure to NPs can result in alterations in gene expression and protein levels. Once NPs are inside the cell, some can intercalate with nucleic acids intracellularly [136]. Furthermore, some studies have indicated that ions are more toxic than their NP counterparts, possibly due to their size, which facilitates penetration into cells, or their ability to complex with other biomolecules such as proteins, nucleic acids, and negatively charged lipids [137]. NPs can profoundly impact fungal hyphae and spores. Exposure to NPs has been shown to deform hyphae, causing them to appear distorted and shrunken [138]. NPs alter growth patterns, leading to clumping and thinning of hyphal fibers. NPs can inhibit the formation of biofilms, as hyphae development is essential for biofilm formation and adherence, which are required for pathogenesis and colonization. The inhibition of filamentation is primarily driven by cell wall disruption. NPs can also affect pre-formed biofilms and deposit onto extracellular polysaccharides (EPS), crucial for structural integrity [139]. Additionally, MNPs can stimulate the production of ROS within fungal cells, leading to oxidative stress and cellular damage [140]. Figure 3 Showed the antifungal mechanism of NPs’.

5.1. Effect of AgNPs against Fungus

AgNPs are thoroughly studied in various scientific fields [141]. The antimicrobial, antioxidant, and anticancer properties of AgNPs, along with their low cost and ease of production, indeed make them appealing to therapeutic agents. However, there is some disagreement regarding their biocompatibility and toxicity. While AgNPs have been utilized and researched extensively against various microorganisms and fungi [142], previous studies showed that several AgNPs have reasonable activities against phytopathogenic fungi [143]. Ag+ ions and AgNPs can modify fungal cell transcriptomes, epigenomes, and metabolomes, leading to crucial functional alterations. This involves the down-regulation of genes related to the tricarboxylic acid cycle, redox metabolism, ergosterol synthesis, and lipid metabolism, ultimately causing structural modifications, primarily in fungal cell membranes [144]. AgNPs have been shown to have powerful antifungal properties based on their structural characteristics [145]. Mostly spherically and small-sized NPs are found to have a potential role against the different phytopathogenic fungal strains [146]. Sizes ranging from 10 to 30 nm have been found to have relatively effective antifungal activities [147]. Due to their small size, AgNPs can easily penetrate cell membranes, and their toxicity is partly linked to the production of ROS. This leads to the unification of fungal hyphae and mycelium, effectively deactivating these pathogens [148]. Alternatively, 40 to 70 nm NPs showed potent inhibitory activities by destroying mycelium and fungal spores, resulting in significant membrane rupture [147]. The concentration of NPs applied is a critical factor in fungal-NP interactions, with a significant impact on fungal strains. AgNPs attach to the fungal surface via electrostatic attraction. As AgNPs accumulate outside the cell, they release Ag+ ions, which enter the cell and neutralize or incapacitate these pathogens [149,150]. To determine the optimal concentration for effective antifungal activity, previous studies have explored various concentrations [151,152]. Surprisingly, lower concentrations have often demonstrated greater potency compared to higher concentrations. For example, AgNPs at a concentration of 20 ppm, produced from extracts of Psidium guajava and Momordica charantia, inhibited the growth of fungal strains including Fusarium oxysporum, Aspergillus flavus, and Aspergillus niger [153]. Similarly, research revealed that a 25-ppm concentration of AgNPs synthesized from extracts of Trichoderma viride completely inhibited the growth of Alternaria solani [154]. In another study, different concentrations of AgNPs (10, 25, 50, and 100 ppm) synthesized from green and black tea were tested against Aspergillus parasiticus, with maximum inhibition noted at a concentration of 100 ppm [155].

5.2. Effect of Copper NPs against Fungus

Copper nanoparticles (CuNPs) are known for their substantial antiseptic properties and cost-effectiveness [156]. CuNPs are employed as antimicrobial agents, benefiting from their substantial surface-to-volume ratio, which enables interactions with other particles, enhancing antimicrobial efficacy. Frequently, they are protected by polymers or surfactants to prevent oxidation. Chitosan-coupled CuNPs represent a highly promising nanocomposite, demonstrating remarkable antifungal activity against phytopathogens [157,158]. CuNPs function as fungicides by generating highly reactive hydroxyl radicals (•OH), which have the potential to inflict damage on biological macromolecules like the DNA of fungal pathogens [159]. While numerous studies have highlighted the potential of Cu NPs as effective antifungal agents, compared to research on other MNPs, there is a limited number of studies elucidating the antimicrobial mechanism of CuNPs. It has been suggested that the antimicrobial mechanism of CuNPs shares similarities with that of AgNPs; further research is required to fully elucidate their precise antifungal mechanisms and potential applications [160,161]. The efficacy of CuNPs against microbes depends significantly on various factors, including particle size, morphology, and concentration [162]. There is a huge diversity in size and antifungal activity. The variety of sizes of NPs gives a different extent of antifungal activity, making it difficult to evaluate the exact size of the particles that is the most effective in action [41]. Small size NPs may breach the cellular membrane, causing the leakage of the cellular contents [163]. In terms of shape, mostly spherical NPs have the most potent antifungal properties [164]. Other shapes that are found to show activities against the fungal spores are truncated octahedral, which is highly effective against Fusarium oxysporum, Fusarium solani, and Neofusicoccum sp. The faceted shape was also found to have reasonable activities against Fusarium oxysporum, and Fusarium solani [147]. To evaluate the optimum concentration for the CuNPs, low, medium, and high concentrations were applied to the population of phytopathogenic fungi. To check the effect, lower concentrations of 0.1, 0.25, and 0.5 ppm of the CuNPs were applied. The 0.1 ppm concentration appeared to promote hard oxidative stress inside the mycelium, while 0.5 ppm concentration was found to have antifungal activities against Fusarium oxysporum [165]. In medium concentrations, 5, 10, and 20 ppm of CuNPs were applied to the culture of the Phytopthora capsici and Fusarium oxysporum. On the third day of the application, antifungal activities were noted. The 5-ppm concentration applied against the fungi showed 49% inhibition of the culture, while 20 ppm of CuNPs was found to prevent 63% of the phytopathogenic fungus [161]. Another study applied CuNPs at 35, 25, 15, and 5 ppm concentrations against Sparassis crispa, Phytophthora cactorum, Grifola frondose, Megaloceros giganteus, Fusarium redolens, Fasciola hepatica, and Megaloceros giganteus. Among the applied concentrations, the 35 ppm concentration was found to be the most potent that was able to completely inhibit the development of plant pathogenic fungi [166]. The maximum concentrations of 300, 380, and 450 ppm showed excellent antifungal activities against Fusarium oxysporum with maximum antifungal activity noted at 450 ppm concentration [167]. In another study, 50, 100, 500, and 1000 ppm concentrations were applied against Alternaria alternata, Botrytis cinerea, Colletotrichum gloeosporioides, Monilinia fructicola, Fusarium solani, Fusarium oxysporum, and Verticillium dahlia. All the phytopathogenic fungi were found to be inhibited at 1000 ppm concentration of CuNPs [168].

5.3. Effect of Selenium NPs against Fungus

Selenium nanoparticles (SeNPs) possess broad biomedical applications, serving as antibacterial, antifungal, antioxidant, and anticancer agents, and their biologically synthesized variants demonstrate enhanced compatibility with human tissues. With their size, shape, and synthesis methods being actively investigated for their utility in biological systems, owing to their biocompatibility, low toxicity, and high bioavailability, which render them increasingly valuable in diverse biomedical contexts [169]. For example, biogenically synthesized SeNPs, produced by Ralstonia eutropha with a size range of 40–120 nm, exhibited inhibitory effects on the growth of the fungus Aspergillus clavatus at a concentration of 500 µg/mL [170]. SeNPs, produced through Bacillus thuringiensis with an average size of 50 to 200 nm, exhibited antifungal properties against Malassezia and Aspergillus by inhibiting spore germination [171]. Studies have evaluated the fungicidal activity of SeNPs synthesized within a size range varying from 50 to 400 nm. These NPs were employed to prevent the formation of Candida albicans biofilms. SeNPs exhibited a strong adherence to biofilm, enabling penetration into the pathogenic agents and causing structural damage through sulfur substitution [172]. Trichoderma-mediated SeNPs were tested against Sclerospora graminicola in doses ranging from 0 to 1000 ppm. Six different strains of Trichoderma spp., including Trichoderma harzianum, Trichoderma virens, Trichoderma asperellum, Trichoderma longibrachiatum, and Trichoderma brevicompactum were used. Trichoderma asperellum showed the efficient synthesis of SeNPs in the form of culture filtrate in the context of fungicidal capacity [173]. SeNPs synthesized through Trichoderma viride following a biological method were applied in vitro at various concentrations (50, 100, 200, 300, 400, 500, 600, 700, and 800 ppm) on the plant, and then treated with Alternaria solani. Results demonstrated that fungal growth was inhibited by SeNPs at 800 ppm [154]. Yet, inanother study, different concentrations of chemically synthesized SeNPs were evaluated, including 0.1, 0.5, 1, 5, 10, 50, and 100 ppm, against Macrophomina phaseolina, Sclerotinia sclerotiorum, and Diaporthe longicolla. At 10 ppm and above, SeNPs inhibited Diaporthe longicolla, and at 50 and 100 ppm they showed activity against Macrophomina phaseolina. Sclerotinia sclerotiorum grew and developed unhindered under different concentrations of SeNPs [152].

5.4. Effect of FeNPs against Fungus

FeNPs have the potential to induce microbial toxicity through a series of interactions, including membrane depolarization, which compromises cell integrity [174], the generation of ROS resulting in lipid peroxidation and DNA damage, and the release of metal ions that disrupt cellular homeostasis and protein coordination [175]. Due to their unique properties, such as biocompatibility, stability, and magnetic characteristics, biomedical fields are increasingly interested in FeNPs, making them promising candidates for applications in antibacterial, antifungal, and anticancer contexts [176,177]. FeNPs were found to inhibit spore germination, substantially reduce mycelium proliferation, and limit oxygen supply for respiration due to their higher surface-to-volume ratio, resulting in the complete coating of fungal microorganisms [178]. FeNPs can be synthesized using plant extracts, enabling the production of a large quantity of antimicrobial agents. As an example, the sensitivity of phytopathogenic fungi to FeNPs synthesized using green and black tea leaves was tested. Various concentrations of FeNPs at 10, 25, 50, and 100 ppm were used against fungi Aspergillus flavus and Aspergillus parasiticus in vitro. The results showed inhibition activity of 43.5% with FeNPs synthesized in green tea leaf extract and 51.6% inhibition activity with those synthesized in black tea leaf extract both at a dose of 100 ppm [155]. In another study, Fe2O3 particles with sizes ranging from 10 to 30 nm using a green approach were tested for their fungicide efficacy in opposition to Alternaria alternata, Cladosporium herbarum, Trichothecium roseum, Penicillium chrysogenum, and Aspergillus niger. It was noted that Fe2O3 significantly reduced the growth of all the fungal infections that were examined. Trichotherma roseum and Cladosporium herbarum were determined to be inhibited by 87.74% and 84.89% through the highest levels of spore germination inhibition. Penicillium chrysogenum had the maximum inhibitory zone (28.67 mm) caused by iron oxide NPs, followed by the Aspergillus niger (26.33 mm), Trichotherma roseum (22.67 mm), Alternaria alternata (21.33 mm), and Cladosporium herbarum (18.00 mm) [178]. Table 2 describes methods of synthesis and antifungal effects of different types of metallic NPs.

