Next Article in Journal
2-Methoxy-4-Vinylphenol as a Biobased Monomer Precursor for Thermoplastics and Thermoset Polymers
Next Article in Special Issue
Development and Characterization of Bio-Based Composite Films for Food Packing Applications Using Boiled Rice Water and Pistacia vera Shells
Previous Article in Journal
Rheological Properties of Small-Molecular Liquids at High Shear Strain Rates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Metal–Polymer Nanocomposites: A Promising Approach to Antibacterial Materials

1
Faculty of Pharmacy, Al-Ahliyya Amman University, Amman 19328, Jordan
2
Pharmacological and Diagnostic Research Center, Faculty of Pharmacy and Allied Medical Sciences, Al-Ahliyya Amman University, Amman 19328, Jordan
3
Faculty of Allied Medical Sciences, Al-Ahliyya Amman University, Amman 19328, Jordan
*
Author to whom correspondence should be addressed.
Polymers 2023, 15(9), 2167; https://doi.org/10.3390/polym15092167
Submission received: 28 March 2023 / Revised: 24 April 2023 / Accepted: 28 April 2023 / Published: 2 May 2023
(This article belongs to the Special Issue Polymer Composites with Antibacterial Properties)

Abstract

:
There has been a new approach in the development of antibacterials in order to enhance the antibacterial potential. The nanoparticles are tagged on to the surface of other metals or metal oxides and polymers to achieve nanocomposites. These have shown significant antibacterial properties when compared to nanoparticles. In this article we explore the antibacterial potentials of metal-based and metal–polymer-based nanocomposites, various techniques which are involved in the synthesis of the metal–polymer, nanocomposites, mechanisms of action, and their advantages, disadvantages, and applications.

1. Introduction

Antibiotics are designed to prevent bacterial infection, depending on the four mechanisms or the activities that the bacteria carry out. However, there are some limitations, such as difficulty in penetrating into the bacterial cell and excreting of these antibiotics from the system. To overcome these limitations, there is a new approach in this regard; new antibacterials with nanoparticles have been developed based on the existing mechanism of the antibacterials and these antibacterials loaded with nanoparticles have shown significant antibacterial potencies [1]. In recent years, there has been a new approach developed in this regard to enhance the antibacterial potency by tagging the nanoparticles on the surface of other metal oxides or polymers to acquire nanocomposites. The literature survey and the research on these nanocomposites show that nanocomposites exhibit more efficient antibacterial properties when compared to nanoparticles alone. Composite materials combine the favorable properties of a matrix and filler, and can lead to new functionalities, especially due to the size of the filers which are in nanoscale. These composite materials are now popularly known as nanocomposites. One of the approaches towards synthesizing the nanocomposites is to tag the nanoparticles with either metal oxide or polymers. Polymers are one of the choices; polymers are a popular choice as the matrix component due to their low cost, versatility, and ability to be easily processed into various shapes, including thin films. While much research has been carried out on structural polymer-based nanocomposites, there have been fewer investigations on polymer nanocomposites for functional applications. Of particular interest are nanocomposites that contain metal nanoparticles dispersed in a dielectric matrix, which offer novel properties and have various new applications such as electronic conductivity, optical features, magnetic properties, and enhanced catalytic activity.
It may be noted that the investigation of antibacterial properties of metal nanoparticles started with silver and copper in 2003, followed by gold nanoparticles in the same year. The number of Scopus publications related to metal nanoparticles research has increased significantly since then, with silver being the most studied metal and having the highest percentage of Scopus publications at 56.8%. Zinc, copper, and gold were also investigated, but to a lesser extent, with 16.1%, 10%, and 8.6% of the Scopus publications, respectively (Figure 1). Iron and other metals were also studied but had lower percentages of Scopus publications. The data suggest that the scientific community has shown significant interest in investigating the antibacterial properties of metal nanoparticles. Silver has emerged as the most extensively studied metal nanoparticle for antibacterial purposes, possibly due to its broad-spectrum antimicrobial activity, low toxicity, and relatively lower cost than gold. The lower number of publications for other metals may suggest that they are less promising compared to silver, but it is essential to note that this may not reflect the quality or significance of the research.
This review highlights the potential of metal nanoparticles as promising antibacterial agents. Metal-based nanocomposites’ bacterial action, their mechanism of action, various mechanisms that are currently employed in the synthesis of the nanocomposites, potential of metal–polymer-based nanocomposites, their challenges and their application in various fields are investigated, and this sets the groundwork for further research in this field. We conclude the article with a positive approach in developing therapeutically enhanced antibacterials.

2. Antibacterial Mechanism of Action of Metal–Polymer Nanocomposites

Antibacterial mechanisms of metal nanoparticles and metal oxide nanoparticles nanocomposites are discussed in this section. Among metals, silver, copper, zinc, and gold nanocomposites are discussed here in this section. We also briefly explain the antiviral property of some of the NPs.
Gold and silver nanoparticles have been shown to possess significant amounts of antibacterial activity when compared to other metal nanoparticles. The antibacterial activity of the nanocomposite depends on the interaction between the chemical released by the nanocomposite and between the bacteria. Figure 2 shows the different mechanisms of action of metallic nanoparticles as antibacterials.

2.1. Antibacterial Mechanism of Action of Silver Nanocomposites

The silver nanoparticle exhibits antibacterial activity through various mechanisms such as interfering with the metabolic activity, enzymes, denaturation of protein, condensation of DNA, decrease in ATP production, production of ROS, and causing oxidative stress. Bao et al. [2] showed that silver-based graphene oxide nanocomposite sheets showed high rates of antibacterial activity. They attributed this to the release of silver nanoparticles from the graphene nanocomposite sheets that caused the destruction of the protein and the DNA of the bacteria, resulting in death of the bacteria. In a similar study involving silver graphene oxide nanocomposite sheets, it showed bactericidal properties. In another study, the antibacterial activity of graphene-oxide-based silver nanocomposites and graphene oxide decorated with L-cysteine nanocomposites exhibited activity through different mechanisms against Gram-positive and Gram-negative bacteria. It had bacteriostatic effect against Gram-positive bacteria, and Gram-negative bacteria died of cell wall destruction. This difference in the mechanism was attributed to the difference in the cell wall component and the structure [3]. The cysteine and silver nanocomposites interfere with the metabolic processes in Gram-positive bacteria by interacting with the thiol group of the enzyme, and fragmentation of DNA was caused by interacting with phosphorus in the nucleic acid. In Gram-negative, it caused damage to the cell wall by causing oxidative and mechanical stress to the cell wall [4]. Considering copper nanoparticles, it produces destruction in the helical structured links of DNA within and in between its strands.

2.2. Antibacterial Mechanism of Action of Copper Nanocomposites

Copper nanoparticles are shown to possess the antibacterial properties due to the release of monovalent and divalent copper ions from the nanoparticles that damage the cell wall of the bacteria. Usually, metals such as gold and silver are tagged to enhance the antibacterial properties, but as gold and silver are costlier than other metals, copper would be an excellent choice in the synthesis of nanocomposites. Carboxymethyl cellulose coupled with copper oxide nanocomposites exhibited an appreciable amount of antibacterial activity against both Gram-positive and Gram-negative bacteria. The mechanism by which it caused the antibacterial activity caused the leakage of the cellular constituents that killed the bacteria [5]. The antibacterial mechanism of zeolite nanocomposite along with zinc oxide and copper oxide was due to the release of the ions from nanoparticles (zinc and copper) that damaged the surface of the bacteria; zeolite further enhanced this property that killed the bacteria. Zeolite also acted as the stabilizing agent, preventing the agglomeration of the copper and zinc ions [6]. Copper oxide nanocomposites synthesized from copper oxide and zinc oxide nanoparticles with graphene oxide produced higher zones of inhibition against Staphylococcus aureus. Copper oxide nanocomposites (produced by cuprous oxide on the graphene nanocomposites sheets) are also known to produce bactericidal effects by producing reactive oxygen species such as hydrogen peroxide and super oxide anions. Hydrogen peroxide has a 70% bactericidal effect on Gram-negative bacteria (Escherichia coli) and a 65% bactericidal effect on Gram-positive bacteria (Staphylococcus aureus), respectively.

2.3. Antibacterial Mechanism of Action of Zinc Nanocomposites

Ghosh et al. [7], in their study, reported that the mechanism by which the synthesized silver zinc oxide nanocomposite exhibited antibacterial activity may be due to the binding of nanocomposites to the cell wall of the bacteria; they were unable to predict the exact mechanism. To understand the mechanism of the zinc oxide nanoparticles coupled with graphene oxide nanosheets nanocomposite, Prema et al. [8] carried out analysis for DNA fragmentation, lactate dehydrogenase leakage, and ROS analysis. Prema and her colleagues reported that the chromosomal DNA was the same, but there was increase in lactate dehydrogenase leakage and reactive oxygen species in all the microbes upon exposure to zinc oxide nanocomposites. Thambidurai et al. [9] attributed the antibacterial activity of nickel and zinc oxide nanocomposites to their large surface area and increase in the production of the ROS that killed the bacteria.

2.4. Antibacterial Mechanism of Action of Gold Nanocomposites

Huang et al. [10] synthesized gold nanoparticles using a photothermal agent under the near infrared (NIR), and it had a bactericidal effect on the pathogenic bacteria. They synthesized polygonal shaped nanoparticles and appended them with vancomycin. These nanocomposites of gold vancomycin binded to the peptide part D-Ala-D-Ala in the pathogenic bacteria cell and caused the transfer of heat from the light source into the cell, causing its bactericidal effect; this killed more than 99% of the pathogenic bacteria, both Gram-positive and Gram-negative, and antibiotic-resistant bacteria [11].
In another study, Perni and his research team synthesized light-induced polymer nanocomposites made up of methylene blue and gold nanoparticles [11]. These nanocomposites’ gold nanoparticles exhibited bactericidal activity against methicillin-resistant Gram-negative bacteria Escherichia coli and Gram-positive bacteria Staphylococcus aureus. Attributed to the release of singlet oxygen and ROS from methylene blue due to interaction of light on the surface of the polymer, the presence of the gold nanoparticles enhanced the release of reactive oxygen species, further enhancing the bactericidal property. This may be of benefit in biomedical applications. Nirmala et al. [12] reported the bactericidal activity of gold nanoparticles coated with hydroxyapatite immaculate; they reported that the hydroxyapatite in its unaltered form had no harmful effect on the bacterial, but together with metal nanoparticles it had the bactericidal properties. The bactericidal effect was due to the production of electrons or ROS that bind to the membrane or surface of the plasma membrane and penetrate inside, causing the bactericidal effect. These nanocomposites have large surface area volume and are hyperbranched, which offers electrons to acts as a best antibacterial agent, but these nanoparticles are known to cause toxicity to osteoblast cells in the tissue. Hence, their use may be limited due to this factor.
In another study, the nanocomposites synthesized from gold nanoparticles with poly-thiophene exhibited a significant bactericidal effect. The bactericidal properties were attributed to their ability to penetrate the cell membrane and kill the bacteria. This may be an excellent choice in preventing biofilm formation and inhibiting the gastrointestinal pathogens. Gold poly-thiophene nanocomposites have significant potency in eliminating the pathogen and are not toxic to human cells. Hence, this nanocomposite may be an excellent choice for therapeutic purposes for humans.
Polymer-based gold nanocomposites have shown bactericidal activity. Regiel-Futyra et al. (2015) synthesized polymer-based gold nanocomposites; they used chitosan flakes and gold chloride solution, and chitosan-based nanocomposites exhibited bactericidal activity against Staphylococcus aureus (ATTC 25923) and Pseudomonas aeruginosa (ATTC 27853) [13]. The polymer puffed up to reach the cell membrane of the bacteria, causing damage to the cell. A similar type of work was reported by Mendoza et al. [14]; they synthesized chitosan-based gold nanocomposites with two different gold solutions. These nanocomposites exhibited significant bactericidal effects without causing any damage to the human cell.
When antimicrobial peptide daptomycin was combined with gold nanoclusters to form hybrid covalent conjugates of nanocomposites, these hybrid nanocomposites were able to produce pores or openings in the membrane through which they entered and damaged the DNA, killing the bacterial cell. The mechanism by which nanocomposites prepared by conjugating iron oxide with gold nanoparticles exhibited the antibacterial activity was attributed to the attraction between positively charged ions (iron and gold) with the negatively charged ions on the bacterial surface, causing them to stick to the surface and resulting in death of the bacteria.
Metal and metal-based NPs exhibit antimicrobial properties; they exhibit antibacterial antiviral properties. The antiviral or the virucidal properties of some of the nanoparticles and nanocomposites are discussed briefly in the following paragraph.
In the Guangdong province of China, in 2003, COVID-19 caused severe respiratory syndrome, affecting 8000 people in 37 countries with a low mortality of 10%. In 2012, severe respiratory syndrome MERS (Middle East respiratory syndrome) [15,16] affected individuals on the Arabian Peninsula with a high mortality rate (40%). In December 2019, in Hubei province, China, COVID-19 first appeared and spread across the world rapidly. Later, WHO declared it a pandemic [17,18]. Among nanoparticles, metal and metal oxide nanoparticles based on gold, silver, copper, and zinc are known to exhibit extensive antimicrobial properties, including viruses, bacteria, and fungi [19]. Silver nanoparticles are known to cause destruction to viruses [20,21], metal oxides, copper, and iron [22]. When silver and gold nanoparticles are associated with polymers and textiles, they confer viricidal properties to them [23]. Graphene oxide (GO) showed potent antiviral activity with nonionic polymer polyvinylpyrrolidone (PVP), but with cationic polymer poly-diallyl-dimethylammonium (PDDA) it did not show antiviral activity. GO also caused destruction of virus before the entry, further emphasizing the antiviral activity [24]. These are some of the methods that directly act on the virus inside the host. Another way to control the viral infection is to disinfect the surfaces that act as sources of contamination due to discharge from affected persons. The development of antiviral surfaces is a new concept that is currently picking up owing to the rise in viral infection. Lishchynskyi et al. (2022) extensively reviewed articles, highlighting the classification of modern antiviral surfaces, and the impacts of nonspecific and specific interaction on virus deposition. They also outlined the synthesis and characterization of nanostructured antiviral surfaces and lastly on the effects of different antiviral surfaces on different viruses [25].