6. NMs with Antibiotics against Antimicrobial Resistance

In conjunction with antibiotics, NMs display a high degree of antibacterial activity to prevent bacteria from displaying tolerance to different antibiotics [43]. This concludes that NPs synthesis represents the best cure for enhanced bacterial antibiotic resistance [191]. NMs, combined with antibiotics or other antimicrobial agents, can overcome the limitations in their antibacterial potentials when they are used separately. The antibacterial effects of NMs can be improved by combining different antimicrobial agents with them, facilitating their intracellular targeting and improving their drug stabilization [192]. For example, there were combined effects of cephalexin antibiotics plus AgNPs for their improved antibiotic activity and antibacterial activity against S. aureus [193]. Another study showed that enhanced antibacterial activity in AgNPs conjugated with streptomycin was reported for S. aureus and E. coli bacteria [194]. The antimicrobial potential of ZnONP has been well explored in recent decades, although very limited literature is available on its synergistic effects with antibiotics. In the investigation of ZnO-NP, 9AA-HCl, and their conjugates’ effectiveness against E. coli, it was noted that the bacterial cell killing achieved by ZnO-NP-9AA-HCl was nearly 100%, a significant improvement compared to their individual use [195]. Microdilution was employed to assess the minimal inhibitory concentration (MIC) for ZnO NPs and various antibiotics (meropenem, ciprofloxacin, and colistin). The MIC values ranged from 2000 to 4000 μg/mL against P. aeruginosa when considering the combination of antibiotics and ZnO NPs conjugation [196].
An enhancement in the stability, selectivity, or functionality of antibiotics is the major benefit when they get attached to NPs [197]. Conjugating antibiotics with AuNPs is one of the strategies to improve the potency of the current antibiotic treatment [198]. Also, the conjugated NPs will target the drug in a way that other systematic compounds will not [199]. Different studies reveal that when antibiotics are conjugated with AuNPs, they show enhanced antimicrobial activity compared to that of antibiotics alone [200]. According to the report, antibiotics loaded onto Au-NPs exhibit greater efficacy against both Gram-positive and Gram-negative bacteria compared to the same dose of antibiotics used alone [201]. Loading of drugs to NPs has been done for many antibiotics like ciprofloxacin, neomycin, 5-fluorouracil (anticancer compound), ampicillin, kanamycin, gentamycin, and streptomycin. In a study, vancomycin was conjugated to Au-NPs for the destruction of bacteria resistant to vancomycin, i.e., Enterococcus faecium and Enterococcus faecalis. The results of the study have shown a 50-fold increase in the antibacterial activity of vancomycin [202]. FeNPs are less toxic and have many applications in the formation of bioproducts [203]. When antibiotics are attached to FeNPs, their constancy, and capabilities increase [204]. FeNPs also act as nanovehicles for carrying antibiotics because iron is very important for bacterial cell viability. For example, when FeNPs were combined with erythromycin, researchers observed enhanced antibacterial activity of erythromycin against Streptococcus pneumonia. This indicated that FeNPs acted as nanovehicles for erythromycin delivery [205]. Combining FeNPs with antibiotics reduced the required drug dose [206]. FeNPs conjugates also reduce toxicity of the NPs by enhancing intracellular targeting ability. The inhibitory action of FeNPs and cephalexin conjugation against several types of multidrug-resistant bacteria (E. coli, Bacillus sp., S. aureus, and Salmonella sp.) showed that FeNPs conjugated with antibiotics gave zone of inhibition greater than cephalexin alone [207].

7. Current Challenges and Future Perspectives

The role of nanotechnology in different areas such as medicine, vaccination, diagnostics, etc. is noticeable [208]. However, the potential of antibacterial and antifungal applications of NMs is currently limited by certain challenges [209]. If the life science research communities manage to overcome these challenges, the applicability and effectiveness of NMs could then help efficiently restore the lost activity of antimicrobials. The potential toxicity of NMs is one of the biggest challenges nanotechnologies are currently facing [210]. NMs are not just potentially toxic to humans but also to the environment. As most NMs are metallic, unfortunately, their toxicity is not well understood [211]. Metallic NMs are very carcinogenic and therefore, using them in immunodeficient patients can result in adverse effects. The possible accumulation of NMs in patients’ bodies can result in human health compromise [212,213]. Therefore, the toxicity of NMs needs to be minimized. NMs, especially physically and chemically synthesized NMs, have adverse environmental effects [11]. The inappropriate disposal of NMs can cause various forms of environmental pollution as they are not easily degraded. Therefore, these NMs get suspended in the air and can travel long distances [214]. When NMs encounter the body, they easily penetrate the skin due to their nanoscale sizes [215]. NMs have harmful effects on air, soil, and groundwater [216]. There is limited literature available regarding the potential risks associated with NMs. The use of NMs that are not well-understood could potentially lead to adverse health effects when employed for therapeutic purposes in immunodeficient patients [217]. This challenge can be addressed through global knowledge-sharing and collaborative efforts. Researchers, nanotechnology experts, and various research organizations from around the world can collectively focus on advancing this research field, acknowledging that it is still in its developmental stages. This collaborative approach holds the potential to enhance the antimicrobial effectiveness of NMs without adverse consequences.

8. Conclusions

The widespread use of antimicrobials has given rise to a challenging global problem: the development of antimicrobial-resistant pathogens, which has become a worldwide issue. Although various strategies are available to combat these AMR pathogens, the problem persists due to their labor-intensive nature and the significant environmental concerns associated with their implementation. Nanotechnology offers a promising alternative to these conventional strategies, presenting a distinct advantage. Two major distinct strategies for NP synthesis exist: green synthesis and chemical synthesis. Among these, green synthesis stands out as a better alternative that addresses environmental concerns. This type of NP synthesis involves plants, which are integral parts of the environment, providing a strong basis for mitigating the environmental concerns associated with chemical strategies. Furthermore, the production of NPs using plants is less labor-intensive and requires a smaller workforce. Another crucial aspect of this alternative is its superiority over other conventional strategies for tackling AMR. NPs demonstrate remarkable effectiveness against AMR pathogens, irrespective of the medium, their structural characteristics, colony size, or other defensive mechanisms that microbes develop over time and exposure to specific treatments. Notably, microbes cannot develop mechanisms to counteract the action of NPs. Nanotechnology has emerged as a viable solution to the global challenge of antibacterial and antifungal resistance. Nevertheless, further experimental support and systematic clinical trials are required to fully elucidate the precise mechanism underlying the antimicrobial characteristics of MNPs.