3. Types of Metal–Polymer Nanocomposites with Antibacterial Properties Based on Metals (Silver, Copper, Zinc, and Gold)

3.1. Silver-Based Nanocomposites

Silver nanoparticles have been widely used as an antibacterial agent in metal–polymer nanocomposites [26]. Silver has a broad-spectrum antibacterial effect and can disrupt the bacterial cell membrane [27,28], leading to cell death [29]. Silver-based nanocomposites have been used in medical applications [29] such as wound dressings [30], orthopedic implants [31], and catheters [32]. The use of metal nanoparticles, specifically silver nanoparticles (AgNPs), as antimicrobial agents in medicine is a crucial application. When AgNPs are dissolved in an aqueous solution, Ag+ ions interact with microorganisms and have various antibacterial mechanisms. One of these mechanisms involves Ag+ ions binding to sulfur and phosphorus groups in the proteins of the cell wall and plasma membrane of bacteria, causing protein dysfunction and threatening the organism’s life [33,34]. Ag+ ions can also create holes in the microorganism’s membrane by binding to negatively charged parts, leading to the flow of cytoplasmic contents out of the cell, dissipating the proton gradient across the membrane, and ultimately causing cell death [35,36]. The biohazards of silver nanoparticles are well documented [37]. Ag+ ions inside the cell can disturb the function of the bacteria’s electron transport chain, and they can interact with bacterial DNA and RNA, inhibiting cell division [38,39].
Several factors affect the activity of AgNPs, including size, shape, and coating, as well as the medium’s parameters such as the presence of light, oxidative species, other potential ligands for silver, and ionic strength. These factors can impact various phenomena that contribute to the increase or decrease of the antibacterial activity of AgNPs through complex pathways. Table 1 shows the biomedical applications of Ag–polymer nanocomposites and their sources and properties.

3.2. Copper-Based Nanocomposites

Copper has also been used as an antibacterial agent in metal–polymer nanocomposites [61]. Copper ions can disrupt the bacterial cell wall and inhibit bacterial growth by inducing cell membrane [62] or intracellular protein and DNA disruption [63] and generation of reactive oxide species [64].
The incorporation of antimicrobial metals into a polymer can result in the creation of composite materials, but the effectiveness of this method is dependent on both the intended use of the composite and the type of polymer used. The metal can be added to the surface of the polymer or embedded within the matrix [65]. The use of polymers in creating nanocomposites with antimicrobial properties is not limited to providing support for nanoparticles as the polymer itself can enhance the antibacterial performance of the composite. This is due to the synergy between the polymer and the nanoparticles, the long-term ion release ability of the polymer, and the increased surface area resulting from the fine dispersion of nanoparticles within the polymer [65]. There are two general approaches for preparing polymer/metal nanocomposites, based on where the nanoparticles are produced. The first is in situ, which involves using the polymer matrix as the reaction medium. The second is ex situ, in which the particles are synthesized separately and then added to the polymer with a dispersion medium [65]. Table 2 illustrates the several synthetic and natural sources of copper–polymer nanocomposites.

3.3. Zinc-Based Nanocomposites

Zinc has also shown antibacterial properties and has been used in metal–polymer nanocomposites. Zinc ions can interfere with bacterial metabolism and inhibit bacterial growth. Zinc-based nanocomposites have been used in medical applications such as dental materials and wound dressings (Table 3).
The misuse of antimicrobial agents has led to the emergence of multidrug-resistant bacterial strains at an alarming rate, resulting in an increase in infectious diseases and associated mortality [103]. This resistance can arise through mutations, horizontal gene transfer, or adaptive changes [104]. The chronic and intense use of antifungal agents also leads to decreased sensitivity and the production of resistant strains, along with severe toxic effects. Biofilm formation by microbes also shields them from the action of antibiotics [105], making it crucial to investigate novel antimicrobial agents that can bypass multidrug resistance and biofilm-mediated protection [106]. Inorganic metal oxides, such as ZnO, have shown promise as antibacterial agents by damaging the cell membrane and inducing oxidative stress [107]. ZnONP, in particular, exhibits higher photocatalytic activity, increased biocompatibility, and selectivity, along with heat resistance and durability. Incorporating ZnONP into antimicrobial therapeutics could reduce health hazards associated with traditional antimicrobial agents [108]. The mechanism by which NPs exhibit photocatalytic activity is that when the NP absorbs a particular wavelength of light, it causes the excitation of the particles and causes the movement of electrons from the valence band to the conduction band, resulting in the pore formation and generation of free hydroxyl ions and radical anions, which may cause the death of bacteria [109].
Table 3. Zinc–polymer nanocomposites and their biomedical applications.
Table 3. Zinc–polymer nanocomposites and their biomedical applications.
No.Polymer NameSourcePropertiesUsesRef.
1Chitosan hydrogelSynthetic30 nmAntimicrobial[110]
2CelluloseSynthetic65 nmPhotocatalytic and antibacterial[111]
3Agar biopolymerSynthetic20 nmAntibacterial and anticancer[112]
4Gelatin/tragacanthSynthetic10.6 nmAntimicrobial biomaterials for food packaging[113]
5ChitosanNatural25 to 70 nmAntimicrobial[114]
6ChitosanSynthetic20–150 nmAntibacterial and photocatalytic[115]
7Low density polyethylene (LDPE)Synthetic~17 nmAntimicrobial[116]
8ChitosanNatural24 nmAntibacterial and photocatalytic[117]
9Alginate beadsNatural20–100 nmAntimicrobial agent for water disinfection[118]
10Polyvinyl alcoholSynthetic-Antimicrobial coating[119]
11PolyanilineSynthetic61.6 nmAntibacterial[120]
12Poly(3-hydroxybutyrate-co-3-hydroxyvalerate)Synthetic3.5, 25 nmFood packaging and food contact surface applications[121]
13Multiwalled carbon nanotubes (MWCNTs)Synthetic-Food packaging[122]
14Hydroxyethyl cellulose, carboxymethyl chitosan (CMCS) composite/filmSynthetic-Antibacterial, food packaging[123]
15Graphene oxide, composite resinsSynthetic-Antimicrobial[124]
16Mahua-oil-based polyurethane/chitosan/nano ZnO compositeSynthetic30 nmFood packaging[125]
17Polyvinyl (alcohol)/chitosan/nano zinc oxide hydrogelsSynthetic30 nmWound healing[126]
18AZO-Np in Guar gum/polyvinyl alcohol composite fiber matsSemisyntheticLess than 50 nmAntibacterial[127]
19Chitosan–alginate–gelatin and chitosan–bentonite–gelatin films with ZnONaturalNanoscaledSkin burn healing[128]
20Aminoalkylsilane-grafted bacterial nanocelluloseSynthetic30 nmMultifunctional wound dressing, antibacterial[129]
21PVA and xylan, nanoscaled ZnOSynthetic-Bacteriostatic films[130]
22Oxidized sodium alginate and its electrospun bio-hybridsSynthetic-Wound healing[130]

3.4. Gold-Based Nanocomposites

Gold nanoparticles have been studied for their antibacterial properties. Gold nanoparticles can damage bacterial DNA and inhibit bacterial growth. Gold-based nanocomposites have potential applications in medical devices and sensors.
Gold nanoparticles have emerged as a primary candidate in a multitude of fields, such as nanobiotechnology, tissue engineering, and drug delivery, due to their excellent properties. Their strong affinity with organic species and high electrical conductivity make them suitable for applications in chemical sensing and drug delivery technologies.
In therapeutic, medical, and pharmaceutical fields, gold nanoparticles are used for a variety of purposes. Gold nanorods have proven to be effective in photothermal tumor therapy [131], while biosensors have been developed using gold nanoparticles to detect Bacillus anthrax [132]. Gold nanostars are also used for cancer therapy [132,133] and biological labeling [134]. Nanobelt gold nanoparticles serve as transducers and resonators and can be used as nanoscale sensors in medical approaches [132,135]. Nanoshells have potential for photonic crystals, fluorescent diagnostics, improving photoluminescent labels, catalysis, avoiding photodegradation, research in bioconjugates, and chemical and colloidal studies [136]. Additionally, gold nanoparticles are now widely used as antibacterial agents against a broad range of microorganisms [137] due to their biocompatibility, low cost, simple production, and high impact. Gold nanoparticle nanocomposites offer several advantages [138,139]. As shown in Table 4, gold–polymer nanocomposites are used as antimicrobials and for wound healing.

4. Synthesis of Metal–Polymeric Nanocomposites

Various approaches produce functional polymer composites containing metal nanoparticles (Figure 3), including reduction of a metal salt [154], ball milling [155], plasma polymerization [156], co-evaporation [157], and co-sputtering [158]. The goal is to achieve good control of the metal filling factor, filling factor profile, and particle composition [16,17,18,19,20,21,22,23,24].

4.1. In Situ Polymerization

The in-situ polymerization method involves the simultaneous polymerization of monomers and the reduction of metal ions to form copper nanoparticles within the polymer matrix (Figure 4). One common in situ polymerization methods is emulsion polymerization [159]. This method involves the dispersion of a hydrophobic monomer in water using surfactants. Metal salts are then added to the emulsion, followed by the addition of a reducing agent to initiate polymerization and nanoparticles formation [160,161].

4.2. Chemical Reduction

The reduction of chemically bound atoms in nonpolar media or metal cations in polar media (usually aqueous) is the most commonly used chemical method for obtaining metal–polymer nanocomposites [162]. This method offers several benefits such as speed, simplicity, and cost-effectiveness and enables the production of NPs with well-controlled sizes and shapes, facilitated by reducing and stabilizing agents that prevent NP agglomeration. The size and distribution of particles are influenced by factors such as redox potential, temperature, reagent and additive concentrations, and solvent type. It is worth noting that the in situ synthesis of MPNs also relies on the molecular weight, concentration, and chemical properties of the polymer. Additionally, the hydrophilicity, hydrophobicity, or amphiphilicity of the polymers, along with their subsequent configuration in a solvent, also have a crucial impact [163].

4.2.1. Engineering Polymers

Several studies have explored the use of various polymers to immobilize silver nanoparticles for their antimicrobial properties. One study described the coating of PET fibers with sericin-capped silver nanoparticles, where the fibers were first modified with a silane coupling agent and then coated using a dip-coating method [164]. Another study discussed the synthesis of chitosan-silk fibroin nanofibers containing silver nanoparticles for use as an antibacterial wound dressing, where the AgNPs were incorporated via electrospinning along with curcumin for sustained release [165].

4.2.2. Carbohydrates and Biopolymers

AgNPs can be combined with polymers using lignin as a stabilizer, which is a natural polymer found in plants. Lignin can be modified to make it compatible with other materials. In a study highlighted by the authors, lignin was used to stabilize AgNPs, resulting in a material with good antibacterial activity against E. coli [166]. The addition of organo-clay improved the interfacial adhesion between the AgNPs and the polymer matrix, resulting in a better dispersion and improved mechanical properties of the nanocomposite films. Thymol, which was also dispersed in the polymer matrix, improved the thermal stability and antioxidant properties of the films [167].

4.2.3. Dendrimers as Templates and Hosts for Metal Nanoparticles

Two studies described different methods for synthesizing hybrid nanoparticles. One study used a DNA dendrimer template to control the size and shape of copper nanoparticles [168], while the other study described the grafting of a PPI dendrimer onto gold nanoparticles using a multistep process. Both methods involved the use of functional groups for subsequent modifications [169].

4.3. Electrospinning

Electrospinning is a popular method for synthesizing metal–polymer nanocomposites, where an electric field is applied to a polymer solution or melt, resulting in the formation of nanofibers with diameters in the nanometer range. The incorporation of metal nanoparticles or nanowires into the polymer matrix during the electrospinning process enables the creation of nanocomposites with uniform metal distribution. These materials offer improved mechanical, electrical, and optical properties, making them attractive for various applications, including sensors, catalysis, and energy storage. One paper described a synthesis method for producing flexible nylon–ZnO core–shell nanofiber mats with enhanced photocatalytic activity. The method involves electrospinning to produce the nylon core fibers, followed by atomic layer deposition (ALD) to coat the ZnO shell onto the nylon fibers. The resulting nanofiber mats exhibit high flexibility and mechanical strength due to the presence of the nylon core, and the ZnO shell provides photocatalytic activity for potential environmental applications [170].
Another method involved dissolving PVP in a solvent system, generating gold nanoparticles by laser ablation in a PVP solution, and then electrospinning the resulting PVP/Au NP solution. The nanofibers were deposited on a cylindrical metal collector, and the electrospinning was carried out under controlled conditions. For comparison, electrospinning of PVP in the solvent system without laser ablation was also performed [171].

4.4. Layer-by-Layer Assembly

The layer-by-layer (LbL) assembly technique is a versatile and powerful method used to synthesize polymer nanocomposites by sequentially depositing oppositely charged species [172]. The technique has been used to fabricate various nanocomposites with tunable compositions and structures, such as a humidity sensor Co3O4/PSS [173], a bio-interface GO/PAH with excellent mechanical properties [174], a gas sensor based on polypyrrole (PPy)/TiO2 [175], a supercapacitor halloysite/polyaniline, and a foam coating nanoclay/PEI with customizable properties. The LbL assembly technique is capable of achieving nanoscale precision over the thickness and high content of well-dispersed nanomaterials. By incorporating nanomaterials into the polymer matrix, the physicochemical properties of polymer composites can be modified or altered [176].
In the field of LbL assembly, nanoparticles have evolved from being simple additives to become key components that drive the development of various applications, including smart coatings and drug delivery systems. Despite some disadvantages, the advantages of using nanoparticles in LbL assembly outweigh them, and future trends suggest that they are indispensable for achieving advanced coatings with controllable properties and responsiveness to various stimuli [177].
In a recent study, LbL assembly was employed to generate nanoscale antioxidant coatings using cerium oxide nanoparticles (CONPs) and alginate on various materials. The uniformity and scale of the LbL assembly were characterized using ellipsometry, which demonstrated predictable multilayer formation. The versatility of CONP coatings on three-dimensional surfaces was explored by coating silica and alginate microbeads with increasing numbers of alternating layers of CONP/alginate. Confocal microscopy images showed ultrathin coatings formed onto the different spherical materials, with increasing fluorescence intensity correlating with increasing layer number. The resulting nanoscale antioxidant coatings were able to protect encapsulated cells from reactive oxide species (ROS)-mediated damage while preserving capsule permeability and cellular function. This approach provides retention of catalytic activity, control of localization, and mitigation of cytotoxicity, making it a promising strategy for a variety of applications in tissue engineering and drug delivery [178].
Another study described a spray-assisted method for the alignment of silver nanowires in LbL assembled films. The method involves the use of a spray gun to deposit the LbL films onto a substrate while simultaneously applying a magnetic field to align the silver nanowires. The authors showed that this approach could result in highly anisotropic nanocomposite films with controlled thickness and composition [179].
Figure 5 illustrates the process of layer-by-layer assembly mechanisms, which demonstrate the alternate deposition of positively and negatively charged components onto a substrate. The LBL assembly allows for the introduction of various nanospecies, such as inorganic nanoparticles (NPs) and polymers, at the same time with high loading in the assembled layers. The assembly of each layer can be easily modulated to achieve the optimal structure for a specific application, and the prepared layers can be assembled with controlled variable thicknesses for generating biocompatible coatings. The ability of LBL to control the coating thickness, properties of the nanocomponents, and economic use of raw materials make the assembly tool greatly superior to other methods. This approach provides the opportunity to combine the electronic, optical, and magnetic properties of inorganic nanostructures with unique physical responses. The resulting LBL-based thin films have been applied to various applications, including drug delivery systems and thin membranes.

4.5. Template Synthesis

Template synthesis of metal–polymer nanocomposites is a powerful approach that allows the production of highly ordered nanoscale hybrid materials with unique properties. The method involves the use of a template, usually a self-assembled monolayer or a block copolymer, to direct the formation of the metal nanoparticles within the polymer matrix.
The template provides a highly ordered nanoscale architecture, which enables the control of the size, shape, and distribution of the metal nanoparticles within the polymer matrix. This control over the metal nanoparticle morphology leads to improved mechanical, electrical, and optical properties, making them useful in a wide range of applications [180]. There exist two primary categories of template-based synthesis methods, namely, the hard template approach, which employs a physical mold for the growth of nanostructures within composite polymers [181], and the soft template approach, which utilizes the self-assembly of composite polymers [182] (Figure 6).

4.5.1. Hard Template Approach

The hard template technique is a versatile and effective method for synthesizing metal–polymer nanocomposites. This method involves using a physical template, such as colloidal nanoparticles or nanosized channels, to control the shape and size of the resulting nanocomposites. The hard template technique can involve both chemical and electrochemical polymerization processes, depending on the specific reaction parameters. Colloidal nanoparticles have been commonly used as templates for synthesizing 0D metal–polymer nanocomposites, with the resulting core–shell structures highly dependent on the size of the nanoparticles. This method offers advantages such as ease of synthesis, even size dispersion, and large quantity availability [183].

4.5.2. Soft Template Approach

Soft template synthesis is a technique used to fabricate conducting polymer nanocomposites (CPNCs) by confining CP polymerization within surfactant micelles. This method is cheaper and less intricate than hard template synthesis and can produce CPNCs in large quantities. The morphology and size of the CPNCs depend on parameters such as microstructure, morphology, and concentration. Cationic surfactants such as octyltrimethyl ammonium bromide (OTAB), decyltrimethyl ammonium bromide (DTAB), and cetyltrimethyl ammonium bromide (CTAB) are used in micelle formation [184]. Polypyrrole (PPy) nanostructures of different shapes can be produced by varying the concentration of monomer and surfactant. Recently, PPy-based materials have been used to fabricate organic polymeric-oriented thermoelectric hybrid materials and composites, as well as supercapattery attributes of CeO2 nanoparticles. The inclusion of Cu–Al2O3 nanoparticles in PPy/Cu–Al2O3 nanocomposites enhances their electrical conductivity and gas sensing abilities [185].

4.6. Coprecipitation

The coprecipitation method involves obtaining a uniform mixture of two or more cations in a solution during the precipitation reaction. This method is widely used for synthesizing composites that contain multiple metal elements. The coprecipitation reaction consists of several processes, including nucleation, growth, coarsening, and agglomeration [186]. It is a significant technique for synthesizing composites containing multiple metal elements in a homogeneous solution during precipitation. This method is preferred over solid and vapor phase techniques due to its advantageous properties, such as short diffusion paths, high product purity, less agglomeration, simplicity, uniform and controllable particle size, and cost-effectiveness [9,116,187,188,189]. One recent example of this method is the fabrication of mesoporous iron–manganese bimetal oxide nanocomposites using the aeration coprecipitation technique. The method involves separately preparing a molar ratio of 2:1 of manganese sulfate and ferric chloride, mixing them at 70 °C with stirring at 50 rpm, adding 10% NH4OH to obtain a pH of 9, and keeping it at 50 rpm for 4 h to obtain the precipitated nanocomposites [190]. Similarly, a NiFeO4 nanocomposite was synthesized using the coprecipitation method, where a 1:1 molar ratio of FeSO4.7H2O and NiSO4.7H2O was used at a pH of 12, and hydrazine hydrate acted as the reducing agent [191].

4.7. Sol–Gel Process

The sol–gel method is a wet-chemical technique that employs a chemical solution or colloidal particles to create a cohesive network or gel. It uses metal alkoxides or metal chlorides as precursor materials. This process is also referred to as chemical solution deposition, which comprises various steps such as hydrolysis, polycondensation, gelation, aging, drying, densification, and crystallization. The sol–gel process is advantageous for producing nanocrystalline powders at relatively lower temperatures and pressures [186]. The size, morphology, dimensions, composition, and structure of nanomaterials play a crucial role in their various applications such as catalysis, sensing, and antibacterial. The sol–gel process is an important technique for the synthesis of nanomaterials due to its ease of operation, low-temperature reaction conditions, and ability to produce products with defined size, high purity, and homogeneity [192,193,194,195]. Ternary CuO/TiO2/ZnO nanocomposites were successfully synthesized using the sol–gel method. This transition metal oxide-based nanocomposite was found to enhance the photocatalytic degradation of methylene blue (MB), an organic dye [196]. In another study, a NiO/TiO2 nanocomposite was also prepared using the sol–gel method and was used for the degradation of MB organic dye [197]. Similarly, a ZnO/CuO nanomaterial was prepared using the sol–gel technique, where calculated amounts of zinc nitrate and copper oxide were dissolved and dispersed in the solvent, followed by the addition of PVA and stirring in an ultrasound bath at 353 K to obtain a homogeneous gel. The gel was then dried at 773 K for 8 h to obtain the required nanocomposite [198]. In yet another study, a TiO4–Ag nanocomposite was reported to exhibit antibacterial activity against E. coli and was also synthesized using the sol–gel route of synthesis [199].

4.8. Gamma Radiation

Gamma irradiation has become a highly effective method for synthesizing metal–polymer composites by creating random reducing and oxidizing agents with significant redox potentials in aqueous solutions [200], as depicted in Figure 3 and Figure 7. To enhance the concentration of these agents, additives such as alcohols or NO2 can be employed [201], initiating chemical reactions involving metal ions, metal complexes, monomers, and natural or synthetic polymers [202]. Gamma-irradiation-based synthesis of NPs in water generates highly reactive initiators for forming both inorganic and organic NPs. However, this approach is challenging because the reaction pathways are not always well established and can be quite complex. Two primary methods of synthesizing nanocomposites with gamma rays have been identified: one-step and two-step. One-step synthesis generates complete nanocomposites by combining a monomer or polymer with metallic ions and exposing them to gamma radiation in an inert atmosphere, while two-step synthesis involves first creating nuclei, either metallic or polymeric, and then completing the shell.
In summary, gamma ray synthesis of nanocomposites offers a versatile and effective way to create new materials with unique properties and potential applications in various fields. The choice of metallic nucleus, polymeric shell, and synthesis methodology can influence the type and shape of the resulting nanoparticles.

5. Challenges of Synthetic Approaches

During the synthesis of nanoparticles (NPs), one of the main difficulties is the unwanted aggregation of particles, which can sometimes be irreversible even when using a continuous distribution medium. Therefore, achieving a homogeneous redispersion of NPs is not always easy and can even lead to degradation and loss of functionality. To preserve the properties and purity of NPs, it is highly advantageous to use approaches that do not require solvent extraction.
To address this challenge, several methods have been developed to synthesize NPs atom by atom in gas, solid, or liquid phases [203]. In the liquid phase, metallic NPs are chemically synthesized using traditional methods such as reducing precipitation and coprecipitation. For organic NPs, crosslinking of organic macromolecules and polymerization are commonly used, with environmentally friendly synthesis methods involving micro/nanoemulsions assisted by photochemical and electrochemical methodologies. These methods require precursors, reducing agents, particle capping agents, stabilizers, and solvents [204].
However, common chemical processes in the synthesis of nanomaterials can be complicated by the contamination of NPs with chemical precursors and additives during the synthetic procedures. This aspect is particularly important for applications of NPs in electronics or healthcare. As the complexity of synthesis increases, so too do the costs, toxicity, operational safety, and environmental risks. Additionally, there is little known about the toxicity of NPs due to the lack of specific toxicity data.

6. Characterization of Metal–Polymeric Nanocomposites

Proper characterization of metallic and polymeric nanoparticles is crucial for understanding their characteristics, particularly in terms of their particle size, distribution, and shape [205]. Various techniques are typically employed for nanoparticles characterization, including transmission electron microscopy (TEM), atomic force microscopy (AFM), and scanning tunneling microscopic (STM) imaging to determine size and surface morphology. Additionally, electron spin resonance spectroscopy (ESR), nuclear magnetic resonance spectroscopy (NMR), extended X-ray absorption fine structure (EXAFS), X-ray, energy dispersive spectroscopy (EDS), and energy-dispersive X-ray spectroscopy (EDAX) are commonly used for determining structure and composition. Other methods such as UV/vis and infrared (IR) spectroscopy provide valuable information about the plasmonic surface as well as chemical groups and bonding on the particle. Furthermore, gel permeation chromatography (GPC) in combination with a static light scattering (SLS) detector is used to determine the average molecular weight of polymeric NPs. Overall, appropriate characterization methods are essential for gaining a comprehensive understanding of metallic and polymeric nanoparticles.