Author Contributions

S.W.: conceptualization, methodology, formal analysis, literature review, writing-original draft. A.S.: methodology, literature review, writing and editing. Z.K.: methodology, literature review, writing and editing. S.K.: methodology, literature review, writing and editing. C.K.: writing-review& editing, supervision. S.-I.Y.: writing review & editing, supervision, project administration, funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Basic Science Research Program through National Research Foundation of Korea, NRF-2021R1A2C1094316.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Authors acknowledge funding support from Basic Science Research Program through the National Research Foundation of Korea, NRF for successful execution of this review work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fauci, A.S. Infectious Diseases: Considerations for the 21st Century. Clin. Infect. Dis. 2001, 32, 675–685. [Google Scholar] [CrossRef] [PubMed]
  2. Dhingra, S.; Rahman, N.A.A.; Peile, E.; Rahman, M.; Sartelli, M.; Hassali, M.A.; Islam, T.; Islam, S.; Haque, M. Microbial Resistance Movements: An Overview of Global Public Health Threats Posed by Antimicrobial Resistance, and How Best to Counter. Front. Public Health 2020, 8, 535668. [Google Scholar] [CrossRef]
  3. Worthington, R.J.; Melander, C. Combination approaches to combat multidrug-resistant bacteria. Trends Biotechnol. 2013, 31, 177–184. [Google Scholar] [CrossRef] [PubMed]
  4. Morrill, H.J.; Caffrey, A.R.; Jump, R.L.; Dosa, D.; LaPlante, K.L. Antimicrobial Stewardship in Long-Term Care Facilities: A Call to Action. J. Am. Med. Dir. Assoc. 2016, 17, 183.e1–183.e16. [Google Scholar] [CrossRef]
  5. Coates, A.; Hu, Y.; Bax, R.; Page, C. The future challenges facing the development of new antimicrobial drugs. Nat. Rev. Drug Discov. 2002, 1, 895–910. [Google Scholar] [CrossRef] [PubMed]
  6. Ribeiro, C.F.A.; Silveira, G.G.d.O.S.; Cândido, E.d.S.; Cardoso, M.H.; Carvalho, C.M.E.; Franco, O.L. Effects of Antibiotic Treatment on Gut Microbiota and How to Overcome Its Negative Impacts on Human Health. ACS Infect. Dis. 2020, 6, 2544–2559. [Google Scholar] [CrossRef] [PubMed]
  7. Michael, C.A.; Dominey-Howes, D.; Labbate, M. The Antimicrobial Resistance Crisis: Causes, Consequences, and Management. Front. Public Health 2014, 2, 145. [Google Scholar] [CrossRef] [PubMed]
  8. Cunningham, M.; Azcarate-Peril, M.A.; Barnard, A.; Benoit, V.; Grimaldi, R.; Guyonnet, D.; Holscher, H.D.; Hunter, K.; Manurung, S.; Obis, D.; et al. Shaping the Future of Probiotics and Prebiotics. Trends Microbiol. 2021, 29, 667–685. [Google Scholar] [CrossRef]
  9. de la Fuente-Núñez, C.; Reffuveille, F.; Fernández, L.; Hancock, R.E. Bacterial biofilm development as a multicellular adaptation: Antibiotic resistance and new therapeutic strategies. Curr. Opin. Microbiol. 2013, 16, 580–589. [Google Scholar] [CrossRef] [PubMed]
  10. Beyth, N.; Houri-Haddad, Y.; Domb, A.; Khan, W.; Hazan, R. Alternative Antimicrobial Approach: Nano-Antimicrobial Materials. Evid.-Based Complement. Altern. Med. 2015, 2015, 246012. [Google Scholar] [CrossRef]
  11. Saleh, T.A. Nanomaterials: Classification, properties, and environmental toxicities. Environ. Technol. Innov. 2020, 20, 101067. [Google Scholar] [CrossRef]
  12. Kalhapure, R.S.; Suleman, N.; Mocktar, C.; Seedat, N.; Govender, T. Nanoengineered Drug Delivery Systems for Enhancing Antibiotic Therapy. J. Pharm. Sci. 2015, 104, 872–905. [Google Scholar] [CrossRef] [PubMed]
  13. Singh, A.; Gautam, P.K.; Verma, A.; Singh, V.; Shivapriya, P.M.; Shivalkar, S.; Sahoo, A.K.; Samanta, S.K. Green synthesis of metallic nanoparticles as effective alternatives to treat antibiotics resistant bacterial infections: A review. Biotechnol. Rep. 2020, 25, e00427. [Google Scholar] [CrossRef] [PubMed]
  14. Das, R.; Kotra, K.; Singh, P.; Loh, B.; Leptihn, S.; Bajpai, U. Alternative Treatment Strategies for Secondary Bacterial and Fungal Infections Associated with COVID-19. Infect. Dis. Ther. 2022, 11, 79–80. [Google Scholar] [CrossRef] [PubMed]
  15. Wahab, S.; Khan, T.; Adil, M.; Khan, A. Mechanistic aspects of plant-based silver nanoparticles against multi-drug resistant bacteria. Heliyon 2021, 7, e07448. [Google Scholar] [CrossRef] [PubMed]
  16. Jamkhande, P.G.; Ghule, N.W.; Bamer, A.H.; Kalaskar, M.G. Metal nanoparticles synthesis: An overview on methods of preparation, advantages and disadvantages, and applications. J. Drug Deliv. Sci. Technol. 2019, 53, 101174. [Google Scholar] [CrossRef]
  17. Ahmed, S.F.; Mofijur, M.; Rafa, N.; Chowdhury, A.T.; Chowdhury, S.; Nahrin, M.; Islam, A.S.; Ong, H.C. Green approaches in synthesising nanomaterials for environmental nanobioremediation: Technological advancements, applications, benefits and challenges. Environ. Res. 2021, 204, 111967. [Google Scholar] [CrossRef] [PubMed]
  18. Folorunso, A.; Akintelu, S.; Oyebamiji, A.K.; Ajayi, S.; Abiola, B.; Abdusalam, I.; Morakinyo, A. Biosynthesis, characterization and antimicrobial activity of gold nanoparticles from leaf extracts of Annona muricata. J. Nanostruct. Chem. 2019, 9, 111–117. [Google Scholar] [CrossRef]
  19. Yang, Y.; Waterhouse, G.I.; Chen, Y.; Sun-Waterhouse, D.; Li, D. Microbial-enabled green biosynthesis of nanomaterials: Current status and future prospects. Biotechnol. Adv. 2022, 55, 107914. [Google Scholar] [CrossRef] [PubMed]
  20. Hu, X.; Zhang, H.; Wang, Y.; Shiu, B.-C.; Lin, J.-H.; Zhang, S.; Lou, C.-W.; Li, T.-T. Synergistic antibacterial strategy based on photodynamic therapy: Progress and perspectives. Chem. Eng. J. 2022, 450, 138129. [Google Scholar] [CrossRef]
  21. Chakraborty, N.; Jha, D.; Roy, I.; Kumar, P.; Gaurav, S.S.; Marimuthu, K.; Ng, O.-T.; Lakshminarayanan, R.; Verma, N.K.; Gautam, H.K. Nanobiotics against antimicrobial resistance: Harnessing the power of nanoscale materials and technologies. J. Nanobiotechnol. 2022, 20, 1–25. [Google Scholar] [CrossRef] [PubMed]
  22. Amin, M.T.; El Aty, M.A.A.; Ahmed, S.M.; Elsedfy, G.O.; Hassanin, E.S.; El-Gazzar, A.F. Over prescription of antibiotics in children with acute upper respiratory tract infections: A study on the knowledge, attitude and practices of non-specialized physicians in Egypt. PLoS ONE 2022, 17, e0277308. [Google Scholar] [CrossRef]
  23. Qadri, H.; Shah, A.H.; Ahmad, S.M.; Alshehri, B.; Almilaibary, A.; Mir, M.A. Natural products and their semi-synthetic derivatives against antimicrobial-resistant human pathogenic bacteria and fungi. Saudi J. Biol. Sci. 2022, 29, 103376. [Google Scholar] [CrossRef] [PubMed]
  24. Caruso, D.J.; Palombo, E.A.; Moulton, S.E.; Zaferanloo, B. Exploring the Promise of Endophytic Fungi: A Review of Novel Antimicrobial Compounds. Microorganisms 2022, 10, 1990. [Google Scholar] [CrossRef] [PubMed]
  25. Haider, A.; Ikram, M.; Rafiq, A. Green Nanomaterials as Potential Antimicrobials; Springer: Berlin/Heidelberg, Germany, 2023. [Google Scholar] [CrossRef]
  26. Uruén, C.; Chopo-Escuin, G.; Tommassen, J.; Mainar-Jaime, R.C.; Arenas, J. Biofilms as Promoters of Bacterial Antibiotic Resistance and Tolerance. Antibiotics 2020, 10, 3. [Google Scholar] [CrossRef]
  27. Palma, E.; Tilocca, B.; Roncada, P. Antimicrobial Resistance in Veterinary Medicine: An Overview. Int. J. Mol. Sci. 2020, 21, 1914. [Google Scholar] [CrossRef]
  28. Varela, M.F.; Stephen, J.; Lekshmi, M.; Ojha, M.; Wenzel, N.; Sanford, L.M.; Hernandez, A.J.; Parvathi, A.; Kumar, S.H. Bacterial Resistance to Antimicrobial Agents. Antibiotics 2021, 10, 593. [Google Scholar] [CrossRef]
  29. Ebbensgaard, A.E.; Løbner-Olesen, A.; Frimodt-Møller, J. The Role of Efflux Pumps in the Transition from Low-Level to Clinical Antibiotic Resistance. Antibiotics 2020, 9, 855. [Google Scholar] [CrossRef]
  30. Wright, G.D. Antibiotic adjuvants: Rescuing antibiotics from resistance. Trends Microbiol. 2016, 24, 862–871. [Google Scholar] [CrossRef]
  31. Guitor, A.K.; Wright, G.D. Antimicrobial Resistance and Respiratory Infections. Chest 2018, 154, 1202–1212. [Google Scholar] [CrossRef]
  32. Ciofu, O.; Tolker-Nielsen, T.; Jensen, P.; Wang, H.; Høiby, N. Antimicrobial resistance, respiratory tract infections and role of biofilms in lung infections in cystic fibrosis patients. Adv. Drug Deliv. Rev. 2015, 85, 7–23. [Google Scholar] [CrossRef] [PubMed]
  33. Huh, A.J.; Kwon, Y.J. “Nanoantibiotics”: A new paradigm for treating infectious diseases using nanomaterials in the antibiotics resistant era. J. Control. Release 2011, 156, 128–145. [Google Scholar] [CrossRef] [PubMed]
  34. Vijayaraghavan, K.; Nalini, S.K.; Prakash, N.U.; Madhankumar, D. One step green synthesis of silver nano/microparticles using extracts of Trachyspermum ammi and Papaver somniferum. Colloids Surf. B Biointerfaces 2012, 94, 114–117. [Google Scholar] [CrossRef] [PubMed]
  35. Makabenta, J.M.V.; Nabawy, A.; Li, C.-H.; Schmidt-Malan, S.; Patel, R.; Rotello, V.M. Nanomaterial-based therapeutics for antibiotic-resistant bacterial infections. Nat. Rev. Microbiol. 2020, 19, 23–36. [Google Scholar] [CrossRef] [PubMed]
  36. Joshi, A.S.; Singh, P.; Mijakovic, I. Interactions of Gold and Silver Nanoparticles with Bacterial Biofilms: Molecular Interactions behind Inhibition and Resistance. Int. J. Mol. Sci. 2020, 21, 7658. [Google Scholar] [CrossRef]
  37. 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]
  38. Ruddaraju, L.K.; Pammi, S.V.N.; Guntuku, G.S.; Padavala, V.S.; Kolapalli, V.R.M. A review on anti-bacterials to combat resistance: From ancient era of plants and metals to present and future perspectives of green nano technological combinations. Asian J. Pharm. Sci. 2019, 15, 42–59. [Google Scholar] [CrossRef]
  39. Dharmaraja, A.T. Role of Reactive Oxygen Species (ROS) in Therapeutics and Drug Resistance in Cancer and Bacteria. J. Med. Chem. 2017, 60, 3221–3240. [Google Scholar] [CrossRef]
  40. Kessler, A.; Hedberg, J.; Blomberg, E.; Odnevall, I. Reactive Oxygen Species Formed by Metal and Metal Oxide Nanoparticles in Physiological Media—A Review of Reactions of Importance to Nanotoxicity and Proposal for Categorization. Nanomaterials 2022, 12, 1922. [Google Scholar] [CrossRef]
  41. Nisar, P.; Ali, N.; Rahman, L.; Ali, M.; Shinwari, Z.K. Antimicrobial activities of biologically synthesized metal nanoparticles: An insight into the mechanism of action. JBIC J. Biol. Inorg. Chem. 2019, 24, 929–941. [Google Scholar] [CrossRef]
  42. Kang, Y.; Liu, J.; Jiang, Y.; Yin, S.; Huang, Z.; Zhang, Y.; Wu, J.; Chen, L.; Shao, L. Understanding the interactions between inorganic-based nanomaterials and biological membranes. Adv. Drug Deliv. Rev. 2021, 175, 113820. [Google Scholar] [CrossRef] [PubMed]