7. Applications and Limitations

The biomedical field is using nanotechnology to prevent microbial infection and the disease caused by the infection. This is carried out by using the atomic size of the active material. For example, AgNPs have strong antibacterial properties [206]. Nevertheless, they also have cytotoxic properties. This can be reduced by implanting or inserting the, with suitable polymer grids [207]. It (the presence of Ag in polymers) has also improved L929 cell attachment and development [208]. The antibacterial properties are due to the release of silver ions from the polymer nanocomposite [209]. Similarly, coating in light of polymer/Cu nanocomposites was due to the discharge of metal ions that exhibited significant, or prevented the development of, fungi and pathogenic microorganisms [210]. Controlled discharge of silver ions from metal–polymer nanocomposites can be used in various biomedical applications or fields, such as wound dressings, orthopedic implants [31], and catheters [32], paint or coating biomedical facilities or doctor’s facilities, households, or aviation. The antibacterial efficiency of the silver polymer nanocomposites on Gram-positive and Gram-negative bacteria was summarized above in this article. The other applications of polymer-based silver nanocomposites may be in the prevention of microbial infection in the food industries, in decontamination of water, in wound dressing, or inhibiting the growth of bacteria in injuries. Temperature-responsive nanohybrids synthesized form AgNO3 may be used in the preservation of solid substrates, as photothermal agents against microorganisms and against tumors, and in the development of surgically modified face masks [211]. Research using LDPE/Cu nanocomposite as intrauterine material has been explored [212]. Studies have demonstrated its effectiveness, and it may be utilized in place of conventional material. Gold-based nanocomposites have potential in medical devices, sensors, tissue engineering, and drug delivery. Gold nanorods have proven effective in photothermal tumor therapy [131], while biosensors have been developed using gold nanoparticles to detect Bacillus anthrax [132]. Gold nanostars are also used for cancer therapy [132,133] and biological labeling [134]. Nanobelt gold nanoparticles serve as transducers and resonators and can be used as nanoscale sensors in medical approaches [132,135].
The antibacterial property of the metal/polymer nanocomposites may be challenged by some of the properties of the nanoparticles, such as their ability to aggregate together due to high surface free energy; they may oxidize and pollute air, and these properties may diminish the antibacterial properties of the metal–polymer-based nanocomposites. One of the areas of concern for the nanocomposites is the toxicological aspects, and many studies are being carried out in this respect. This may hinder its biomedical application. Hence, in vitro studies are required as in vivo studies may be difficult to apprehend. There are no reports yet on the unfriendly behavior of the nanoparticles, owing to their difference in physical properties such as the shape, size, chemical nature (hydrophobicity, etc.), chemistry, and the method that has been used in the synthesis, the target organs, or the cells. Similarly, there are difficulties with the in vivo test. Hence, standard tests are required to assess the exact fate of these nanoparticles in vivo. The nanoparticles may not be targeted and recognized by the body’s immune system due to its size and may cause toxic response [213]. It seems that there is no standard test to evaluate the bioconsistent reaction to nanoparticles. Other things that must be taken into consideration are that nanoparticles are smaller than the chemical compounds, the study period must be longer, and alteration in the cytotoxicity test is needed. The physical and chemical parameters are also to be considered, such as size, shape, and the surface. These drawbacks are to be considered and assessed in all research; similarly, the toxicity posed by the polymer needs to be considered and assessed in the future to keep the threats below the recommended limits.
Future strategies for the development of more enhanced antibacterial nanocomposites depend on several factors. First, the use of a suitable polymer with therapeutic potential against multidrug-resistant microorganisms, which may be an alternate to the microbes resistant to antibiotics and against various microbes (fungi, viruses, etc.); secondly, to use green synthesis, plant-based polymers with antibacterial properties, and with fewer side effects, a new bactericidal nanocomposite may be developed. The other approach would include the use of smart polymer nanocomposites or stimuli-responsive polymer nanocomposites, which are those that contain one chemical and one physical property controlled in response to the stimulus. The stimulus may be external or internal. The external stimuli are concerned with temperature, light, electric current, and magnetic field strength. The internal stimuli are concerned with pH biological recognition, solvent type, and chemical recognition. Thus, by utilizing various properties of NPs, selective release of drug to the target can be achieved in the incorporation of external stimuli such as light pH, and sound; this may also increase the antibacterial efficacy and reduce side effects [214].
For any nanocomposites to be used in clinical trials, they must be studied extensively for their cytotoxic effects using effective in vivo studies. It is important to choose a suitable method for the synthesis of nanocomposites to be developed on a large scale for commercial purposes in order to substitute it for present existing antibacterial or antimicrobial agent. The advantages of the methods used in the synthesis are discussed above. More research investments are needed in this area to develop a potent antibacterial.

8. Conclusions

The metal-based and metal–polymer-based nanocomposites show significant antibacterial potential and have various biomedical applications. These metal-based nanocomposites have shown significant antibacterial properties; however, some of the metal-oxide-based nanocomposites are toxic to human cells, whereas most of them show negligible toxicity. As mentioned above, the goal is to develop good control of metal filling factor, filling factor profile, and particle composition. There are various approaches in developing a desired metal–polymer nanocomposite. The chemical-reduction method offers various benefits in synthesis such as speed, simple technique, and cost-effectiveness with well-controlled shapes and sizes of the NPs. However, there are several factors that must be kept in mind, to name a few: temperature, redox reaction, nature of the polymer, etc., which can be addressed. Techniques such as electrospin can be used to create nanofibers, which can be loaded with antibacterial activity and can be used as wound dressings to cure injuries. In the layer-by-layer technique, nanoscale precision is achieved with high content of well-dispersed nanomaterials, and the polymer can be modified by altering its physicochemical properties. It can be used in the coating of the drugs; despite its limitations, the benefits the LBL offers outweigh the limitations, which may explain the choice to develop nanocomposites using this technique. Template techniques enable one to develop well-structured nanocomposites with desired shape and size of nanoparticles, and their well-organized distribution in the polymer matrix of this method allows the coprecipitation that contains a mixture of one or more elements. The product is pure with the least contamination and desired shapes and sizes; with few agglomerations, this method is simple and less costly. Methods such as the gamma ray method offer an effective way to develop versatile nanocomposites with various properties that can be employed for various applications. The metal-based and metal–polymer-based nanocomposites have shown various benefits in biomedical applications and as potential antibacterials. Further extensive scientific research is required to test their toxicity and carry out the in vivo study, so that they can be used for clinical trials and can be marketed as therapeutically enhanced antibacterials.

Author Contributions

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

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Data available within the text.

Acknowledgments

We thank the Deanship, Faculty of Pharmacy, Allied Medical Sciences, and the Dean of Research and Higher Education of Al-Ahliyya Amman University, Amman, Jordan, for providing the necessary facilities. Article processing fees (APC) were supported by the Deanship of Scientific Research, Al-Ahliyya Amman University, Amman, Jordan.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kapoor, G.; Saigal, S.; Elongavan, A. Action and resistance mechanisms of antibiotics: A guide for clinicians. J. Anaesthesiol. Clin. Pharmacol. 2017, 33, 300–305. [Google Scholar] [CrossRef]
  2. Bao, Q.; Zhang, D.; Qi, P. Synthesis and characterization of silver nanoparticle and graphene oxide nanosheet composites as a bactericidal agent for water disinfection. J. Colloid Interface Sci. 2011, 360, 463–470. [Google Scholar] [CrossRef] [PubMed]
  3. Chandraker, K.; Nagwanshi, R.; Jadhav, S.K.; Ghosh, K.K.; Satnami, M.L. Antibacterial properties of amino acid functionalized silver nanoparticles decorated on graphene oxide sheets. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2017, 181, 47–54. [Google Scholar] [CrossRef]
  4. Prasad, K.; Lekshmi, G.S.; Ostrikov, K.; Lussini, V.; Blinco, J.; Mohandas, M.; Vasilev, K.; Bottle, S. Synergic bactericidal effects of reduced graphene oxide and silver nanoparticles against Gram-positive and Gram-negative bacteria. Sci. Rep. 2017, 7, 1591. [Google Scholar] [CrossRef] [PubMed]
  5. Yadollahi, M.; Gholamali, I.; Namazi, H.; Aghazadeh, M. Synthesis and characterization of antibacterial carboxymethyl cellulose/ZnO nanocomposite hydrogels. Int. J. Biol. Macromol. 2015, 74, 136–141. [Google Scholar] [CrossRef]
  6. Alswat, A.A.; Ahmad, M.B.; Saleh, T.A. Preparation and Characterization of Zeolite\Zinc Oxide-Copper Oxide Nanocomposite: Antibacterial Activities. Colloid Interface Sci. Commun. 2017, 16, 19–24. [Google Scholar] [CrossRef]
  7. Ghosh, T.; Das, A.B.; Jena, B.; Pradhan, C. Antimicrobial effect of silver zinc oxide (Ag-ZnO) nanocomposite particles. Front. Life Sci. 2015, 8, 47–54. [Google Scholar] [CrossRef]
  8. Prema, D.; Prakash, J.; Vignesh, S.; Veluchamy, P.; Ramachandran, C.; Samal, D.B.; Oh, D.-H.; Sahabudeen, S.; Devanand Venkatasubbu, G. Mechanism of inhibition of graphene oxide/zinc oxide nanocomposite against wound infection causing pathogens. Appl. Nanosci. 2020, 10, 827–849. [Google Scholar] [CrossRef]
  9. Thambidurai, S.; Gowthaman, P.; Venkatachalam, M.; Suresh, S. Enhanced bactericidal performance of nickel oxide-zinc oxide nanocomposites synthesized by facile chemical co-precipitation method. J. Alloys Compd. 2020, 830, 154642. [Google Scholar] [CrossRef]
  10. Huang, W.C.; Tsai, P.J.; Chen, Y.C. Functional gold nanoparticles as photothermal agents for selective-killing of pathogenic bacteria. Nanomedicine 2007, 2, 777–787. [Google Scholar] [CrossRef]
  11. Perni, S.; Piccirillo, C.; Pratten, J.; Prokopovich, P.; Chrzanowski, W.; Parkin, I.P.; Wilson, M. The antimicrobial properties of light-activated polymers containing methylene blue and gold nanoparticles. Biomaterials 2009, 30, 89–93. [Google Scholar] [CrossRef] [PubMed]
  12. Nirmala, R.; Park, H.M.; Kalpana, D.; Kang, H.S.; Navamathavan, R.; Lee, Y.S.; Kim, H.Y. Bactericidal activity and in vitro cytotoxicity assessment of hydroxyapatite containing gold nanoparticles. J. Biomed. Nanotechnol. 2011, 7, 342–350. [Google Scholar] [CrossRef] [PubMed]
  13. Regiel-Futyra, A.; Kus-Liśkiewicz, M.; Sebastian, V.; Irusta, S.; Arruebo, M.; Stochel, G.; Kyzioł, A. Development of noncytotoxic chitosan-gold nanocomposites as efficient antibacterial materials. ACS Appl. Mater. Interfaces 2015, 7, 1087–1099. [Google Scholar] [CrossRef] [PubMed]
  14. Mendoza, G.; Regiel-Futyra, A.; Andreu, V.; Sebastián, V.; Kyzioł, A.; Stochel, G.; Arruebo, M. Bactericidal Effect of Gold–Chitosan Nanocomposites in Coculture Models of Pathogenic Bacteria and Human Macrophages. ACS Appl. Mater. Interfaces 2017, 9, 17693–17701. [Google Scholar] [CrossRef]
  15. Chan, J.F.; Lau, S.K.; To, K.K.; Cheng, V.C.; Woo, P.C.; Yuen, K.Y. Middle East respiratory syndrome coronavirus: Another zoonotic betacoronavirus causing SARS-like disease. Clin. Microbiol. Rev. 2015, 28, 465–522. [Google Scholar] [CrossRef]
  16. Ellis, B.R.; Barrett, A.D. The enigma of yellow fever in East Africa. Rev. Med. Virol. 2008, 18, 331–346. [Google Scholar] [CrossRef]
  17. Andersen, K.G.; Rambaut, A.; Lipkin, W.I.; Holmes, E.C.; Garry, R.F. The proximal origin of SARS-CoV-2. Nat. Med. 2020, 26, 450–452. [Google Scholar] [CrossRef]
  18. Jiang, S.; Shi, Z.; Shu, Y.; Song, J.; Gao, G.F.; Tan, W.; Guo, D. A distinct name is needed for the new coronavirus. Lancet 2020, 395, 949. [Google Scholar] [CrossRef]
  19. Moritz, M.; Geszke-Moritz, M. The newest achievements in synthesis, immobilization and practical applications of antibacterial nanoparticles. Chem. Eng. J. 2013, 228, 596–613. [Google Scholar] [CrossRef]
  20. Lara, H.H.; Ayala-Nuñez, N.V.; Ixtepan-Turrent, L.; Rodriguez-Padilla, C. Mode of antiviral action of silver nanoparticles against HIV-1. J. Nanobiotechnol. 2010, 8, 1. [Google Scholar] [CrossRef]
  21. Lara, H.H.; Garza-Treviño, E.N.; Ixtepan-Turrent, L.; Singh, D.K. Silver nanoparticles are broad-spectrum bactericidal and virucidal compounds. J. Nanobiotechnol. 2011, 9, 30. [Google Scholar] [CrossRef] [PubMed]
  22. Kumar, R.; Nayak, M.; Sahoo, G.C.; Pandey, K.; Sarkar, M.C.; Ansari, Y.; Das, V.N.; Topno, R.K.; Madhukar, M.; Das, P. Iron oxide nanoparticles based antiviral activity of H1N1 influenza A virus. J. Infect. Chemother. 2019, 25, 325–329. [Google Scholar] [CrossRef]
  23. Simoncic, B.; Tomsic, B. Structures of Novel Antimicrobial Agents for Textiles—A Review. Text. Res. J. 2010, 80, 1721–1737. [Google Scholar] [CrossRef]
  24. Ye, S.; Shao, K.; Li, Z.; Guo, N.; Zuo, Y.; Li, Q.; Lu, Z.; Chen, L.; He, Q.; Han, H. Antiviral Activity of Graphene Oxide: How Sharp Edged Structure and Charge Matter. ACS Appl. Mater. Interfaces 2015, 7, 21571–21579. [Google Scholar] [CrossRef]
  25. Lishchynskyi, O.; Shymborska, Y.; Stetsyshyn, Y.; Raczkowska, J.; Skirtach, A.G.; Peretiatko, T.; Budkowski, A. Passive antifouling and active self-disinfecting antiviral surfaces. Chem. Eng. J. 2022, 446, 137048. [Google Scholar] [CrossRef]
  26. Dallas, P.; Sharma, V.K.; Zboril, R. Silver polymeric nanocomposites as advanced antimicrobial agents: Classification, synthetic paths, applications, and perspectives. Adv. Colloid Interface Sci. 2011, 166, 119–135. [Google Scholar]
  27. Hoseini-Alfatemi, S.M.; Karimi, A.; Armin, S.; Fakharzadeh, S.; Fallah, F.; Kalanaky, S. Antibacterial and antibiofilm activity of nanochelating based silver nanoparticles against several nosocomial pathogens. Appl. Organomet. Chem. 2018, 32, e4327. [Google Scholar]
  28. Al-Ramamneh, E.A.M.; Ghrair, A.M.; Shakya, A.K. Efficacy of Sterculia diversifolia Leaf Extracts: Volatile Compounds, Antioxidant and Anti-Inflammatory Activity, and Green Synthesis of Potential Antibacterial Silver Nanoparticles. Plants 2022, 11, 2492. [Google Scholar] [CrossRef] [PubMed]
  29. Tang, S.; Zheng, J. Antibacterial activity of silver nanoparticles: Structural effects. Adv. Healthc. Mater. 2018, 7, 1701503. [Google Scholar] [CrossRef]
  30. Innes, M.E.; Umraw, N.; Fish, J.S.; Gomez, M.; Cartotto, R.C. The use of silver coated dressings on donor site wounds: A prospective, controlled matched pair study. Burns 2001, 27, 621–627. [Google Scholar] [CrossRef]
  31. Zheng, Z.; Liu, P.; Zhang, X.; Zou, X.; Mei, X.; Zhang, S.; Zhang, S. Strategies to improve bioactive and antibacterial properties of polyetheretherketone (PEEK) for use as orthopedic implants. Mater. Today Bio 2022, 16, 100402. [Google Scholar] [CrossRef] [PubMed]
  32. Thokala, N.; Kealey, C.; Kennedy, J.; Brady, D.B.; Farrell, J. Comparative activity of silver-based antimicrobial composites for urinary catheters. Int. J. Antimicrob. Agents 2018, 52, 166–171. [Google Scholar] [CrossRef]
  33. Priyadarshini, S.; Gopinath, V.; Meera Priyadharsshini, N.; MubarakAli, D.; Velusamy, P. Synthesis of anisotropic silver nanoparticles using novel strain, Bacillus flexus and its biomedical application. Colloids Surf. B Biointerfaces 2013, 102, 232–237. [Google Scholar] [CrossRef]
  34. Yamanaka, M.; Hara, K.; Kudo, J. Bactericidal actions of a silver ion solution on Escherichia coli, studied by energy-filtering transmission electron microscopy and proteomic analysis. Appl. Environ. Microbiol. 2005, 71, 7589–7593. [Google Scholar] [CrossRef]
  35. Jung, W.K.; Koo, H.C.; Kim, K.W.; Shin, S.; Kim, S.H.; Park, Y.H. Antibacterial activity and mechanism of action of the silver ion in Staphylococcus aureus and Escherichia coli. Appl. Environ. Microbiol. 2008, 74, 2171–2178. [Google Scholar] [CrossRef] [PubMed]
  36. Slavin, Y.N.; Asnis, J.; Häfeli, U.O.; Bach, H. Metal nanoparticles: Understanding the mechanisms behind antibacterial activity. J. Nanobiotechnol. 2017, 15, 65. [Google Scholar] [CrossRef] [PubMed]
  37. Maurya, P.; Singh, S.; Naik, R.R.; Shakya, A.K. Biohazards of Nanomaterials. In Integrative Nanomedicine for New Therapies; Krishnan, A., Chuturgoon, A., Eds.; Springer International Publishing: Cham, Switzerland, 2020; pp. 39–70. [Google Scholar]
  38. Dakal, T.C.; Kumar, A.; Majumdar, R.S.; Yadav, V. Mechanistic Basis of Antimicrobial Actions of Silver Nanoparticles. Front. Microbiol. 2016, 7, 1831. [Google Scholar] [CrossRef] [PubMed]
  39. Yang, W.; Shen, C.; Ji, Q.; An, H.; Wang, J.; Liu, Q.; Zhang, Z. Food storage material silver nanoparticles interfere with DNA replication fidelity and bind with DNA. Nanotechnology 2009, 20, 085102. [Google Scholar] [CrossRef]
  40. Zarnegar, Z.; Safari, J.; Zahraei, Z. Design, synthesis and antimicrobial evaluation of silver decorated magnetic polymeric nanocomposites. Nano-Struct. Nano-Objects 2019, 19, 100368. [Google Scholar] [CrossRef]
  41. Hasan, N.; Cao, J.; Lee, J.; Hlaing, S.P.; Oshi, M.A.; Naeem, M.; Ki, M.H.; Lee, B.L.; Jung, Y.; Yoo, J.W. Bacteria-Targeted Clindamycin Loaded Polymeric Nanoparticles: Effect of Surface Charge on Nanoparticle Adhesion to MRSA, Antibacterial Activity, and Wound Healing. Pharmaceutics 2019, 11, 236. [Google Scholar] [CrossRef]
  42. Oliani, W.L.; Parra, D.F.; Komatsu, L.G.H.; Lincopan, N.; Rangari, V.K.; Lugao, A.B. Fabrication of polypropylene/silver nanocomposites for biocidal applications. Mater. Sci. Eng. C 2017, 75, 845–853. [Google Scholar] [CrossRef] [PubMed]
  43. Rehan, M.; Nada, A.A.; Khattab, T.A.; Abdelwahed, N.A.; Abou El-Kheir, A.A. Development of multifunctional polyacrylonitrile/silver nanocomposite films: Antimicrobial activity, catalytic activity, electrical conductivity, UV protection and SERS-active sensor. J. Mater. Res. Technol. 2020, 9, 9380–9394. [Google Scholar]
  44. Narayanan, K.B.; Han, S.S. Dual-crosslinked poly (vinyl alcohol)/sodium alginate/silver nanocomposite beads–A promising antimicrobial material. Food Chem. 2017, 234, 103–110. [Google Scholar] [CrossRef]
  45. Spagnol, C.; Fragal, E.H.; Pereira, A.G.; Nakamura, C.V.; Muniz, E.C.; Follmann, H.D.; Silva, R.; Rubira, A.F. Cellulose nanowhiskers decorated with silver nanoparticles as an additive to antibacterial polymers membranes fabricated by electrospinning. J. Colloid Interface Sci. 2018, 531, 705–715. [Google Scholar] [CrossRef]
  46. Xie, Y.; Liao, X.; Zhang, J.; Yang, F.; Fan, Z. Novel chitosan hydrogels reinforced by silver nanoparticles with ultrahigh mechanical and high antibacterial properties for accelerating wound healing. Int. J. Biol. Macromol. 2018, 119, 402–412. [Google Scholar] [CrossRef]
  47. Hajji, S.; Khedir, S.B.; Hamza-Mnif, I.; Hamdi, M.; Jedidi, I.; Kallel, R.; Boufi, S.; Nasri, M. Biomedical potential of chitosan-silver nanoparticles with special reference to antioxidant, antibacterial, hemolytic and in vivo cutaneous wound healing effects. Biochim. Et Biophys. Acta (BBA)-Gen. Subj. 2019, 1863, 241–254. [Google Scholar] [CrossRef]
  48. Ye, H.; Cheng, J.; Yu, K. In situ reduction of silver nanoparticles by gelatin to obtain porous silver nanoparticle/chitosan composites with enhanced antimicrobial and wound-healing activity. Int. J. Biol. Macromol. 2019, 121, 633–642. [Google Scholar] [CrossRef]