  43. Gupta, A.; Mumtaz, S.; Li, C.-H.; Hussain, I.; Rotello, V.M. Combatting antibiotic-resistant bacteria using nanomaterials. Chem. Soc. Rev. 2019, 48, 415–427. [Google Scholar] [CrossRef] [PubMed]
  44. Lartigue, L.; Alloyeau, D.; Kolosnjaj-Tabi, J.; Javed, Y.; Guardia, P.; Riedinger, A.; Péchoux, C.; Pellegrino, T.; Wilhelm, C.; Gazeau, F. Biodegradation of Iron Oxide Nanocubes: High-Resolution In Situ Monitoring. ACS Nano 2013, 7, 3939–3952. [Google Scholar] [CrossRef] [PubMed]
  45. Lesniak, A.; Salvati, A.; Santos-Martinez, M.J.; Radomski, M.W.; Dawson, K.A.; Åberg, C. Nanoparticle Adhesion to the Cell Membrane and Its Effect on Nanoparticle Uptake Efficiency. J. Am. Chem. Soc. 2013, 135, 1438–1444. [Google Scholar] [CrossRef]
  46. Wang, T.; Rong, F.; Tang, Y.; Li, M.; Feng, T.; Zhou, Q.; Li, P.; Huang, W. Targeted polymer-based antibiotic delivery system: A promising option for treating bacterial infections via macromolecular approaches. Prog. Polym. Sci. 2021, 116, 101389. [Google Scholar] [CrossRef]
  47. Rohde, M.; Fischetti, V.A.; Novick, R.P.; Ferretti, J.J.; Portnoy, D.A.; Braunstein, M.; Rood, J.I. The Gram-Positive Bacterial Cell Wall; Wiley: Hoboken, NJ, USA, 2019; pp. 3–18. [Google Scholar] [CrossRef]
  48. 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. Nanomed. Nanotechnol. Biol. Med. 2010, 6, 103–109. [Google Scholar] [CrossRef]
  49. Yu, J.; Zhang, W.; Li, Y.; Wang, G.; Yang, L.; Jin, J.; Chen, Q.; Huang, M. Synthesis, characterization, antimicrobial activity and mechanism of a novel hydroxyapatite whisker/nano zinc oxide biomaterial. Biomed. Mater. 2014, 10, 015001. [Google Scholar] [CrossRef]
  50. Massironi, A.; Franco, A.R.; Babo, P.S.; Puppi, D.; Chiellini, F.; Reis, R.L.; Gomes, M.E. Development and Characterization of Highly Stable Silver NanoParticles as Novel Potential Antimicrobial Agents for Wound Healing Hydrogels. Int. J. Mol. Sci. 2022, 23, 2161. [Google Scholar] [CrossRef]
  51. Salleh, A.; Naomi, R.; Utami, N.D.; Mohammad, A.W.; Mahmoudi, E.; Mustafa, N.; Fauzi, M.B. The Potential of Silver Nanoparticles for Antiviral and Antibacterial Applications: A Mechanism of Action. Nanomaterials 2020, 10, 1566. [Google Scholar] [CrossRef]
  52. Lok, C.-N.; Ho, C.-M.; Chen, R.; He, Q.-Y.; Yu, W.-Y.; Sun, H.; Tam, P.K.-H.; Chiu, J.-F.; Che, C.-M. Proteomic Analysis of the Mode of Antibacterial Action of Silver Nanoparticles. J. Proteome Res. 2006, 5, 916–924. [Google Scholar] [CrossRef]
  53. Agnihotri, S.; Mukherji, S.; Mukherji, S. Immobilized silver nanoparticles enhance contact killing and show highest efficacy: Elucidation of the mechanism of bactericidal action of silver. Nanoscale 2013, 5, 7328–7340. [Google Scholar] [CrossRef] [PubMed]
  54. Krishnaraj, C.; Kaliannagounder, V.K.; Rajan, R.; Ramesh, T.; Kim, C.S.; Park, C.H.; Liu, B.; Yun, S.-I. Silver nanoparticles decorated reduced graphene oxide: Eco-friendly synthesis, characterization, biological activities and embryo toxicity studies. Environ. Res. 2022, 210, 112864. [Google Scholar] [CrossRef] [PubMed]
  55. Xu, Z.; Zhang, C.; Wang, X.; Liu, D. Release Strategies of Silver Ions from Materials for Bacterial Killing. ACS Appl. Bio Mater. 2021, 4, 3985–3999. [Google Scholar] [CrossRef] [PubMed]
  56. Padmanabhan, A.; Kaushik, M.; Niranjan, R.; Richards, J.S.; Ebright, B.; Venkatasubbu, G.D. Zinc oxide nanoparticles induce oxidative and proteotoxic stress in ovarian cancer cells and trigger apoptosis independent of p53-mutation status. Appl. Surf. Sci. 2019, 487, 807–818. [Google Scholar] [CrossRef]
  57. Bhunia, A.K.; Pradhan, S.S.; Bhunia, K.; Pradhan, A.K.; Saha, S. Study of the optical properties and frequency-dependent electrical modulus spectrum to the analysis of electric relaxation and conductivity effect in zinc oxide nanoparticles. J. Mater. Sci. Mater. Electron. 2021, 32, 22561–22578. [Google Scholar] [CrossRef]
  58. Gordon, T.; Perlstein, B.; Houbara, O.; Felner, I.; Banin, E.; Margel, S. Synthesis and characterization of zinc/iron oxide composite nanoparticles and their antibacterial properties. Colloids Surf. A Physicochem. Eng. Asp. 2010, 374, 1–8. [Google Scholar] [CrossRef]
  59. Ohira, T.; Yamamoto, O.; Iida, Y.; Nakagawa, Z.-E. Antibacterial activity of ZnO powder with crystallographic orientation. J. Mater. Sci. Mater. Med. 2007, 19, 1407–1412. [Google Scholar] [CrossRef] [PubMed]
  60. Talebian, N.; Amininezhad, S.M.; Doudi, M. Controllable synthesis of ZnO nanoparticles and their morphology-dependent antibacterial and optical properties. J. Photochem. Photobiol. B Biol. 2013, 120, 66–73. [Google Scholar] [CrossRef]
  61. Yang, H.; Liu, C.; Yang, D.; Zhang, H.; Xi, Z. Comparative study of cytotoxicity, oxidative stress and genotoxicity induced by four typical nanomaterials: The role of particle size, shape and composition. J. Appl. Toxicol. 2009, 29, 69–78. [Google Scholar] [CrossRef]
  62. Peng, X.; Palma, S.; Fisher, N.S.; Wong, S.S. Effect of morphology of ZnO nanostructures on their toxicity to marine algae. Aquat. Toxicol. 2011, 102, 186–196. [Google Scholar] [CrossRef]
  63. Sehmi, S.K.; Noimark, S.D.; Pike, S.; Bear, J.C.; Peveler, W.J.; Williams, C.K.; Shaffer, M.; Allan, E.; Parkin, I.P.; MacRobert, A.J. Enhancing the Antibacterial Activity of Light-Activated Surfaces Containing Crystal Violet and ZnO Nanoparticles: Investigation of Nanoparticle Size, Capping Ligand, and Dopants. ACS Omega 2016, 1, 334–343. [Google Scholar] [CrossRef] [PubMed]
  64. Dadi, R.; Azouani, R.; Traore, M.; Mielcarek, C.; Kanaev, A. Antibacterial activity of ZnO and CuO nanoparticles against gram positive and gram negative strains. Mater. Sci. Eng. C 2019, 104, 109968. [Google Scholar] [CrossRef]
  65. Elumalai, K.; Velmurugan, S. Green synthesis, characterization and antimicrobial activities of zinc oxide nanoparticles from the leaf extract of Azadirachta indica (L.). Appl. Surf. Sci. 2015, 345, 329–336. [Google Scholar] [CrossRef]
  66. Sukri, S.N.A.M.; Shameli, K.; Wong, M.M.-T.; Teow, S.-Y.; Chew, J.; Ismail, N.A. Cytotoxicity and antibacterial activities of plant-mediated synthesized zinc oxide (ZnO) nanoparticles using Punica granatum (pomegranate) fruit peels extract. J. Mol. Struct. 2019, 1189, 57–65. [Google Scholar] [CrossRef]
  67. Hussain, A.; Oves, M.; Alajmi, M.F.; Hussain, I.; Amir, S.; Ahmed, J.; Rehman, T.; El-Seedi, H.R.; Ali, I. Biogenesis of ZnO nanoparticles using Pandanus odorifer leaf extract: Anticancer and antimicrobial activities. RSC Adv. 2019, 9, 15357–15369. [Google Scholar] [CrossRef]
  68. Mahmud, S. One-dimensional growth of zinc oxide nanostructures from large micro-particles in a highly rapid synthesis. J. Alloys Compd. 2011, 509, 4035–4040. [Google Scholar] [CrossRef]
  69. Tariq, M.; Khan, A.U.; Rehman, A.U.; Ullah, S.; Jan, A.U.; Zakareya; Khan, Z.U.H.; Muhammad, N.; Islam, Z.U.; Yuan, Q. Green synthesis of Zno@GO nanocomposite and its’ efficient antibacterial activity. Photodiagnosis Photodyn. Ther. 2021, 35, 102471. [Google Scholar] [CrossRef]
  70. Smaoui, S.; Chérif, I.; Ben Hlima, H.; Khan, M.U.; Rebezov, M.; Thiruvengadam, M.; Sarkar, T.; Shariati, M.A.; Lorenzo, J.M. Zinc oxide nanoparticles in meat packaging: A systematic review of recent literature. Food Packag. Shelf Life 2023, 36, 101045. [Google Scholar] [CrossRef]
  71. Ahmed, B.; Dwivedi, S.; Abdin, M.Z.; Azam, A.; Al-Shaeri, M.; Khan, M.S.; Saquib, Q.; Al-Khedhairy, A.A.; Musarrat, J. Mitochondrial and Chromosomal Damage Induced by Oxidative Stress in Zn(2+) Ions, ZnO-Bulk and ZnO-NPs treated Allium cepa roots. Sci. Rep. 2017, 7, 40685. [Google Scholar] [CrossRef] [PubMed]
  72. Zare, M.; Namratha, K.; Ilyas, S.; Sultana, A.; Hezam, A.; L, S.; Surmeneva, M.A.; Surmenev, R.A.; Nayan, M.B.; Ramakrishna, S.; et al. Emerging Trends for ZnO Nanoparticles and Their Applications in Food Packaging. ACS Food Sci. Technol. 2022, 2, 763–781. [Google Scholar] [CrossRef]
  73. Lee, K.X.; Shameli, K.; Yew, Y.P.; Teow, S.-Y.; Jahangirian, H.; Rafiee-Moghaddam, R.; Webster, T.J. Recent Developments in the Facile Bio-Synthesis of Gold Nanoparticles (AuNPs) and Their Biomedical Applications. Int. J. Nanomed. 2020, 15, 275–300. [Google Scholar] [CrossRef] [PubMed]
  74. Santhosh, P.B.; Genova, J.; Chamati, H. Green Synthesis of Gold Nanoparticles: An Eco-Friendly Approach. Chemistry 2022, 4, 345–369. [Google Scholar] [CrossRef]
  75. Zhang, J.; Mou, L.; Jiang, X. Surface chemistry of gold nanoparticles for health-related applications. Chem. Sci. 2020, 11, 923–936. [Google Scholar] [CrossRef] [PubMed]
  76. Konvičková, Z.; Schröfel, A.; Kolenčík, M.; Dědková, K.; Peikertová, P.; Žídek, M.; Seidlerová, J.; Kratošová, G. Antimicrobial bionanocomposite—From precursors to the functional material in one simple step. J. Nanoparticle Res. 2016, 18, 368. [Google Scholar] [CrossRef]
  77. Sani, A.; Cao, C.; Cui, D. Toxicity of gold nanoparticles (AuNPs): A review. Biochem. Biophys. Rep. 2021, 26, 100991. [Google Scholar] [CrossRef]
  78. Al Hagbani, T.; Rizvi, S.M.D.; Hussain, T.; Mehmood, K.; Rafi, Z.; Moin, A.; Abu Lila, A.S.; Alshammari, F.; Khafagy, E.-S.; Rahamathulla, M.; et al. Cefotaxime Mediated Synthesis of Gold Nanoparticles: Characterization and Antibacterial Activity. Polymers 2022, 14, 771. [Google Scholar] [CrossRef]
  79. Srivastava, N.; Mukhopadhyay, M. Biosynthesis and Characterization of Gold Nanoparticles Using Zooglea ramigera and Assessment of Its Antibacterial Property. J. Clust. Sci. 2014, 26, 675–692. [Google Scholar] [CrossRef]
  80. Li, X.; Robinson, S.M.; Gupta, A.; Saha, K.; Jiang, Z.; Moyano, D.F.; Sahar, A.; Riley, M.A.; Rotello, V.M. Functional Gold Nanoparticles as Potent Antimicrobial Agents against Multi-Drug-Resistant Bacteria. ACS Nano 2014, 8, 10682–10686. [Google Scholar] [CrossRef]
  81. Zhou, Y.; Kong, Y.; Kundu, S.; Cirillo, J.D.; Liang, H. Antibacterial activities of gold and silver nanoparticles against Escherichia coli and bacillus Calmette-Guérin. J. Nanobiotechnol. 2012, 10, 19. [Google Scholar] [CrossRef]
  82. Wang, X.; Wang, X.; Bai, X.; Yan, L.; Liu, T.; Wang, M.; Song, Y.; Hu, G.; Gu, Z.; Miao, Q.; et al. Nanoparticle Ligand Exchange and Its Effects at the Nanoparticle–Cell Membrane Interface. Nano Lett. 2018, 19, 8–18. [Google Scholar] [CrossRef]
  83. Zhao, Y.; Tian, Y.; Cui, Y.; Liu, W.; Ma, W.; Jiang, X. Small Molecule-Capped Gold Nanoparticles as Potent Antibacterial Agents That Target Gram-Negative Bacteria. J. Am. Chem. Soc. 2010, 132, 12349–12356. [Google Scholar] [CrossRef] [PubMed]
  84. Shaikh, S.; Nazam, N.; Rizvi, S.M.D.; Ahmad, K.; Baig, M.H.; Lee, E.J.; Choi, I. Mechanistic Insights into the Antimicrobial Actions of Metallic Nanoparticles and Their Implications for Multidrug Resistance. Int. J. Mol. Sci. 2019, 20, 2468. [Google Scholar] [CrossRef]