  49. Hernández-Rangel, A.; Silva-Bermudez, P.; Espana-Sanchez, B.; Luna-Hernández, E.; Almaguer-Flores, A.; Ibarra, C.; Garcia-Perez, V.; Velasquillo, C.; Luna-Barcenas, G. Fabrication and in vitro behavior of dual-function chitosan/silver nanocomposites for potential wound dressing applications. Mater. Sci. Eng. C 2019, 94, 750–765. [Google Scholar]
  50. Suteewong, T.; Wongpreecha, J.; Polpanich, D.; Jangpatarapongsa, K.; Kaewsaneha, C.; Tangboriboonrat, P. PMMA particles coated with chitosan-silver nanoparticles as a dual antibacterial modifier for natural rubber latex films. Colloids Surf. B Biointerfaces 2019, 174, 544–552. [Google Scholar] [CrossRef]
  51. Lee, D.; Lee, S.J.; Moon, J.-H.; Kim, J.H.; Heo, D.N.; Bang, J.B.; Lim, H.-N.; Kwon, I.K. Preparation of antibacterial chitosan membranes containing silver nanoparticles for dental barrier membrane applications. J. Ind. Eng. Chem. 2018, 66, 196–202. [Google Scholar] [CrossRef]
  52. Bahrami, A.; Rezaei Mokarram, R.; Sowti Khiabani, M.; Ghanbarzadeh, B.; Salehi, R. Physico-mechanical and antimicrobial properties of tragacanth/hydroxypropyl methylcellulose/beeswax edible films reinforced with silver nanoparticles. Int. J. Biol. Macromol. 2019, 129, 1103–1112. [Google Scholar] [CrossRef]
  53. Haider, M.K.; Ullah, A. Fabricating Antibacterial and Antioxidant Electrospun Hydrophilic Polyacrylonitrile Nanofibers Loaded with AgNPs by Lignin-Induced In-Situ Method. Polymers 2021, 13, 748. [Google Scholar] [CrossRef]
  54. Hsueh, Y.H.; Hsieh, C.T.; Chiu, S.T.; Tsai, P.H.; Liu, C.Y.; Ke, W.J. Antibacterial Property of Composites of Reduced Graphene Oxide with Nano-Silver and Zinc Oxide Nanoparticles Synthesized Using a Microwave-Assisted Approach. Int. J. Mol. Sci. 2019, 20, 5394. [Google Scholar] [CrossRef]
  55. Liu, C.; Ling, J.; Yang, L.Y.; Ouyang, X.K.; Wang, N. Chitosan-based carbon nitride-polydopamine-silver composite dressing with antibacterial properties for wound healing. Carbohydr. Polym. 2023, 303, 120436. [Google Scholar] [CrossRef]
  56. Massey, S.; Iqbal, F.; Rehman, A.U.; Iqbal, M.S. Preparation, characterization and biological evaluation of silver nanoparticles and drug loaded composites for wound dressings formed from Lallemantia royleana seeds’ mucilage. J. Biomater. Sci. Polym. Ed. 2022, 33, 481–498. [Google Scholar] [CrossRef] [PubMed]
  57. Rather, A.H.; Khan, R.S.; Wani, T.U.; Rafiq, M.; Jadhav, A.H.; Srinivasappa, P.M.; Abdal-Hay, A.; Sultan, P.; Rather, S.U.; Macossay, J.; et al. Polyurethane and cellulose acetate micro-nanofibers containing rosemary essential oil, and decorated with silver nanoparticles for wound healing application. Int. J. Biol. Macromol. 2023, 226, 690–705. [Google Scholar] [CrossRef] [PubMed]
  58. Khan, M.U.A.; Abd Razak, S.I.; Mehboob, H.; Abdul Kadir, M.R.; Anand, T.J.S.; Inam, F.; Shah, S.A.; Abdel-Haliem, M.E.F.; Amin, R. Synthesis and Characterization of Silver-Coated Polymeric Scaffolds for Bone Tissue Engineering: Antibacterial and In Vitro Evaluation of Cytotoxicity and Biocompatibility. ACS Omega 2021, 6, 4335–4346. [Google Scholar] [CrossRef] [PubMed]
  59. Zhou, M.; Lin, F.; Li, W.; Shi, L.; Li, Y.; Shan, G. Development of nanosilver doped carboxymethyl chitosan-polyamideamine alginate composite dressing for wound treatment. Int. J. Biol. Macromol. 2021, 166, 1335–1351. [Google Scholar] [CrossRef]
  60. Stetsyshyn, Y.; Awsiuk, K.; Kusnezh, V.; Raczkowska, J.; Jany, B.R.; Kostruba, A.; Harhay, K.; Ohar, H.; Lishchynskyi, O.; Shymborska, Y.; et al. Shape-Controlled synthesis of silver nanoparticles in temperature-responsive grafted polymer brushes for optical applications. Appl. Surf. Sci. 2019, 463, 1124–1133. [Google Scholar] [CrossRef]
  61. Wichai, S.; Chuysinuan, P.; Chaiarwut, S.; Ekabutr, P.; Supaphol, P. Development of bacterial cellulose/alginate/chitosan composites incorporating copper (II) sulfate as an antibacterial wound dressing. J. Drug Deliv. Sci. Technol. 2019, 51, 662–671. [Google Scholar] [CrossRef]
  62. 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]
  63. Halbus, A.F.; Horozov, T.S.; Paunov, V.N. Strongly enhanced antibacterial action of copper oxide nanoparticles with boronic acid surface functionality. ACS Appl. Mater. Interfaces 2019, 11, 12232–12243. [Google Scholar] [CrossRef]
  64. Bezza, F.A.; Tichapondwa, S.M.; Chirwa, E.M. Fabrication of monodispersed copper oxide nanoparticles with potential application as antimicrobial agents. Sci. Rep. 2020, 10, 16680. [Google Scholar] [CrossRef]
  65. Palza, H. Antimicrobial polymers with metal nanoparticles. Int. J. Mol. Sci. 2015, 16, 2099–2116. [Google Scholar] [PubMed]
  66. Arjunan, N.; Singaravelu, C.M.; Kulanthaivel, J.; Kandasamy, J. A potential photocatalytic, antimicrobial and anticancer activity of chitosan-copper nanocomposite. Int. J. Biol. Macromol. 2017, 104, 1774–1782. [Google Scholar] [CrossRef] [PubMed]
  67. Jayaramudu, T.; Varaprasad, K.; Reddy, K.K.; Pyarasani, R.D.; Akbari-Fakhrabadi, A.; Amalraj, J. Chitosan-pluronic based Cu nanocomposite hydrogels for prototype antimicrobial applications. Int. J. Biol. Macromol. 2020, 143, 825–832. [Google Scholar] [CrossRef] [PubMed]
  68. Hasanin, M.; Al Abboud, M.A.; Alawlaqi, M.M.; Abdelghany, T.M.; Hashem, A.H. Ecofriendly synthesis of biosynthesized copper nanoparticles with starch-based nanocomposite: Antimicrobial, antioxidant, and anticancer activities. Biol. Trace Elem. Res. 2021, 200, 2099–2112. [Google Scholar] [CrossRef]
  69. Phan, D.-N.; Dorjjugder, N.; Saito, Y.; Khan, M.Q.; Ullah, A.; Bie, X.; Taguchi, G.; Kim, I.-S. Antibacterial mechanisms of various copper species incorporated in polymeric nanofibers against bacteria. Mater. Today Commun. 2020, 25, 101377. [Google Scholar]
  70. Castro Mayorga, J.L.; Fabra Rovira, M.J.; Cabedo Mas, L.; Sánchez Moragas, G.; Lagarón Cabello, J.M. Antimicrobial nanocomposites and electrospun coatings based on poly (3-hydroxybutyrate-co-3-hydroxyvalerate) and copper oxide nanoparticles for active packaging and coating applications. J. Appl. Polym. Sci. 2018, 135, 45673. [Google Scholar] [CrossRef]
  71. Ashjari, H.R.; Dorraji, M.S.S.; Fakhrzadeh, V.; Eslami, H.; Rasoulifard, M.H.; Rastgouy-Houjaghan, M.; Gholizadeh, P.; Kafil, H.S. Starch-based polyurethane/CuO nanocomposite foam: Antibacterial effects for infection control. Int. J. Biol. Macromol. 2018, 111, 1076–1082. [Google Scholar] [CrossRef]
  72. El Nahrawy, A.M.; Hammad, A.B.A.; Youssef, A.M.; Mansour, A.M.; Othman, A.M. Thermal, dielectric and antimicrobial properties of polystyrene-assisted/ITO:Cu nanocomposites. Appl. Phys. A 2019, 125, 46. [Google Scholar] [CrossRef]
  73. Araújo, I.M.S.; Silva, R.R.; Pacheco, G.; Lustri, W.R.; Tercjak, A.; Gutierrez, J.; Júnior, J.R.S.; Azevedo, F.H.C.; Figuêredo, G.S.; Vega, M.L.; et al. Hydrothermal synthesis of bacterial cellulose–copper oxide nanocomposites and evaluation of their antimicrobial activity. Carbohydr. Polym. 2018, 179, 341–349. [Google Scholar] [CrossRef]
  74. Marković, D.; Deeks, C.; Nunney, T.; Radovanović, Ž.; Radoičić, M.; Šaponjić, Z.; Radetić, M. Antibacterial activity of Cu-based nanoparticles synthesized on the cotton fabrics modified with polycarboxylic acids. Carbohydr. Polym. 2018, 200, 173–182. [Google Scholar] [CrossRef]
  75. Nouri, A.; Yaraki, M.T.; Ghorbanpour, M.; Agarwal, S.; Gupta, V.K. Enhanced Antibacterial effect of chitosan film using Montmorillonite/CuO nanocomposite. Int. J. Biol. Macromol. 2018, 109, 1219–1231. [Google Scholar] [CrossRef] [PubMed]
  76. Chen, M.; Li, Z.; Chen, L. Highly antibacterial rGO/Cu2O nanocomposite from a biomass precursor: Synthesis, performance, and mechanism. Nano Mater. Sci. 2020, 2, 172–179. [Google Scholar] [CrossRef]
  77. Logpriya, S.; Bhuvaneshwari, V.; Vaidehi, D.; SenthilKumar, R.P.; Nithya Malar, R.S.; Pavithra Sheetal, B.; Amsaveni, R.; Kalaiselvi, M. Preparation and characterization of ascorbic acid-mediated chitosan–copper oxide nanocomposite for anti-microbial, sporicidal and biofilm-inhibitory activity. J. Nanostruct. Chem. 2018, 8, 301–309. [Google Scholar] [CrossRef]
  78. Sportelli, M.C.; Izzi, M.; Volpe, A.; Lacivita, V.; Clemente, M.; Di Franco, C.; Conte, A.; Del Nobile, M.A.; Ancona, A.; Cioffi, N. A new nanocomposite based on LASiS-generated CuNPs as a preservation system for fruit salads. Food Packag. Shelf Life 2019, 22, 100422. [Google Scholar] [CrossRef]
  79. Solairaj, D.; Rameshthangam, P.; Arunachalam, G. Anticancer activity of silver and copper embedded chitin nanocomposites against human breast cancer (MCF-7) cells. Int. J. Biol. Macromol. 2017, 105, 608–619. [Google Scholar] [CrossRef]
  80. Tabesh, E.; Salimijazi, H.R.; Kharaziha, M.; Mahmoudi, M.; Hejazi, M. Development of an in-situ chitosan-copper nanoparticle coating by electrophoretic deposition. Surf. Coat. Technol. 2019, 364, 239–247. [Google Scholar] [CrossRef]
  81. Muthulakshmi, L.; Rajini, N.; Nellaiah, H.; Kathiresan, T.; Jawaid, M.; Rajulu, A.V. Preparation and properties of cellulose nanocomposite films with in situ generated copper nanoparticles using Terminalia catappa leaf extract. Int. J. Biol. Macromol. 2017, 95, 1064–1071. [Google Scholar] [CrossRef]
  82. Abdollahi, Z.; Zare, E.N.; Salimi, F.; Goudarzi, I.; Tay, F.R.; Makvandi, P. Bioactive Carboxymethyl Starch-Based Hydrogels Decorated with CuO Nanoparticles: Antioxidant and Antimicrobial Properties and Accelerated Wound Healing In Vivo. Int. J. Mol. Sci. 2021, 22, 2531. [Google Scholar] [CrossRef] [PubMed]
  83. Abu-Elala, N.M.; AbuBakr, H.O.; Khattab, M.S.; Mohamed, S.H.; El-Hady, M.A.; Ghandour, R.A.; Morsi, R.E. Aquatic environmental risk assessment of chitosan/silver, copper and carbon nanotube nanocomposites as antimicrobial agents. Int. J. Biol. Macromol. 2018, 113, 1105–1115. [Google Scholar] [CrossRef] [PubMed]
  84. Ahmed, S.B.; Mohamed, H.I.; Al-Subaie, A.M.; Al-Ohali, A.I.; Mahmoud, N.M.R. Investigation of the antimicrobial activity and hematological pattern of nano-chitosan and its nano-copper composite. Sci. Rep. 2021, 11, 9540. [Google Scholar] [CrossRef]
  85. Al-Enizi, A.M.; Ahamad, T.; Al-Hajji, A.B.; Ahmed, J.; Chaudhary, A.A.; Alshehri, S.M. Cellulose gum and copper nanoparticles based hydrogel as antimicrobial agents against urinary tract infection (UTI) pathogens. Int. J. Biol. Macromol. 2018, 109, 803–809. [Google Scholar] [CrossRef] [PubMed]
  86. Al-Saeedi, S.I.; Al-Kadhi, N.S.; Al-Senani, G.M.; Almaghrabi, O.A.; Nafady, A. Antibacterial potency, cell viability and morphological implications of copper oxide nanoparticles encapsulated into cellulose acetate nanofibrous scaffolds. Int. J. Biol. Macromol. 2021, 182, 464–471. [Google Scholar] [CrossRef]
  87. Bagchi, B.; Salvadores Fernandez, C.; Bhatti, M.; Ciric, L.; Lovat, L.; Tiwari, M.K. Copper nanowire embedded hypromellose: An antibacterial nanocomposite film. J. Colloid Interface Sci. 2022, 608, 30–39. [Google Scholar] [CrossRef]
  88. des Ligneris, E.; Dumée, L.F.; Al-Attabi, R.; Castanet, E.; Schütz, J.; Kong, L. Mixed Matrix Poly(Vinyl Alcohol)-Copper Nanofibrous Anti-Microbial Air-Microfilters. Membranes 2019, 9, 87. [Google Scholar] [CrossRef]
  89. He, W.; Huang, X.; Zheng, Y.; Sun, Y.; Xie, Y.; Wang, Y.; Yue, L. In situ synthesis of bacterial cellulose/copper nanoparticles composite membranes with long-term antibacterial property. J. Biomater. Sci. Polym. Ed. 2018, 29, 2137–2153. [Google Scholar] [CrossRef]
  90. Jayaramudu, T.; Varaprasad, K.; Pyarasani, R.D.; Reddy, K.K.; Akbari-Fakhrabadi, A.; Carrasco-Sánchez, V.; Amalraj, J. Hydroxypropyl methylcellulose-copper nanoparticle and its nanocomposite hydrogel films for antibacterial application. Carbohydr. Polym. 2021, 254, 117302. [Google Scholar] [CrossRef]
  91. Jayaramudu, T.; Varaprasad, K.; Pyarasani, R.D.; Reddy, K.K.; Kumar, K.D.; Akbari-Fakhrabadi, A.; Mangalaraja, R.V.; Amalraj, J. Chitosan capped copper oxide/copper nanoparticles encapsulated microbial resistant nanocomposite films. Int. J. Biol. Macromol. 2019, 128, 499–508. [Google Scholar] [CrossRef]
  92. Kruk, T.; Gołda-Cępa, M.; Szczepanowicz, K.; Szyk-Warszyńska, L.; Brzychczy-Włoch, M.; Kotarba, A.; Warszyński, P. Nanocomposite multifunctional polyelectrolyte thin films with copper nanoparticles as the antimicrobial coatings. Colloids Surf. B Biointerfaces 2019, 181, 112–118. [Google Scholar] [CrossRef]