  85. Lee, H.; Lee, D.G. Gold nanoparticles induce a reactive oxygen species-independent apoptotic pathway in Escherichia coli. Colloids Surf. B Biointerfaces 2018, 167, 1–7. [Google Scholar] [CrossRef] [PubMed]
  86. Mahmoud, N.N.; Alkilany, A.M.; Khalil, E.A.; Al-Bakri, A.G. Nano-Photothermal ablation effect of Hydrophilic and Hydrophobic Functionalized Gold Nanorods on Staphylococcus aureus and Propionibacterium acnes. Sci. Rep. 2018, 8, 6881. [Google Scholar] [CrossRef]
  87. Zhang, X.-F.; Shen, W.; Gurunathan, S. Biologically Synthesized Gold Nanoparticles Ameliorate Cold and Heat Stress-Induced Oxidative Stress in Escherichia coli. Molecules 2016, 21, 731. [Google Scholar] [CrossRef] [PubMed]
  88. Patra, J.K.; Baek, K.-H. Green biosynthesis of magnetic iron oxide (Fe3O4) nanoparticles using the aqueous extracts of food processing wastes under photo-catalyzed condition and investigation of their antimicrobial and antioxidant activity. J. Photochem. Photobiol. B Biol. 2017, 173, 291–300. [Google Scholar] [CrossRef]
  89. Kajani, A.A.; Bordbar, A.-K. Biogenic magnetite nanoparticles: A potent and environmentally benign agent for efficient removal of azo dyes and phenolic contaminants from water. J. Hazard. Mater. 2018, 366, 268–274. [Google Scholar] [CrossRef]
  90. Bahadar, H.; Maqbool, F.; Niaz, K.; Abdollahi, M. Toxicity of Nanoparticles and an Overview of Current Experimental Models. Iran. Biomed. J. 2016, 20, 1–11. [Google Scholar]
  91. Huang, Y.; Hsu, J.C.; Koo, H.; Cormode, D.P. Repurposing ferumoxytol: Diagnostic and therapeutic applications of an FDA-approved nanoparticle. Theranostics 2022, 12, 796–816. [Google Scholar] [CrossRef]
  92. Senthil, M.; Ramesh, C. Biostructures, Biogenic Synthesis Of Fe3O4 Nanoparticles Using Tridax Procumbens Leaf Extract And Its Antibacterial Activity On Pseudomonas aeruginosa. Dig. J. Nanomater. Biostruct. 2012, 7, 4. [Google Scholar]
  93. Alam, T.; Khan, R.A.A.; Ali, A.; Sher, H.; Ullah, Z.; Ali, M. Biogenic synthesis of iron oxide nanoparticles via Skimmia laureola and their antibacterial efficacy against bacterial wilt pathogen Ralstonia solanacearum. Mater. Sci. Eng. C 2018, 98, 101–108. [Google Scholar] [CrossRef] [PubMed]
  94. Jagathesan, G.; Rajiv, P. Biosynthesis and characterization of iron oxide nanoparticles using Eichhornia crassipes leaf extract and assessing their antibacterial activity. Biocatal. Agric. Biotechnol. 2018, 13, 90–94. [Google Scholar] [CrossRef]
  95. Naseem, T.; Farrukh, M.A. Antibacterial activity of green synthesis of iron nanoparticles using Lawsonia inermis and Gardenia jasminoides leaves extract. J. Chem. 2015, 2015, 912342. [Google Scholar] [CrossRef]
  96. Irshad, R.; Tahir, K.; Li, B.; Ahmad, A.; Siddiqui, A.R.; Nazir, S. Antibacterial activity of biochemically capped iron oxide nanoparticles: A view towards green chemistry. J. Photochem. Photobiol. B Biol. 2017, 170, 241–246. [Google Scholar] [CrossRef] [PubMed]
  97. Kamran, U.; Bhatti, H.N.; Iqbal, M.; Jamil, S.; Zahid, M. Biogenic synthesis, characterization and investigation of photocatalytic and antimicrobial activity of manganese nanoparticles synthesized from Cinnamomum verum bark extract. J. Mol. Struct. 2019, 1179, 532–539. [Google Scholar] [CrossRef]
  98. Sathiyanarayanan, G.; Dineshkumar, K.; Yang, Y.-H. Microbial exopolysaccharide-mediated synthesis and stabilization of metal nanoparticles. Crit. Rev. Microbiol. 2017, 43, 731–752. [Google Scholar] [CrossRef]
  99. Paunovic, J.; Vucevic, D.; Radosavljevic, T.; Mandić-Rajčević, S.; Pantic, I. Iron-based nanoparticles and their potential toxicity: Focus on oxidative stress and apoptosis. Chem. Interact. 2019, 316, 108935. [Google Scholar] [CrossRef]
  100. Godoy-Gallardo, M.; Eckhard, U.; Delgado, L.M.; de Roo Puente, Y.J.; Hoyos-Nogués, M.; Gil, F.J.; Perez, R.A. Antibacterial approaches in tissue engineering using metal ions and nanoparticles: From mechanisms to applications. Bioact. Mater. 2021, 6, 4470–4490. [Google Scholar] [CrossRef]
  101. Aribisala, J.O.; Sabiu, S. Redox Impact on Bacterial Macromolecule: A Promising Avenue for Discovery and Development of Novel Antibacterials. Biomolecules 2022, 12, 1545. [Google Scholar] [CrossRef]
  102. Bose, D.; Chatterjee, S. Antibacterial Activity of Green Synthesized Silver Nanoparticles Using Vasaka (Justicia adhatoda L.) Leaf Extract. Indian J. Microbiol. 2015, 55, 163–167. [Google Scholar] [CrossRef]
  103. Velmurugan, P.; Iydroose, M.; Lee, S.-M.; Cho, M.; Park, J.-H.; Balachandar, V.; Oh, B.-T. Synthesis of Silver and Gold Nanoparticles Using Cashew Nut Shell Liquid and Its Antibacterial Activity Against Fish Pathogens. Indian J. Microbiol. 2013, 54, 196–202. [Google Scholar] [CrossRef] [PubMed]
  104. Jyoti, K.; Baunthiyal, M.; Singh, A. Characterization of silver nanoparticles synthesized using Urtica dioica Linn. leaves and their synergistic effects with antibiotics. J. Radiat. Res. Appl. Sci. 2016, 9, 217–227. [Google Scholar] [CrossRef]
  105. Ibrahim, H.M. Green synthesis and characterization of silver nanoparticles using banana peel extract and their antimicrobial activity against representative microorganisms. J. Radiat. Res. Appl. Sci. 2015, 8, 265–275. [Google Scholar] [CrossRef]
  106. Geethalakshmi, R.; Sarada, D. Gold and silver nanoparticles from Trianthema decandra: Synthesis, characterization, and antimicrobial properties. Int. J. Nanomed. 2012, 7, 5375–5384. [Google Scholar] [CrossRef]
  107. Patra, J.K.; Baek, K.H. Novel green synthesis of gold nanoparticles using Citrullus lanatus rind and investigation of proteasome inhibitory activity, antibacterial, and antioxidant potential. Int. J. Nanomed. 2015, 10, 7253–7264. [Google Scholar] [CrossRef]
  108. Eskandari-Nojehdehi, M.; Jafarizadeh-Malmiri, H.; Rahbar-Shahrouzi, J. Optimization of processing parameters in green synthesis of gold nanoparticles using microwave and edible mushroom (Agaricus bisporus) extract and evaluation of their antibacterial activity. Nanotechnol. Rev. 2017, 5, 537–548. [Google Scholar] [CrossRef]
  109. Mata, R.; Bhaskaran, A.; Sadras, S.R. Green-synthesized gold nanoparticles from Plumeria alba flower extract to augment catalytic degradation of organic dyes and inhibit bacterial growth. Particuology 2016, 24, 78–86. [Google Scholar] [CrossRef]
  110. Islam, N.U.; Jalil, K.; Shahid, M.; Rauf, A.; Muhammad, N.; Khan, A.; Shah, M.R.; Khan, M.A. Green synthesis and biological activities of gold nanoparticles functionalized with Salix alba. Arab. J. Chem. 2015, 12, 2914–2925. [Google Scholar] [CrossRef]
  111. Piruthiviraj, P.; Margret, A.; Krishnamurthy, P.P. Gold nanoparticles synthesized by Brassica oleracea (Broccoli) acting as antimicrobial agents against human pathogenic bacteria and fungi. Appl. Nanosci. 2015, 6, 467–473. [Google Scholar] [CrossRef]
  112. Bala, N.; Saha, S.; Chakraborty, M.; Maiti, M.; Das, S.; Basu, R.; Nandy, P. Green synthesis of zinc oxide nanoparticles using Hibiscus subdariffa leaf extract: Effect of temperature on synthesis, anti-bacterial activity and anti-diabetic activity. RSC Adv. 2014, 5, 4993–5003. [Google Scholar] [CrossRef]
  113. Datta, A.; Patra, C.; Bharadwaj, H.; Kaur, S.; Dimri, N.; Khajuria, R. Green Synthesis of Zinc Oxide Nanoparticles Using Parthenium hysterophorus Leaf Extract and Evaluation of their Antibacterial Properties. J. Biotechnol. Biomater. 2017, 7, 271. [Google Scholar] [CrossRef]
  114. Senthilkumar, S.; Sivakumar, T. Green tea (Camellia sinensis) mediated synthesis of zinc oxide (ZnO) nanoparticles and studies on their antimicrobial activities. Int. J. Pharm. Pharm. Sci. 2014, 6, 461–465. [Google Scholar]
  115. Katata-Seru, L.; Moremedi, T.; Aremu, O.S.; Bahadur, I. Green synthesis of iron nanoparticles using Moringa oleifera extracts and their applications: Removal of nitrate from water and antibacterial activity against Escherichia coli. J. Mol. Liq. 2018, 256, 296–304. [Google Scholar] [CrossRef]
  116. Groiss, S.; Selvaraj, R.; Varadavenkatesan, T.; Vinayagam, R. Structural characterization, antibacterial and catalytic effect of iron oxide nanoparticles synthesised using the leaf extract of Cynometra ramiflora. J. Mol. Struct. 2017, 1128, 572–578. [Google Scholar] [CrossRef]
  117. Kanagasubbulakshmi, S.; Kadirvelu, K. Green synthesis of Iron oxide nanoparticles using Lagenaria siceraria and evaluation of its Antimicrobial activity. Def. Life Sci. J. 2017, 2, 422–427. [Google Scholar] [CrossRef]
  118. Burygin, G.; Khlebtsov, B.; Shantrokha, A.; Dykman, L.; Bogatyrev, V.; Khlebtsov, N. On the Enhanced Antibacterial Activity of Antibiotics Mixed with Gold Nanoparticles. Nanoscale Res. Lett. 2009, 4, 794–801. [Google Scholar] [CrossRef]
  119. MubarakAli, D.; Thajuddin, N.; Jeganathan, K.; Gunasekaran, M. Plant extract mediated synthesis of silver and gold nanoparticles and its antibacterial activity against clinically isolated pathogens. Colloids Surf. B Biointerfaces 2011, 85, 360–365. [Google Scholar] [CrossRef]
  120. Lee, B.; Lee, D.G. Synergistic antibacterial activity of gold nanoparticles caused by apoptosis-like death. J. Appl. Microbiol. 2019, 127, 701–712. [Google Scholar] [CrossRef]
  121. Ismail, R.A.; Sulaiman, G.M.; Abdulrahman, S.A.; Marzoog, T.R. Antibacterial activity of magnetic iron oxide nanoparticles synthesized by laser ablation in liquid. Mater. Sci. Eng. C 2015, 53, 286–297. [Google Scholar] [CrossRef]
  122. Ansari, S.A.; Oves, M.; Satar, R.; Khan, A.; Ahmad, S.I.; Jafri, M.A.; Zaidi, S.K.; Alqahtani, M.H. Antibacterial activity of iron oxide nanoparticles synthesized by co-precipitation technology against Bacillus cereus and Klebsiella pneumoniae. Pol. J. Chem. Technol. 2017, 19, 110–115. [Google Scholar] [CrossRef]
  123. Yadav, J.P.; Kumar, S.; Budhwar, L.; Yadav, A.; Yadav, M. Characterization and Antibacterial Activity of Synthesized Silver and Iron Nanoparticles using Aloe vera. J. Nanomed. Nanotechnol. 2016, 7, 3. [Google Scholar] [CrossRef]
  124. Jin, S.-E.; Jin, J.E.; Hwang, W.; Hong, S.W. Photocatalytic antibacterial application of zinc oxide nanoparticles and self-assembled networks under dual UV irradiation for enhanced disinfection. Int. J. Nanomed. 2019, 14, 1737–1751. [Google Scholar] [CrossRef] [PubMed]
  125. Salem, W.; Leitner, D.R.; Zingl, F.G.; Schratter, G.; Prassl, R.; Goessler, W.; Reidl, J.; Schild, S. Antibacterial activity of silver and zinc nanoparticles against Vibrio cholerae and enterotoxic Escherichia coli. Int. J. Med. Microbiol. 2014, 305, 85–95. [Google Scholar] [CrossRef]
  126. Coleine, C.; Stajich, J.E.; Selbmann, L. Fungi are key players in extreme ecosystems. Trends Ecol. Evol. 2022, 37, 517–528. [Google Scholar] [CrossRef]
  127. Sun, S.; Hoy, M.J.; Heitman, J. Fungal pathogens. Curr. Biol. 2020, 30, R1163–R1169. [Google Scholar] [CrossRef] [PubMed]
  128. Degani, O.; Kalman, B. Assessment of Commercial Fungicides against Onion (Allium cepa) Basal Rot Disease Caused by Fusarium oxysporum f. sp. cepae and Fusarium acutatum. J. Fungi 2021, 7, 235. [Google Scholar] [CrossRef]
  129. Baibakova, E.V.; Nefedjeva, E.E.; Suska-Malawska, M.; Wilk, M.; Sevriukova, G.A.; Zheltobriukhov, V.F. Modern Fungicides: Mechanisms of Action, Fungal Resistance and Phytotoxic Effects. Annu. Res. Rev. Biol. 2019, 1–16. [Google Scholar] [CrossRef]
  130. Abd-Elsalam, K.A.; Al-Dhabaan, F.A.; Alghuthaymi, M.; Njobeh, P.B.; Almoammar, H. Nanobiofungicides: Present Concept and Future Perspectives in Fungal Control; Academic Press: Cambridge, MA, USA, 2019; pp. 315–351. [Google Scholar] [CrossRef]
  131. Šebesta, M.; Vojtková, H.; Cyprichová, V.; Ingle, A.P.; Urík, M.; Kolenčík, M. Mycosynthesis of metal-containing nanoparticles—Fungal metal resistance and mechanisms of synthesis. Int. J. Mol. Sci. 2022, 23, 14084. [Google Scholar] [CrossRef]
  132. Priyadarshini, E.; Priyadarshini, S.S.; Cousins, B.G.; Pradhan, N. Metal-Fungus interaction: Review on cellular processes underlying heavy metal detoxification and synthesis of metal nanoparticles. Chemosphere 2021, 274, 129976. [Google Scholar] [CrossRef]
  133. Babele, P.K.; Thakre, P.K.; Kumawat, R.; Tomar, R.S. Zinc oxide nanoparticles induce toxicity by affecting cell wall integrity pathway, mitochondrial function and lipid homeostasis in Saccharomyces cerevisiae. Chemosphere 2018, 213, 65–75. [Google Scholar] [CrossRef]
  134. Selvaraj, M.; Pandurangan, P.; Ramasami, N.; Rajendran, S.B.; Sangilimuthu, S.N.; Perumal, P. Highly Potential Antifungal Activity of Quantum-Sized Silver Nanoparticles Against Candida albicans. Appl. Biochem. Biotechnol. 2014, 173, 55–66. [Google Scholar] [CrossRef] [PubMed]
  135. Slavin, Y.N.; Bach, H. Mechanisms of Antifungal Properties of Metal Nanoparticles. Nanomaterials 2022, 12, 4470. [Google Scholar] [CrossRef] [PubMed]
  136. Athie-García, M.S.; Piñón-Castillo, H.A.; Muñoz-Castellanos, L.N.; Ulloa-Ogaz, A.L.; Martínez-Varela, P.I.; Quintero-Ramos, A.; Duran, R.; Murillo-Ramirez, J.G.; Orrantia-Borunda, E. Cell wall damage and oxidative stress in Candida albicans ATCC10231 and Aspergillus niger caused by palladium nanoparticles. Toxicol. Vitr. 2018, 48, 111–120. [Google Scholar] [CrossRef]
  137. Khan, M.; Wahab, S.; Ali, H.M.; Khan, S.; Iqbal, R.; Khan, T. Biogenic Nanomaterials: A Way Forward in Preventing Bacterial Infections: Nanobiotechnology against resistant pathogens. Proc. Pak. Acad. Sci. B Life Environ. Sci. 2023, 60, 3–23. [Google Scholar]
  138. Ouda, S.M. Antifungal Activity of Silver and Copper Nanoparticles on Two Plant Pathogens, Alternaria alternata and Botrytis cinerea. Res. J. Microbiol. 2014, 9, 34–42. [Google Scholar] [CrossRef]
  139. Ramage, G.; Mowat, E.; Jones, B.; Williams, C.; Lopez-Ribot, J. Our Current Understanding of Fungal Biofilms. Crit. Rev. Microbiol. 2009, 35, 340–355. [Google Scholar] [CrossRef]
  140. Lee, Y.-H.; Cheng, F.-Y.; Chiu, H.-W.; Tsai, J.-C.; Fang, C.-Y.; Chen, C.-W.; Wang, Y.-J. Cytotoxicity, oxidative stress, apoptosis and the autophagic effects of silver nanoparticles in mouse embryonic fibroblasts. Biomaterials 2014, 35, 4706–4715. [Google Scholar] [CrossRef]
  141. Pryshchepa, O.; Pomastowski, P.; Buszewski, B. Silver nanoparticles: Synthesis, investigation techniques, and properties. Adv. Colloid Interface Sci. 2020, 284, 102246. [Google Scholar] [CrossRef]
  142. Akpinar, I.; Unal, M.; Sar, T. Potential antifungal effects of silver nanoparticles (AgNPs) of different sizes against phytopathogenic Fusarium oxysporum f. sp. radicis-lycopersici (FORL) strains. SN Appl. Sci. 2021, 3, 1–9. [Google Scholar] [CrossRef]
  143. Matras, E.; Gorczyca, A.; Przemieniecki, S.W.; Oćwieja, M. Surface properties-dependent antifungal activity of silver nanoparticles. Sci. Rep. 2022, 12, 18046. [Google Scholar] [CrossRef]
  144. Mussin, J.; Giusiano, G. Biogenic silver nanoparticles as antifungal agents. Front. Chem. 2022, 10, 1023542. [Google Scholar] [CrossRef]
  145. Gulbagca, F.; Ozdemir, S.; Gulcan, M.; Sen, F. Synthesis and characterization of Rosa canina-mediated biogenic silver nanoparticles for anti-oxidant, antibacterial, antifungal, and DNA cleavage activities. Heliyon 2019, 5, e02980. [Google Scholar] [CrossRef] [PubMed]
  146. Rizwana, H.; Alwhibi, M.S.; Al-Judaie, R.A.; Aldehaish, H.A.; Alsaggabi, N.S. Sunlight-Mediated Green Synthesis of Silver Nanoparticles Using the Berries of Ribes rubrum (Red Currants): Characterisation and Evaluation of Their Antifungal and Antibacterial Activities. Molecules 2022, 27, 2186. [Google Scholar] [CrossRef]
  147. Cruz-Luna, A.R.; Cruz-Martínez, H.; Vásquez-López, A.; Medina, D.I. Metal Nanoparticles as Novel Antifungal Agents for Sustainable Agriculture: Current Advances and Future Directions. J. Fungi 2021, 7, 1033. [Google Scholar] [CrossRef] [PubMed]
  148. Xiang, S.; Ma, X.; Shi, H.; Ma, T.; Tian, C.; Chen, Y.; Chen, H.; Chen, X.; Luo, K.; Cai, L.; et al. Green Synthesis of an Alginate-Coated Silver Nanoparticle Shows High Antifungal Activity by Enhancing Its Cell Membrane Penetrating Ability. ACS Appl. Bio Mater. 2019, 2, 4087–4096. [Google Scholar] [CrossRef]
  149. Mussin, J.E.; Roldán, M.V.; Rojas, F.; de los Ángeles Sosa, M.; Pellegri, N.; Giusiano, G. Antifungal activity of silver nanoparticles in combination with ketoconazole against Malassezia furfur. AMB Express 2019, 9, 131. [Google Scholar] [CrossRef] [PubMed]
  150. Shaik, M.R.; Khan, M.; Kuniyil, M.; Al-Warthan, A.; Alkhathlan, H.Z.; Siddiqui, M.R.H.; Shaik, J.P.; Ahamed, A.; Mahmood, A.; Khan, M.; et al. Plant-Extract-Assisted Green Synthesis of Silver Nanoparticles Using Origanum vulgare L. Extract and Their Microbicidal Activities. Sustainability 2018, 10, 913. [Google Scholar] [CrossRef]
  151. Prittesh, K.; Heena, B.; Rutvi, B.; Sangeeta, J.; Krunal, M. Synthesis and Characterisation of Silver Nanoparticles Using Withania somnifera and Antifungal Effect against Fusarium solani. Int. J. Plant Soil Sci. 2018, 25, 1–6. [Google Scholar] [CrossRef]
  152. Vrandečić, K.; Ćosić, J.; Ilić, J.; Ravnjak, B.; Selmani, A.; Galić, E.; Pem, B.; Barbir, R.; Vinković Vrček, I.; Vinković, T. Antifungal activities of silver and selenium nanoparticles stabilized with different surface coating agents. Pest Manag. Sci. 2020, 76, 2021–2029. [Google Scholar] [CrossRef]
  153. Nguyen, D.H.; Vo, T.N.N.; Nguyen, N.T.; Ching, Y.C.; Thi, T.T.H. Comparison of biogenic silver nanoparticles formed by Momordica charantia and Psidium guajava leaf extract and antifungal evaluation. PLoS ONE 2020, 15, e0239360. [Google Scholar] [CrossRef]
  154. Ismail, A.-W.A.; Sidkey, N.M.; Arafa, R.A.; Fathy, R.M.; El-Batal, A.I. Evaluation of in vitro antifungal activity of silver and selenium nanoparticles against Alternaria solani caused early blight disease on potato. Biotechnol. J. Int. 2016, 12, 1–11. [Google Scholar] [CrossRef]
  155. Zahir, E.; Shahid, S.M.; Khan, M.N.; Asghar, M.A.; Iqbal, J.; Walker, G. Iron, copper and silver nanoparticles: Green synthesis using green and black tea leaves extracts and evaluation of antibacterial, antifungal and aflatoxin B1 adsorption activity. LWT 2018, 90, 98–107. [Google Scholar] [CrossRef]
  156. Darwesh, O.M.; Li, H.; Matter, I.A. Nano-bioremediation of textile industry wastewater using immobilized CuO-NPs myco-synthesized by a novel Cu-resistant Fusarium oxysporum OSF18. Environ. Sci. Pollut. Res. Int. 2023, 30, 16694–16706. [Google Scholar] [CrossRef] [PubMed]
  157. Al-Dhabaan, F.A.; Shoala, T.; Ali, A.A.; Alaa, M.; Abd-Elsalam, K.; Abd-Elsalam, K. Chemically-produced copper, zinc nanoparticles and chitosan-bimetallic nanocomposites and their antifungal activity against three phytopathogenic fungi. Int. J. Agric. Technol. 2017, 13, 753–769. [Google Scholar]
  158. Vanti, G.L.; Masaphy, S.; Kurjogi, M.; Chakrasali, S.; Nargund, V.B. Synthesis and application of chitosan-copper nanoparticles on damping off causing plant pathogenic fungi. Int. J. Biol. Macromol. 2019, 156, 1387–1395. [Google Scholar] [CrossRef]
  159. Esteban-Tejeda, L.; Malpartida, F.; Esteban-Cubillo, A.; Pecharromán, C.; Moya, J.S. Antibacterial and antifungal activity of a soda-lime glass containing copper nanoparticles. Nanotechnology 2009, 20, 505701. [Google Scholar] [CrossRef]
  160. Khan, M.; Khan, A.A.; Parveen, A.; Min, K.; Yadav, V.K.; Khan, A.U.; Alam, M. Mitigating the growth of plant pathogenic bacterium, fungi, and nematode by using plant-mediated synthesis of copper oxide nanoparticles (CuO NPs). Green Chem. Lett. Rev. 2023, 16, 2177520. [Google Scholar] [CrossRef]
  161. Pham, N.-D.; Duong, M.-M.; Le, M.-V.; Hoang, H.A.; Pham, L.-K.-O. Preparation and characterization of antifungal colloidal copper nanoparticles and their antifungal activity against Fusarium oxysporum and Phytophthora capsici. Comptes Rendus Chim. 2019, 22, 786–793. [Google Scholar] [CrossRef]
  162. Verma, N.; Kumar, N. Synthesis and Biomedical Applications of Copper Oxide Nanoparticles: An Expanding Horizon. ACS Biomater. Sci. Eng. 2019, 5, 1170–1188. [Google Scholar] [CrossRef] [PubMed]
  163. Sánchez-López, E.; Gomes, D.; Esteruelas, G.; Bonilla, L.; Lopez-Machado, A.L.; Galindo, R.; Cano, A.; Espina, M.; Ettcheto, M.; Camins, A.; et al. Metal-Based Nanoparticles as Antimicrobial Agents: An Overview. Nanomaterials 2020, 10, 292. [Google Scholar] [CrossRef]
  164. Kuppusamy, P.; Ilavenil, S.; Srigopalram, S.; Maniam, G.P.; Yusoff, M.M.; Govindan, N.; Choi, K.C. Treating of palm oil mill effluent using Commelina nudiflora mediated copper nanoparticles as a novel bio-control agent. J. Clean. Prod. 2017, 141, 1023–1029. [Google Scholar] [CrossRef]
  165. Lopez-Lima, D.; Mtz-Enriquez, A.I.; Carrión, G.; Basurto-Cereceda, S.; Pariona, N. The bifunctional role of copper nanoparticles in tomato: Effective treatment for Fusarium wilt and plant growth promoter. Sci. Hortic. 2021, 277, 109810. [Google Scholar] [CrossRef]
  166. Aleksandrowicz-Trzcińska, M.; Szaniawski, A.; Olchowik, J.; Drozdowski, S. Effects of copper and silver nanoparticles on growth of selected species of pathogenic and wood-decay fungi in vitro. For. Chron. 2018, 94, 109–116. [Google Scholar] [CrossRef]
  167. Viet, P.V.; Nguyen, H.T.; Cao, T.M.; Hieu, L.V. Fusarium Antifungal Activities of Copper Nanoparticles Synthesized by a Chemical Reduction Method. J. Nanomater. 2016, 2016, 1957612. [Google Scholar] [CrossRef]