  93. Li, M.; Liu, X.; Tan, L.; Cui, Z.; Yang, X.; Li, Z.; Zheng, Y.; Yeung, K.W.K.; Chu, P.K.; Wu, S. Noninvasive rapid bacteria-killing and acceleration of wound healing through photothermal/photodynamic/copper ion synergistic action of a hybrid hydrogel. Biomater. Sci. 2018, 6, 2110–2121. [Google Scholar] [CrossRef] [PubMed]
  94. Mehta, R.; Brahmbhatt, H.; Bhojani, G.; Mukherjee, M.; Bhattacharya, A. Poly(piperizinamide) with copper ion composite membranes: Application for mitigation of Hexaconazole from water and combat microbial contamination. J. Hazard. Mater. 2019, 376, 102–111. [Google Scholar] [CrossRef]
  95. Mugesh, S.; Arun, R.; Arunkumar, K.; Murugan, M. Synthesis of Biogenic Copper Nanoparticles Embedded in Graphene Oxide-Chitosan Composite and Its Anti-Bacterial and Cytotoxic Activities. J. Nanosci. Nanotechnol. 2019, 19, 2625–2632. [Google Scholar] [CrossRef]
  96. Prokhorov, E.; España-Sánchez, B.L.; Luna-Bárcenas, G.; Padilla-Vaca, F.; Cruz-Soto, M.E.; Vázquez-Lepe, M.O.; Kovalenko, Y.; Elizalde-Peña, E.A. Chitosan/copper nanocomposites: Correlation between electrical and antibacterial properties. Colloids Surf. B Biointerfaces 2019, 180, 186–192. [Google Scholar] [CrossRef] [PubMed]
  97. Pulit-Prociak, J.; Staroń, A.; Staroń, P.; Chmielowiec-Korzeniowska, A.; Drabik, A.; Tymczyna, L.; Banach, M. Preparation and of PVA-based compositions with embedded silver, copper and zinc oxide nanoparticles and assessment of their antibacterial properties. J. Nanobiotechnol. 2020, 18, 148. [Google Scholar] [CrossRef]
  98. Sathiyavimal, S.; Vasantharaj, S.; Kaliannan, T.; Pugazhendhi, A. Eco-biocompatibility of chitosan coated biosynthesized copper oxide nanocomposite for enhanced industrial (Azo) dye removal from aqueous solution and antibacterial properties. Carbohydr. Polym. 2020, 241, 116243. [Google Scholar] [CrossRef] [PubMed]
  99. Sathiyavimal, S.; Vasantharaj, S.; Kaliannan, T.; Garalleh, H.A.; Garaleh, M.; Brindhadevi, K.; Chi, N.T.L.; Sharma, A.; Pugazhendhi, A. Bio-functionalized copper oxide/chitosan nanocomposite using Sida cordifolia and their efficient properties of antibacterial, anticancer activity against on breast and lung cancer cell lines. Environ. Res. 2023, 218, 114986. [Google Scholar] [CrossRef]
  100. Sun, X.; Dong, M.; Guo, Z.; Zhang, H.; Wang, J.; Jia, P.; Bu, T.; Liu, Y.; Li, L.; Wang, L. Multifunctional chitosan-copper-gallic acid based antibacterial nanocomposite wound dressing. Int. J. Biol. Macromol. 2021, 167, 10–22. [Google Scholar] [CrossRef]
  101. Wang, W.B.; Clapper, J.C. Antibacterial Activity of Electrospun Polyacrylonitrile Copper Nanoparticle Nanofibers on Antibiotic Resistant Pathogens and Methicillin Resistant. Nanomaterials 2022, 12, 2139. [Google Scholar] [CrossRef]
  102. Yang, Y.; Dong, Z.; Li, M.; Liu, L.; Luo, H.; Wang, P.; Zhang, D.; Yang, X.; Zhou, K.; Lei, S. Graphene Oxide/Copper Nanoderivatives-Modified Chitosan/Hyaluronic Acid Dressings for Facilitating Wound Healing in Infected Full-Thickness Skin Defects. Int. J. Nanomed. 2020, 15, 8231–8247. [Google Scholar] [CrossRef] [PubMed]
  103. Kumar, V.; Menon, S.; Agarwal, H.; Gopalakrishnan, D. Characterization and optimization of bacterium isolated from soil samples for the production of siderophores. Resour. Effic. Technol. 2017, 3, 434–439. [Google Scholar] [CrossRef]
  104. Derewacz, D.K.; Goodwin, C.R.; McNees, C.R.; McLean, J.A.; Bachmann, B.O. Antimicrobial drug resistance affects broad changes in metabolomic phenotype in addition to secondary metabolism. Proc. Natl. Acad. Sci. USA 2013, 110, 2336–2341. [Google Scholar] [CrossRef]
  105. Landini, P.; Antoniani, D.; Burgess, J.G.; Nijland, R. Molecular mechanisms of compounds affecting bacterial biofilm formation and dispersal. Appl. Microbiol. Biotechnol. 2010, 86, 813–823. [Google Scholar] [CrossRef]
  106. Rai, M.; Yadav, A.; Gade, A. Silver nanoparticles as a new generation of antimicrobials. Biotechnol. Adv. 2009, 27, 76–83. [Google Scholar] [CrossRef]
  107. Tiwari, V.; Mishra, N.; Gadani, K.; Solanki, P.S.; Shah, N.A.; Tiwari, M. Mechanism of Anti-bacterial Activity of Zinc Oxide Nanoparticle Against Carbapenem-Resistant Acinetobacter baumannii. Front. Microbiol. 2018, 9, 1218. [Google Scholar] [CrossRef]
  108. Singh, T.A.; Sharma, A.; Tejwan, N.; Ghosh, N.; Das, J.; Sil, P.C. A state of the art review on the synthesis, antibacterial, antioxidant, antidiabetic and tissue regeneration activities of zinc oxide nanoparticles. Adv. Colloid Interface Sci. 2021, 295, 102495. [Google Scholar] [CrossRef]
  109. Rajabi, H.R.; Khani, O.; Shamsipur, M.; Vatanpour, V. High-performance pure and Fe3+-ion doped ZnS quantum dots as green nanophotocatalysts for the removal of malachite green under UV-light irradiation. J. Hazard. Mater. 2013, 250–251, 370–378. [Google Scholar] [CrossRef] [PubMed]
  110. Afrasiabi, S.; Bahador, A.; Partoazar, A. Combinatorial therapy of chitosan hydrogel-based zinc oxide nanocomposite attenuates the virulence of Streptococcus mutans. BMC Microbiol. 2021, 21, 62. [Google Scholar] [CrossRef]
  111. Lefatshe, K.; Muiva, C.M.; Kebaabetswe, L.P. Extraction of nanocellulose and in-situ casting of ZnO/cellulose nanocomposite with enhanced photocatalytic and antibacterial activity. Carbohydr. Polym. 2017, 164, 301–308. [Google Scholar] [CrossRef] [PubMed]
  112. Magesh, G.; Bhoopathi, G.; Nithya, N.; Arun, A.P.; Ranjith Kumar, E. Structural, morphological, optical and biological properties of pure ZnO and agar/zinc oxide nanocomposites. Int. J. Biol. Macromol. 2018, 117, 959–966. [Google Scholar] [CrossRef] [PubMed]
  113. Shahvalizadeh, R.; Ahmadi, R.; Davandeh, I.; Pezeshki, A.; Seyed Moslemi, S.A.; Karimi, S.; Rahimi, M.; Hamishehkar, H.; Mohammadi, M. Antimicrobial bio-nanocomposite films based on gelatin, tragacanth, and zinc oxide nanoparticles—Microstructural, mechanical, thermo-physical, and barrier properties. Food Chem. 2021, 354, 129492. [Google Scholar] [CrossRef]
  114. Preethi, S.; Abarna, K.; Nithyasri, M.; Kishore, P.; Deepika, K.; Ranjithkumar, R.; Bhuvaneshwari, V.; Bharathi, D. Synthesis and characterization of chitosan/zinc oxide nanocomposite for antibacterial activity onto cotton fabrics and dye degradation applications. Int. J. Biol. Macromol. 2020, 164, 2779–2787. [Google Scholar] [CrossRef] [PubMed]
  115. Bharathi, D.; Ranjithkumar, R.; Chandarshekar, B.; Bhuvaneshwari, V. Preparation of chitosan coated zinc oxide nanocomposite for enhanced antibacterial and photocatalytic activity: As a bionanocomposite. Int. J. Biol. Macromol. 2019, 129, 989–996. [Google Scholar] [CrossRef]
  116. Rojas, K.; Canales, D.; Amigo, N.; Montoille, L.; Cament, A.; Rivas, L.M.; Gil-Castell, O.; Reyes, P.; Ulloa, M.T.; Ribes-Greus, A.; et al. Effective antimicrobial materials based on low-density polyethylene (LDPE) with zinc oxide (ZnO) nanoparticles. Compos. Part B Eng. 2019, 172, 173–178. [Google Scholar] [CrossRef]
  117. Madhan, G.; Begam, A.A.; Varsha, L.V.; Ranjithkumar, R.; Bharathi, D. Facile synthesis and characterization of chitosan/zinc oxide nanocomposite for enhanced antibacterial and photocatalytic activity. Int. J. Biol. Macromol. 2021, 190, 259–269. [Google Scholar] [CrossRef] [PubMed]
  118. Motshekga, S.C.; Sinha Ray, S.; Maity, A. Synthesis and characterization of alginate beads encapsulated zinc oxide nanoparticles for bacteria disinfection in water. J. Colloid Interface Sci. 2018, 512, 686–692. [Google Scholar] [CrossRef] [PubMed]
  119. Jayakumar, A.; Radoor, S.; Nair, I.C.; Siengchin, S.; Parameswaranpillai, J.; Radhakrishnan, E.K. Lipopeptide and zinc oxide nanoparticles blended polyvinyl alcohol-based nanocomposite films as antimicrobial coating for biomedical applications. Process Biochem. 2021, 102, 220–228. [Google Scholar] [CrossRef]
  120. Mohsen, R.M.; Morsi, S.M.M.; Selim, M.M.; Ghoneim, A.M.; El-Sherif, H.M. Electrical, thermal, morphological, and antibacterial studies of synthesized polyaniline/zinc oxide nanocomposites. Polym. Bull. 2019, 76, 1–21. [Google Scholar] [CrossRef]
  121. Castro-Mayorga, J.L.; Fabra, M.J.; Pourrahimi, A.M.; Olsson, R.T.; Lagaron, J.M. The impact of zinc oxide particle morphology as an antimicrobial and when incorporated in poly(3-hydroxybutyrate-co-3-hydroxyvalerate) films for food packaging and food contact surfaces applications. Food Bioprod. Process. 2017, 101, 32–44. [Google Scholar] [CrossRef]
  122. Wen, Y.-H.; Tsou, C.-H.; de Guzman, M.R.; Huang, D.; Yu, Y.-Q.; Gao, C.; Zhang, X.-M.; Du, J.; Zheng, Y.-T.; Zhu, H.; et al. Antibacterial nanocomposite films of poly(vinyl alcohol) modified with zinc oxide-doped multiwalled carbon nanotubes as food packaging. Polym. Bull. 2022, 79, 3847–3866. [Google Scholar] [CrossRef]
  123. Cen, C.; Wang, F.; Wang, Y.; Li, H.; Fu, L.; Li, Y.; Chen, J.; Wang, Y. Design and characterization of an antibacterial film composited by hydroxyethyl cellulose (HEC), carboxymethyl chitosan (CMCS), and nano ZnO for food packaging. Int. J. Biol. Macromol. 2023, 231, 123203. [Google Scholar] [CrossRef] [PubMed]
  124. Farhangian, Z.; Alaghehmand, H.; Tashakkorian, H.; Mokhtarpour, F.; Davoodabadi, A. Antimicrobial effect of different physical and chemical compounds of zinc oxide and graphene oxide added to composite resins. Dent. Res. J. 2022, 19, 81. [Google Scholar]
  125. Sarojini, S.; Indumathi, M.P.; Rajarajeswari, G.R. Mahua oil-based polyurethane/chitosan/nano ZnO composite films for biodegradable food packaging applications. Int. J. Biol. Macromol. 2019, 124, 163–174. [Google Scholar] [CrossRef]
  126. Khorasani, M.T.; Joorabloo, A.; Adeli, H.; Mansoori-Moghadam, Z.; Moghaddam, A. Design and optimization of process parameters of polyvinyl (alcohol)/chitosan/nano zinc oxide hydrogels as wound healing materials. Carbohydr. Polym. 2019, 207, 542–554. [Google Scholar] [CrossRef]
  127. Lubambo, A.F.; Mattoso, N. In Situ Synthesis of AZO-Np in Guar Gum/PVOH Composite Fiber Mats for Potential Bactericidal Release. Polymers 2022, 14, 4983. [Google Scholar] [CrossRef]
  128. Nozari, M.; Gholizadeh, M.; Zahiri Oghani, F.; Tahvildari, K. Studies on novel chitosan/alginate and chitosan/bentonite flexible films incorporated with ZnO nano particles for accelerating dermal burn healing: In vivo and in vitro evaluation. Int. J. Biol. Macromol. 2021, 184, 235–249. [Google Scholar] [CrossRef]
  129. Shahriari-Khalaji, M.; Hu, G.; Chen, L.; Cao, Z.; Andreeva, T.; Xiong, X.; Krastev, R.; Hong, F.F. Functionalization of Aminoalkylsilane-Grafted Bacterial Nanocellulose with ZnO-NPs-Doped Pullulan Electrospun Nanofibers for Multifunctional Wound Dressing. ACS Biomater. Sci. Eng. 2021, 7, 3933–3946. [Google Scholar] [CrossRef] [PubMed]
  130. Wang, H.; Xue, T.; Wang, S.; Jia, X.; Cao, S.; Niu, B.; Guo, R.; Yan, H. Preparation, characterization and food packaging application of nano ZnO@Xylan/quaternized xylan/polyvinyl alcohol composite films. Int. J. Biol. Macromol. 2022, 215, 635–645. [Google Scholar] [CrossRef]
  131. von Maltzahn, G.; Park, J.H.; Agrawal, A.; Bandaru, N.K.; Das, S.K.; Sailor, M.J.; Bhatia, S.N. Computationally guided photothermal tumor therapy using long-circulating gold nanorod antennas. Cancer Res. 2009, 69, 3892–3900. [Google Scholar] [CrossRef]
  132. Gupta, A.; Pandey, S.; Yadav, J.S. A Review on Recent Trends in Green Synthesis of Gold Nanoparticles for Tuberculosis. Adv. Pharm. Bull. 2021, 11, 10–27. [Google Scholar] [CrossRef]
  133. Krawinkel, J.; Richter, U.; Torres-Mapa, M.L.; Westermann, M.; Gamrad, L.; Rehbock, C.; Barcikowski, S.; Heisterkamp, A. Optical and electron microscopy study of laser-based intracellular molecule delivery using peptide-conjugated photodispersible gold nanoparticle agglomerates. J. Nanobiotechnol. 2016, 14, 2. [Google Scholar] [CrossRef]
  134. Lin, C.A.; Yang, T.Y.; Lee, C.H.; Huang, S.H.; Sperling, R.A.; Zanella, M.; Li, J.K.; Shen, J.L.; Wang, H.H.; Yeh, H.I.; et al. Synthesis, characterization, and bioconjugation of fluorescent gold nanoclusters toward biological labeling applications. ACS Nano 2009, 3, 395–401. [Google Scholar] [CrossRef] [PubMed]
  135. Wang, Z.L. Functional oxide nanobelts: Materials, properties and potential applications in nanosystems and biotechnology. Annu. Rev. Phys. Chem. 2004, 55, 159–196. [Google Scholar] [CrossRef] [PubMed]
  136. Sikdar, D.; Rukhlenko, I.D.; Cheng, W.; Premaratne, M. Optimized gold nanoshell ensembles for biomedical applications. Nanoscale Res. Lett. 2013, 8, 142. [Google Scholar] [CrossRef]