  168. Malandrakis, A.A.; Kavroulakis, N.; Chrysikopoulos, C.V. Use of copper, silver and zinc nanoparticles against foliar and soil-borne plant pathogens. Sci. Total Environ. 2019, 670, 292–299. [Google Scholar] [CrossRef]
  169. Bisht, N.; Phalswal, P.; Khanna, P.K. Selenium nanoparticles: A review on synthesis and biomedical applications. Mater. Adv. 2021, 3, 1415–1431. [Google Scholar] [CrossRef]
  170. Srivastava, N.; Mukhopadhyay, M. Green synthesis and structural characterization of selenium nanoparticles and assessment of their antimicrobial property. Bioprocess Biosyst. Eng. 2015, 38, 1723–1730. [Google Scholar] [CrossRef]
  171. Wadhwani, S.A.; Shedbalkar, U.U.; Singh, R.; Chopade, B.A. Biogenic selenium nanoparticles: Current status and future prospects. Appl. Microbiol. Biotechnol. 2016, 100, 2555–2566. [Google Scholar] [CrossRef]
  172. Guisbiers, G.; Lara, H.H.; Mendoza-Cruz, R.; Naranjo, G.; Vincent, B.A.; Peralta, X.G.; Nash, K.L. Inhibition of Candida albicans biofilm by pure selenium nanoparticles synthesized by pulsed laser ablation in liquids. Nanomed. Nanotechnol. Biol. Med. 2016, 13, 1095–1103. [Google Scholar] [CrossRef]
  173. Nandini, B.; Hariprasad, P.; Prakash, H.S.; Shetty, H.S.; Geetha, N. Trichogenic-selenium nanoparticles enhance disease suppressive ability of Trichoderma against downy mildew disease caused by Sclerospora graminicola in pearl millet. Sci. Rep. 2017, 7, 2612. [Google Scholar] [CrossRef]
  174. Pelgrift, R.Y.; Friedman, A.J. Nanotechnology as a therapeutic tool to combat microbial resistance. Adv. Drug Deliv. Rev. 2013, 65, 1803–1815. [Google Scholar] [CrossRef] [PubMed]
  175. Pan, X.; Redding, J.E.; Wiley, P.A.; Wen, L.; McConnell, J.S.; Zhang, B. Mutagenicity evaluation of metal oxide nanoparticles by the bacterial reverse mutation assay. Chemosphere 2010, 79, 113–116. [Google Scholar] [CrossRef] [PubMed]
  176. Sangaiya, P.; Jayaprakash, R. A Review on Iron Oxide Nanoparticles and Their Biomedical Applications. J. Supercond. Nov. Magn. 2018, 31, 3397–3413. [Google Scholar] [CrossRef]
  177. Sofi, M.A.; Sunitha, S.; Sofi, M.A.; Pasha, S.K.; Choi, D. An overview of antimicrobial and anticancer potential of silver nanoparticles. J. King Saud Univ. Sci. 2022, 34, 101791. [Google Scholar] [CrossRef]
  178. Parveen, S.; Wani, A.H.; Shah, M.A.; Devi, H.S.; Bhat, M.Y.; Koka, J.A. Preparation, characterization and antifungal activity of iron oxide nanoparticles. Microb. Pathog. 2018, 115, 287–292. [Google Scholar] [CrossRef] [PubMed]
  179. Nasrollahi, A.; Pourshamsian, K.; Mansourkiaee, P. Antifungal activity of silver nanoparticles on some of fungi. Int. J. Nano Dimens. 2011, 57, 233–239. [Google Scholar]
  180. López-Lorente, Á.I.; Cárdenas, S.; González-Sánchez, Z.I. Effect of synthesis, purification and growth determination methods on the antibacterial and antifungal activity of gold nanoparticles. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 103, 109805. [Google Scholar] [CrossRef]
  181. Baskar, G.; Chandhuru, J.; Fahad, K.S.; Praveen, A. Mycological synthesis, characterization and antifungal activity of zinc oxide nanoparticles. Asian J. Pharm. 2013, 3, 142–146. [Google Scholar]
  182. Mallmann, E.J.J.; Cunha, F.A.; Castro, B.N.M.F.; Maciel, A.M.; Menezes, E.A.; Fechine, P.B.A. Antifungal Activity of Silver Nanoparticles Obtained by Green Synthesis. Rev. Inst. Med. Trop. Sao Paulo 2015, 57, 165–167. [Google Scholar] [CrossRef]
  183. Ahmad, T.; Wani, I.A.; Lone, I.H.; Ganguly, A.; Manzoor, N.; Ahmad, A.; Ahmed, J.; Al-Shihri, A.S. Antifungal activity of gold nanoparticles prepared by solvothermal method. Mater. Res. Bull. 2013, 48, 12–20. [Google Scholar] [CrossRef]
  184. Pillai, A.M.; Sivasankarapillai, V.S.; Rahdar, A.; Joseph, J.; Sadeghfar, F.; Anuf, A.R.; Rajesh, K.; Kyzas, G.Z. Green synthesis and characterization of zinc oxide nanoparticles with antibacterial and antifungal activity. J. Mol. Struct. 2020, 1211, 128107. [Google Scholar] [CrossRef]
  185. Al-Zubaidi, S.; Al-Ayafi, A.; Abdelkader, H. Biosynthesis, Characterization and Antifungal Activity of Silver Nanoparticles by Aspergillus Niger Isolate. J. Nanotechnol. Res. 2019, 2, 022–035. [Google Scholar] [CrossRef]
  186. Kanhed, P.; Birla, S.; Gaikwad, S.; Gade, A.; Seabra, A.B.; Rubilar, O.; Duran, N.; Rai, M. In vitro antifungal efficacy of copper nanoparticles against selected crop pathogenic fungi. Mater. Lett. 2014, 115, 13–17. [Google Scholar] [CrossRef]
  187. Ponmurugan, P.; Manjukarunambika, K.; Elango, V.; Gnanamangai, B.M. Antifungal activity of biosynthesised copper nanoparticles evaluated against red root-rot disease in tea plants. J. Exp. Nanosci. 2016, 11, 1019–1031. [Google Scholar] [CrossRef]
  188. Kairyte, K.; Kadys, A.; Luksiene, Z. Antibacterial and antifungal activity of photoactivated ZnO nanoparticles in suspension. J. Photochem. Photobiol. B Biol. 2013, 128, 78–84. [Google Scholar] [CrossRef]
  189. He, L.; Liu, Y.; Mustapha, A.; Lin, M. Antifungal activity of zinc oxide nanoparticles against Botrytis cinerea and Penicillium expansum. Microbiol. Res. 2011, 166, 207–215. [Google Scholar] [CrossRef]
  190. Lipovsky, A.; Nitzan, Y.; Gedanken, A.; Lubart, R. Antifungal activity of ZnO nanoparticles—The role of ROS mediated cell injury. Nanotechnology 2011, 22, 105101. [Google Scholar] [CrossRef]
  191. Paladini, F.; Pollini, M. Antimicrobial Silver Nanoparticles for Wound Healing Application: Progress and Future Trends. Materials 2019, 12, 2540. [Google Scholar] [CrossRef]
  192. Arana, L.; Gallego, L.; Alkorta, I. Incorporation of Antibiotics into Solid Lipid Nanoparticles: A Promising Approach to Reduce Antibiotic Resistance Emergence. Nanomaterials 2021, 11, 1251. [Google Scholar] [CrossRef]
  193. Salarian, A.; Mollamahale, Y.B.; Hami, Z.; Soltani-Rezaee-Rad, M. Cephalexin nanoparticles: Synthesis, cytotoxicity and their synergistic antibacterial study in combination with silver nanoparticles. Mater. Chem. Phys. 2017, 198, 125–130. [Google Scholar] [CrossRef]
  194. Shruthi, T.; Meghana, M.; Medha, M.; Sanjana, S.; Navya, P.; Daima, H.K. Streptomycin functionalization on silver nanoparticles for improved antibacterial activity. Mater. Today Proc. 2019, 10, 8–15. [Google Scholar] [CrossRef]
  195. Mitra, P.; Dutta, D.; Das, S.; Basu, T.; Pramanik, A.; Patra, A. Antibacterial and Photocatalytic Properties of ZnO–9-Aminoacridine Hydrochloride Hydrate Drug Nanoconjugates. ACS Omega 2018, 3, 7962–7970. [Google Scholar] [CrossRef] [PubMed]
  196. Fadwa, A.O.; Alkoblan, D.K.; Mateen, A.; Albarag, A.M. Synergistic effects of zinc oxide nanoparticles and various antibiotics combination against Pseudomonas aeruginosa clinically isolated bacterial strains. Saudi J. Biol. Sci. 2020, 28, 928–935. [Google Scholar] [CrossRef] [PubMed]
  197. Pissuwan, D.; Cortie, C.H.; Valenzuela, S.M.; Cortie, M.B. Functionalised gold nanoparticles for controlling pathogenic bacteria. Trends Biotechnol. 2010, 28, 207–213. [Google Scholar] [CrossRef]
  198. Jelinkova, P.; Mazumdar, A.; Sur, V.P.; Kociova, S.; Dolezelikova, K.; Jimenez, A.M.J.; Koudelkova, Z.; Mishra, P.K.; Smerkova, K.; Heger, Z.; et al. Nanoparticle-drug conjugates treating bacterial infections. J. Control. Release 2019, 307, 166–185. [Google Scholar] [CrossRef]
  199. Muhamad, N.; Plengsuriyakarn, T.; Na-Bangchang, K. Application of active targeting nanoparticle delivery system for chemotherapeutic drugs and traditional/herbal medicines in cancer therapy: A systematic review. Int. J. Nanomed. 2018, 13, 3921–3935. [Google Scholar] [CrossRef]
  200. Zhang, Y.; Dasari, T.P.S.; Deng, H.; Yu, H. Antimicrobial Activity of Gold Nanoparticles and Ionic Gold. J. Environ. Sci. Heath Part C Environ. Carcinog. Ecotoxicol. Rev. 2015, 33, 286–327. [Google Scholar] [CrossRef]
  201. Saha, B.; Bhattacharya, J.; Mukherjee, A.; Ghosh, A.; Santra, C.; Dasgupta, A.K.; Karmakar, P. In Vitro Structural and Functional Evaluation of Gold Nanoparticles Conjugated Antibiotics. Nanoscale Res. Lett. 2007, 2, 614–622. [Google Scholar] [CrossRef]
  202. Gu, H.; Ho, P.L.; Tong, E.; Wang, L.; Xu, B. Presenting Vancomycin on Nanoparticles to Enhance Antimicrobial Activities. Nano Lett. 2003, 3, 1261–1263. [Google Scholar] [CrossRef]
  203. Aisida, S.O.; Akpa, P.A.; Ahmad, I.; Zhao, T.-K.; Maaza, M.; Ezema, F.I. Bio-inspired encapsulation and functionalization of iron oxide nanoparticles for biomedical applications. Eur. Polym. J. 2019, 122, 109371. [Google Scholar] [CrossRef]
  204. Arias, L.S.; Pessan, J.P.; Vieira, A.P.M.; de Lima, T.M.T.; Delbem, A.C.B.; Monteiro, D.R. Iron Oxide Nanoparticles for Biomedical Applications: A Perspective on Synthesis, Drugs, Antimicrobial Activity, and Toxicity. Antibiotics 2018, 7, 46. [Google Scholar] [CrossRef] [PubMed]
  205. Caamano, M.A.; Morales, M.C. Iron Oxide Nanoparticle Improve the Antibacterial Activity of Erythromycin. J. Bacteriol. Parasitol. 2016, 7, 2. [Google Scholar] [CrossRef]
  206. Ribeiro, A.I.; Dias, A.M.; Zille, A. Synergistic Effects Between Metal Nanoparticles and Commercial Antimicrobial Agents: A Review. ACS Appl. Nano Mater. 2022, 5, 3030–3064. [Google Scholar] [CrossRef] [PubMed]
  207. Rayegan, A.; Allafchian, A.; Sarsari, I.A.; Kameli, P. Synthesis and characterization of basil seed mucilage coated Fe3O4 magnetic nanoparticles as a drug carrier for the controlled delivery of cephalexin. Int. J. Biol. Macromol. 2018, 113, 317–328. [Google Scholar] [CrossRef] [PubMed]
  208. Weiss, C.; Carriere, M.; Fusco, L.; Capua, I.; Regla-Nava, J.A.; Pasquali, M.; Scott, J.A.; Vitale, F.; Unal, M.A.; Mattevi, C.; et al. Toward Nanotechnology-Enabled Approaches against the COVID-19 Pandemic. ACS Nano 2020, 14, 6383–6406. [Google Scholar] [CrossRef] [PubMed]
  209. Schrand, A.M.; Rahman, M.F.; Hussain, S.M.; Schlager, J.J.; Smith, D.A.; Syed, A.F. Metal-based nanoparticles and their toxicity assessment. WIREs Nanomed. Nanobiotechnol. 2010, 2, 544–568. [Google Scholar] [CrossRef] [PubMed]
  210. Foulkes, R.; Man, E.; Thind, J.; Yeung, S.; Joy, A.; Hoskins, C. The regulation of nanomaterials and nanomedicines for clinical application: Current and future perspectives. Biomater. Sci. 2020, 8, 4653–4664. [Google Scholar] [CrossRef]
  211. Najahi-Missaoui, W.; Arnold, R.D.; Cummings, B.S. Safe Nanoparticles: Are We There Yet? Int. J. Mol. Sci. 2020, 22, 385. [Google Scholar] [CrossRef]
  212. Matusiewicz, H. Potential release of in vivo trace metals from metallic medical implants in the human body: From ions to nanoparticles–A systematic analytical review. Acta Biomater. 2014, 10, 2379–2403. [Google Scholar] [CrossRef]
  213. Asmatulu, E.; Andalib, M.N.; Subeshan, B.; Abedin, F. Impact of nanomaterials on human health: A review. Environ. Chem. Lett. 2022, 20, 2509–2529. [Google Scholar] [CrossRef]
  214. Intisar, A.; Ramzan, A.; Sawaira, T.; Kareem, A.T.; Hussain, N.; Din, M.I.; Bilal, M.; Iqbal, H.M.J.C. Occurrence, toxic effects, and mitigation of pesticides as emerging environmental pollutants using robust nanomaterials—A review. Chemosphere 2022, 293, 133538. [Google Scholar] [CrossRef] [PubMed]
  215. Wu, J.; Liu, W.; Xue, C.; Zhou, S.; Lan, F.; Bi, L.; Xu, H.; Yang, X.; Zeng, F.-D. Toxicity and penetration of TiO2 nanoparticles in hairless mice and porcine skin after subchronic dermal exposure. Toxicol. Lett. 2009, 191, 1–8. [Google Scholar] [CrossRef] [PubMed]
  216. Kumar, L.; Ragunathan, V.; Chugh, M.; Bharadvaja, N. Nanomaterials for remediation of contaminants: A review. Environ. Chem. Lett. 2021, 19, 3139–3163. [Google Scholar] [CrossRef]
  217. Sharifi, S.; Behzadi, S.; Laurent, S.; Forrest, M.L.; Stroeve, P.; Mahmoudi, M. Toxicity of nanomaterials. Chem. Soc. Rev. 2012, 41, 2323–2343. [Google Scholar] [CrossRef]
Figure 1. Schematic Illustration of Antibiotic Resistance Mechanisms in Bacteria.
Figure 1. Schematic Illustration of Antibiotic Resistance Mechanisms in Bacteria.
Ijms 24 14897 g001
Figure 2. Illustration of the Mechanistic Approach of NPs Against Bacteria.
Figure 2. Illustration of the Mechanistic Approach of NPs Against Bacteria.
Ijms 24 14897 g002
Figure 3. Illustration of NPs Antifungal Mechanism.
Figure 3. Illustration of NPs Antifungal Mechanism.
Ijms 24 14897 g003
Table 1. Describes the methods of synthesis and antibacterial effects of different types of metallic NPs.
Table 1. Describes the methods of synthesis and antibacterial effects of different types of metallic NPs.
S. NoNPsSynthesis MethodBacterial StrainsSummary of ResultsReferences
1SilverJusticia adhatoda L. leavesPseudomonas aeruginosaInhibit bacterial growth[102]
2GoldCashew nutshell extract Anacardim occidentalePseudomons fluorescens
Aeromonas bestiarum
For Aermonas bestiarum: MIC values = 294 ± 12.8 μg/mL MBC values = 363 ± 16.2 μg/mL
For Pseudomonas fluorescens: MIC values = 386 ± 12.7 μg/mL MBC values = 294 ± 9.42 μg/mL
[103]
3SilverLeaves extract of Urtica dioica (Linn.)Bacillus cereus, Bacillus subtilisFor Bacillus cereus:
MIC values = 284 ± 12.6 μg/mL MBC values = 361 ± 15.7 μg/mL
For Bacillus subtilis:
MIC values = 376 ± 12.5 μg/mL MBC values = 282 ± 9.43 μg/mL
[104]
4SilverBanana peel extractBacillus subtilis, Staphylococcus aureusMICs were 1.70, 5.1, 6.8, and 3.4 μg/mL
MBCs value of AuNPs is 0.2, 10.2, 5.1 μg/mL
[105]
5GoldFrom Trianthema decandraStaphylococcs aureus, Streptococcus faecalisFor Staphylococcs aureus:
MIC values = 113 ± 9.1 μg/mL MBC values = 111 ± 13.5 μg/mL For Streptococcus faecalis:
MIC values = 245 ± 12.2 μg/mL MBC values = 176 ± 12.9 μg/mL
[106]
6GoldCitrullus lanatus rindEscherichia coli, Bacillus cereusFor Bacillus cereus:
MIC value = 50 μg/mL
MBC value = 100 μg/mL
For Escherichia coli:
MIC value = 50 μg/mL
[107]
7GoldAgaricus bisporus extractsStaphylococcus aureus, Escherichia coliMIC values = 50.99 μg/mL for E. coli
MIC values = 198.2 μg/mL for S. aureus
[108]
8GoldUsing aqueous Plumeria alba flower extractEscherichia coliMIC value = 400 μg/mL[109]
9GoldSalix albaKlebsiella pneumoniae, Bacillus subtilis, Staphylococcus aureus-----[110]
10GoldBrassica oleraceaStaphylococcus aureus, Klebsiella pneumoniaeMIC values = 25 μg/mL for S. aureus
MIC values = 50 μg/mL for K. pneumoniae
[111]
11Zinc oxideHibiscus subdarifa leaf extractEscherichia coli, Staphylococucs aureusFor Escherichia coli with PZN60: MIC value = 24 ± 1 μg/mL
MBC value = 50 ± 1 μg/mL
For Staphylococcs aureus with PZN60:
MIC value = 50 ± 1 μg/mL
MBC value = 50 ± 1 μg/mL
[112]
12Zinc oxideParthenium hysterophorus extractStaphylococcus aureus, Bacillus subtilisMIC value = 11 ± 0.28 μg/mL for Staphylococcus aureus
MIC value = 10 ± 0.16 μg/mL for Bacillus subtilis
[113]
13Zinc oxideCamellia sinensis extractsKlebsiella pneumoniae, Pseudomonas aeruginosa, Escherichia coliMIC value = 10.3 ± 0.57 μg/mL for K. pneumoniae
MIC value = 3.3 ± 0.57 μg/mL for P. aeruginosa
[114]
14IronMoringa oleifera extractsEscherichia coliMIC value = 59 ± 1.22 μg/mL for Escherichia coli[115]
15Iron oxideCynometra ramifloraEscherichia coli, S. epidermidis--------[116]
16Iron oxideLagenaria siceraria leaves extractEscherichia coli, Staphylococcus aureus--------[117]
17GoldReduction of tetra chloroauric acid with sodium nitrateE. coli k12MIC value = 7.4 μg/mL[118]
18GoldM. piperitaE. coli, S. aureusShowed activity against E. coli and no activity against S. aureus[119]
19GoldPurchasedSalmonella typhi, Salmonella enteritisMIC values = 2.5–5 μg/mL[120]
20Iron oxideChemically by laser ablation in liquidSerratia marcescens, Escherichia coli, Pseudomonas aeruginosa, Staphylococcus aureusThe disc diffusion method found iron oxide NPs inhibitory zones against Gram-negative and Gram-positive bacteria.[121]
21Iron oxideSynthesized from ferric chloride and ferrous chloride using the co-precipitation methodB. cereus, Klebsiella pneumoniaeAt 40 μg/well concentration of Fe3O4-NPs, the inhibitory zone was 15 mm against K. pneumoniae and 13 mm against B. cereus
At 80 μg/well concentration of Fe3O4-NPs, the inhibitory zone was 26 mm against K. pneumoniae and 22 mm against B. cereus
[122]
22IronAloe vera leavesProteus mirabilis, Escherichia coli, Klebsiella pneumoniae, Pseudomonas aeruginosa, Shigella flexneri, Serratia marcescenes, Salmonella typhi, Enterococcus faecalis, Staphylococcus aureusZone of inhibition (mm) at 40 µg/well:
Escherichia coli = 15 ± 0.11 Proteus mirabilis = 16 ± 0.21 Klebsiella pneumoniae = 17 ± 0.54 Pseudomonas aeruginosa = 16 ± 0.29
Shigella flexneri = 14 ± 0.61 Serratia marcescenes = 15 ± 0.58 Salmonella typhi = 16 ± 0.66 Enterococcus faecalis = 15 ± 0.13 Staphylococcus aureus = 15 ± 0.79
[123]
23Zinc oxideSynthesized from zinc acetylacetonate hydrate and oleylamineE. coliShowed excellent antibacterial activity (10 CFU/mL)[124]
24Zinc oxideFrom leaves and fruits of C. proceraE. coli, Vibrio choleraeMIC value = 1.6 × 105 − 1.2 × 106 per mL[125]
25SilverSynthesized from C. procera leaves and fruitsVibrio cholerae, E. coliMIC value = 5 × 106 − 1.2 × 107 per mL[125]
Table 2. Describe methods of synthesis and antifungal effects of different types of metallic NPs.
Table 2. Describe methods of synthesis and antifungal effects of different types of metallic NPs.
S. NoNPsSynthesis MethodFungal StrainsSummary of Results and
Antifungal Potency
References
1.SilverUsing PVP as a reducing agentSaccharomyces cerevisiae, Candida albicans. MIC 50 = 0.5 mg/mL and 4 mg/mL against Saccharomyces cerevisiae and Candida albicans, respectively.[179]
2.GoldStainless steel (reducing agent) mediated reductionCandida albicans (ATCC 10231)Cell viability count through the Neubauer chamber gave antifungal activity at 20 mg/mL.[180]
3.GoldFor reducing agent citrate are used. Candida albicans (ATCC 10231)Cell viability count through the Neubauer chamber gave antifungal activity at 40 mg/mL.[180]
4.Zinc oxideMycological synthesis Aspergillus niger, Aspergillus fumigatus, Aspergillus aculeatusA large zone of clearance was obtained with the largest against Aspergillus fumigatus.[181]
5.Silver Reduced with ribose and stabilized with sodium dodecylCandida albicans and Candida tropicalisA high antifungal activity like that of amphotericin B disc.[182]
6. Gold Solvothermal methodCandida isolatesThe zone of clearance observed to be 4.2 mm/mg against Candida albicans & 1.1 mm/mg against Candida glabrata.[183]
7.Zinc oxideUse of leaf extracts of medicinal plants such as Beta vulgaris, Cinnamomum verum, Cinnamomum tamala, and Brassica oleracea var. ItalicaCandida albicans and Aspergillus nigerBeta vulgaris based NPs revealed potency against A. niger, Cinnamomum tamala based NPs showed activity against C. Albicans. Both fungal strains were sensitive to ZnONPsynthesized from Brassica oleracea var. italic.[184]
8.Silver Aspergillus niger fungal isolatesAspergillus flavus, Fusarium oxysporum and Penicillium digitatum.The lower MIC values i.e., 6.75 ± 0.24, 7.45 ± 0.18, and 9.62 ± 0.14 obtained for Penicillium digitatum, Aspergillus flavus, and Fusarium oxysporum, respectively.[185]
9. Copper The use of Cetyl Trimethyl Ammonium Bromide and isopropyl alcohol as reducing agent in the chemical reduction of Cu2+Curvularia lunata Phoma destructiva Alternaria alternata and Fusarium oxysporum.The inhibitory zone was
22 ± 1 mm against Phoma destructiva, 21 ± 0.5 mm against Curvularia lunata, 18 ± 1.1 mm against Alternaria alternata, and against Fusarium oxysporum was 24 ± 0.5 mm.
[186]
10.Copper Extracellular synthesis by Streptomyces griseusRed-root rot disease-causing fungus52.7% of the disease was reduced with the application of 2.5 ppm CuNPs in the selected bushes.[187]
11.CopperUsing CTAB as a reducing agent in chemical reduction methodFusarium sp.93.98% of fungal growth was inhibited with the application of 450 ppm of CuNPs after 9 days of incubation[167]
12.Photo-activated Zinc oxideObtained from Alfa Aesar (NanoShield, Germany)Botrytis cinereaZnONPs were revealed to cause morphological changes to fungus after treatment with photoinactivation (58%) and NPs at a concentration of 5 × 10−3 M. [188]
13.Zinc oxideAlfa Aesar (Ward Hill, MA, USA)Penicillium
expansum and
Botrytis cinerea
Results showed that 3 mmoL/L of ZnO NPs can pointedly stop the growth of P. expansum and B. cinerea.[189]
14.Zinc oxide Ultrasonic methodCandida albicansThe minimum concentration of ZnO required to effectively stop the growth of Candida albicans was found to be 0.1 mg/mL, resulting in more than 95% inhibition[190]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wahab, S.; Salman, A.; Khan, Z.; Khan, S.; Krishnaraj, C.; Yun, S.-I. Metallic Nanoparticles: A Promising Arsenal against Antimicrobial Resistance—Unraveling Mechanisms and Enhancing Medication Efficacy. Int. J. Mol. Sci. 2023, 24, 14897. https://doi.org/10.3390/ijms241914897

AMA Style

Wahab S, Salman A, Khan Z, Khan S, Krishnaraj C, Yun S-I. Metallic Nanoparticles: A Promising Arsenal against Antimicrobial Resistance—Unraveling Mechanisms and Enhancing Medication Efficacy. International Journal of Molecular Sciences. 2023; 24(19):14897. https://doi.org/10.3390/ijms241914897

Chicago/Turabian Style

Wahab, Shahid, Alishba Salman, Zaryab Khan, Sadia Khan, Chandran Krishnaraj, and Soon-Il Yun. 2023. "Metallic Nanoparticles: A Promising Arsenal against Antimicrobial Resistance—Unraveling Mechanisms and Enhancing Medication Efficacy" International Journal of Molecular Sciences 24, no. 19: 14897. https://doi.org/10.3390/ijms241914897

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