  137. Shah, M.; Badwaik, V.; Kherde, Y.; Waghwani, H.K.; Modi, T.; Aguilar, Z.P.; Rodgers, H.; Hamilton, W.; Marutharaj, T.; Webb, C.; et al. Gold nanoparticles: Various methods of synthesis and antibacterial applications. Front. Biosci. 2014, 19, 1320–1344. [Google Scholar] [CrossRef] [PubMed]
  138. Dheyab, M.A.; Aziz, A.A.; Khaniabadi, P.M.; Jameel, M.S.; Ahmed, N.M.; Ali, A.T. Distinct advantages of using sonochemical over laser ablation methods for a rapid-high quality gold nanoparticles production. Mater. Res. Express 2021, 8, 015009. [Google Scholar] [CrossRef]
  139. Mobed, A.; Hasanzadeh, M.; Seidi, F. Anti-bacterial activity of gold nanocomposites as a new nanomaterial weapon to combat photogenic agents: Recent advances and challenges. RSC Adv. 2021, 11, 34688–34698. [Google Scholar] [CrossRef]
  140. Virgili, A.H.; Laranja, D.C.; Malheiros, P.S.; Pereira, M.B.; Costa, T.M.H.; de Menezes, E.W. Nanocomposite film with antimicrobial activity based on gold nanoparticles, chitosan and aminopropylsilane. Surf. Coat. Technol. 2021, 415, 127086. [Google Scholar] [CrossRef]
  141. Hashem, A.H.; Shehabeldine, A.M.; Ali, O.M.; Salem, S.S. Synthesis of chitosan-based gold nanoparticles: Antimicrobial and wound-healing activities. Polymers 2022, 14, 2293. [Google Scholar] [CrossRef]
  142. Hussein, M.A.M.; Grinholc, M.; Dena, A.S.A.; El-Sherbiny, I.M.; Megahed, M. Boosting the antibacterial activity of chitosan–gold nanoparticles against antibiotic–resistant bacteria by Punicagranatum L. extract. Carbohydr. Polym. 2021, 256, 117498. [Google Scholar] [CrossRef] [PubMed]
  143. Doghish, A.S.; Hashem, A.H.; Shehabeldine, A.M.; Sallam, A.-A.M.; El-Sayyad, G.S.; Salem, S.S. Nanocomposite based on gold nanoparticles and carboxymethyl cellulose: Synthesis, characterization, antimicrobial, and anticancer activities. J. Drug Deliv. Sci. Technol. 2022, 77, 103874. [Google Scholar] [CrossRef]
  144. Zhang, K.; Zhao, G. An Effective Wound Healing Material Based on Gold Incorporation into a Heparin-Polyvinyl Alcohol Nanocomposite: Enhanced In Vitro and In Vivo Care of Perioperative Period. J. Clust. Sci. 2022, 33, 1655–1665. [Google Scholar] [CrossRef]
  145. Adil, S.F.; Shaik, M.R.; Nasr, F.A.; Alqahtani, A.S.; Ahmed, M.Z.; Qamar, W.; Kuniyil, M.; Almutairi, A.; Alwarthan, A.; Siddiqui, M.R.H.; et al. Enhanced Apoptosis by Functionalized Highly Reduced Graphene Oxide and Gold Nanocomposites in MCF-7 Breast Cancer Cells. ACS Omega 2021, 6, 15147–15155. [Google Scholar] [CrossRef] [PubMed]
  146. Aksoy, İ.; Küçükkeçeci, H.; Sevgi, F.; Metin, Ö.; Hatay Patir, I. Photothermal Antibacterial and Antibiofilm Activity of Black Phosphorus/Gold Nanocomposites against Pathogenic Bacteria. ACS Appl. Mater. Interfaces 2020, 12, 26822–26831. [Google Scholar] [CrossRef] [PubMed]
  147. Chamundeeswari, M.; Sobhana, S.S.; Jacob, J.P.; Kumar, M.G.; Devi, M.P.; Sastry, T.P.; Mandal, A.B. Preparation, characterization and evaluation of a biopolymeric gold nanocomposite with antimicrobial activity. Biotechnol. Appl. Biochem. 2010, 55, 29–35. [Google Scholar] [CrossRef]
  148. Choodari Gharehpapagh, A.; Farahpour, M.R.; Jafarirad, S. The biological synthesis of gold/perlite nanocomposite using Urtica dioica extract and its chitosan-capped derivative for healing wounds infected with methicillin-resistant Staphylococcus aureus. Int. J. Biol. Macromol. 2021, 183, 447–456. [Google Scholar] [CrossRef]
  149. Jayeoye, T.J.; Eze, F.N.; Singh, S.; Olatunde, O.O.; Benjakul, S.; Rujiralai, T. Synthesis of gold nanoparticles/polyaniline boronic acid/sodium alginate aqueous nanocomposite based on chemical oxidative polymerization for biological applications. Int. J. Biol. Macromol. 2021, 179, 196–205. [Google Scholar] [CrossRef]
  150. Lu, B.; Lu, F.; Ran, L.; Yu, K.; Xiao, Y.; Li, Z.; Dai, F.; Wu, D.; Lan, G. Imidazole-molecule-capped chitosan–gold nanocomposites with enhanced antimicrobial activity for treating biofilm-related infections. J. Colloid Interface Sci. 2018, 531, 269–281. [Google Scholar] [CrossRef]
  151. Mane, P.C.; Chaudhari, R.D.; Shinde, M.D.; Kadam, D.D.; Song, C.K.; Amalnerkar, D.P.; Lee, H. Designing Ecofriendly Bionanocomposite Assembly with Improved Antimicrobial and Potent on-site Zika Virus Vector Larvicidal Activities with its Mode of Action. Sci. Rep. 2017, 7, 15531. [Google Scholar] [CrossRef]
  152. Łyczek, J.; Bończak, B.; Krzymińska, I.; Giżyński, K.; Paczesny, J. Gold–oxoborate nanocomposite-coated orthodontic brackets gain antibacterial properties while remaining safe for eukaryotic cells. J. Biomed. Mater. Res. Part B Appl. Biomater. 2023, 111, 996–1004. [Google Scholar] [CrossRef] [PubMed]
  153. Christau, S.; Möller, T.; Brose, F.; Genzer, J.; Soltwedel, O.; von Klitzing, R. Effect of gold nanoparticle hydrophobicity on thermally induced color change of PNIPAM brush/gold nanoparticle hybrids. Polymer 2016, 98, 454–463. [Google Scholar] [CrossRef]
  154. Tröger, L.; Hünnefeld, H.; Nunes, S.; Oehring, M.; Fritsch, D. Structural characterization of catalytically active metal nanoclusters in poly (amide imide) films with high metal loading. J. Phys. Chem. B 1997, 101, 1279–1291. [Google Scholar] [CrossRef]
  155. Günther-Schade, K.; Castricum, H.; Ziegler, H.; Bakker, H.; Faupel, F. Free volume changes in mechanically milled PS and PC studied by positron annihilation lifetime spectroscopy (PALS). Polym. Eng. Sci. 2004, 44, 1351–1359. [Google Scholar] [CrossRef]
  156. Martinů, L.; Biederman, H.; Zemek, J. Metal doped polymer films prepared by simultaneous plasma polymerization of tetrafluoromethane and evaporation of gold. Vacuum 1985, 35, 171–176. [Google Scholar] [CrossRef]
  157. Hong, J.; Kay, E.; Wang, S.X. Granular magnetic cobalt metal/polymer thin film system. IEEE Trans. Magn. 1996, 32, 4475–4477. [Google Scholar] [CrossRef]
  158. Biederman, H. RF sputtering of polymers and its potential application. Vacuum 2000, 59, 594–599. [Google Scholar] [CrossRef]
  159. Khangarot, R.K.; Khandelwal, M.; Singh, R. Copper-based polymer nanocomposites: Application as sensors. In Metal Nanocomposites for Energy and Environmental Applications; Springer: Berlin/Heidelberg, Germany, 2022; pp. 489–508. [Google Scholar]
  160. Tamayo, L.; Azócar, M.; Kogan, M.; Riveros, A.; Páez, M. Copper-polymer nanocomposites: An excellent and cost-effective biocide for use on antibacterial surfaces. Mater. Sci. Eng. C 2016, 69, 1391–1409. [Google Scholar] [CrossRef] [PubMed]
  161. Solanki, J.N.; Sengupta, R.; Murthy, Z. Synthesis of copper sulphide and copper nanoparticles with microemulsion method. Solid State Sci. 2010, 12, 1560–1566. [Google Scholar] [CrossRef]
  162. Qian, Z.; Guye, K.N.; Masiello, D.J.; Ginger, D.S. Dynamic optical switching of polymer/plasmonic nanoparticle hybrids with sparse loading. J. Phys. Chem. B 2017, 121, 1092–1099. [Google Scholar] [CrossRef]
  163. Dzhardimalieva, G.I.; Uflyand, I.E. Preparation of metal-polymer nanocomposites by chemical reduction of metal ions: Functions of polymer matrices. J. Polym. Res. 2018, 25, 255. [Google Scholar] [CrossRef]
  164. Gün Gök, Z.; Günay, K.; Arslan, M.; Yiğitoğlu, M.; Vargel, İ. Coating of modified poly (ethylene terephthalate) fibers with sericin-capped silver nanoparticles for antimicrobial application. Polym. Bull. 2020, 77, 1649–1665. [Google Scholar] [CrossRef]
  165. Heydari Foroushani, P.; Rahmani, E.; Alemzadeh, I.; Vossoughi, M.; Pourmadadi, M.; Rahdar, A.; Díez-Pascual, A.M. Curcumin sustained release with a hybrid chitosan-silk fibroin nanofiber containing silver nanoparticles as a novel highly efficient antibacterial wound dressing. Nanomaterials 2022, 12, 3426. [Google Scholar] [CrossRef] [PubMed]
  166. Lizundia, E.; Sipponen, M.H.; Greca, L.G.; Balakshin, M.; Tardy, B.L.; Rojas, O.J.; Puglia, D. Multifunctional lignin-based nanocomposites and nanohybrids. Green Chem. 2021, 23, 6698–6760. [Google Scholar] [CrossRef] [PubMed]
  167. Dairi, N.; Ferfera-Harrar, H.; Ramos, M.; Garrigós, M.C. Cellulose acetate/AgNPs-organoclay and/or thymol nano-biocomposite films with combined antimicrobial/antioxidant properties for active food packaging use. Int. J. Biol. Macromol. 2019, 121, 508–523. [Google Scholar] [CrossRef]
  168. Li, M.; Cai, Y.-N.; Peng, C.-F.; Wei, X.-L.; Wang, Z.-P. DNA dendrimer–templated copper nanoparticles: Self-assembly, aggregation-induced emission enhancement and sensing of lead ions. Microchim. Acta 2021, 188, 346. [Google Scholar] [CrossRef]
  169. Najafi, F.; Salami-Kalajahi, M.; Roghani-Mamaqani, H.; Kahaie-Khosrowshahi, A. Effect of grafting ratio of poly (propylene imine) dendrimer onto gold nanoparticles on the properties of colloidal hybrids, their DOX loading and release behavior and cytotoxicity. Colloids Surf. B Biointerfaces 2019, 178, 500–507. [Google Scholar] [CrossRef]
  170. Kayaci, F.; Ozgit-Akgun, C.; Donmez, I.; Biyikli, N.; Uyar, T. Polymer–inorganic core–shell nanofibers by electrospinning and atomic layer deposition: Flexible nylon–ZnO core–shell nanofiber mats and their photocatalytic activity. ACS Appl. Mater. Interfaces 2012, 4, 6185–6194. [Google Scholar] [CrossRef]
  171. Deniz, A.E.; Vural, H.A.; Ortaç, B.; Uyar, T. Gold nanoparticle/polymer nanofibrous composites by laser ablation and electrospinning. Mater. Lett. 2011, 65, 2941–2943. [Google Scholar] [CrossRef]
  172. Qi, W.; Zhang, X.; Wang, H. Self-assembled polymer nanocomposites for biomedical application. Curr. Opin. Colloid Interface Sci. 2018, 35, 36–41. [Google Scholar] [CrossRef]
  173. Zhang, D.; Jiang, C.; Zhou, Q. Layer-by-layer self-assembly of tricobalt tetroxide-polymer nanocomposite toward high-performance humidity-sensing. J. Alloys Compd. 2017, 711, 652–658. [Google Scholar] [CrossRef]
  174. Qi, W.; Xue, Z.; Yuan, W.; Wang, H. Layer-by-layer assembled graphene oxide composite films for enhanced mechanical properties and fibroblast cell affinity. J. Mater. Chem. B 2014, 2, 325–331. [Google Scholar] [CrossRef] [PubMed]
  175. Cui, S.; Yang, L.; Wang, J.; Wang, X. Fabrication of a sensitive gas sensor based on PPy/TiO2 nanocomposites films by layer-by-layer self-assembly and its application in food storage. Sens. Actuators B Chem. 2016, 233, 337–346. [Google Scholar] [CrossRef]
  176. Capadona, J.R.; Van Den Berg, O.; Capadona, L.A.; Schroeter, M.; Rowan, S.J.; Tyler, D.J.; Weder, C. A versatile approach for the processing of polymer nanocomposites with self-assembled nanofibre templates. Nat. Nanotechnol. 2007, 2, 765–769. [Google Scholar] [CrossRef]
  177. Lengert, E.V.; Koltsov, S.I.; Li, J.; Ermakov, A.V.; Parakhonskiy, B.V.; Skorb, E.V.; Skirtach, A.G. Nanoparticles in polyelectrolyte multilayer layer-by-layer (LbL) films and capsules—Key enabling components of hybrid coatings. Coatings 2020, 10, 1131. [Google Scholar] [CrossRef]
  178. Abuid, N.J.; Gattás-Asfura, K.M.; Schofield, E.A.; Stabler, C.L. Layer-by-Layer Cerium Oxide Nanoparticle Coating for Antioxidant Protection of Encapsulated Beta Cells. Adv. Healthc. Mater. 2019, 8, 1801493. [Google Scholar] [CrossRef]
  179. Hu, H.; Pauly, M.; Felix, O.; Decher, G. Spray-assisted alignment of Layer-by-Layer assembled silver nanowires: A general approach for the preparation of highly anisotropic nano-composite films. Nanoscale 2017, 9, 1307–1314. [Google Scholar] [CrossRef] [PubMed]
  180. Meyer, B.; Croce, F. MATERIALS|Nanofibers. In Encyclopedia of Electrochemical Power Source; Elsevier: Amsterdam, The Netherlands, 2009; pp. 607–612. [Google Scholar] [CrossRef]
  181. Marcos-Hernández, M.; Villagrán, D. Mesoporous composite nanomaterials for dye removal and other applications. In Composite Nanoadsorbents; Elsevier: Amsterdam, The Netherlands, 2019; pp. 265–293. [Google Scholar]
  182. Poolakkandy, R.R.; Menamparambath, M.M. Soft-template-assisted synthesis: A promising approach for the fabrication of transition metal oxides. Nanoscale Adv. 2020, 2, 5015–5045. [Google Scholar] [CrossRef]
  183. Idumah, C.I.; Ezeani, E.; Nwuzor, I. A review: Advancements in conductive polymers nanocomposites. Polym. Plast. Technol. Mater. 2021, 60, 756–783. [Google Scholar] [CrossRef]
  184. Khasim, S. Polyaniline-Graphene nanoplatelet composite films with improved conductivity for high performance X-band microwave shielding applications. Results Phys. 2019, 12, 1073–1081. [Google Scholar] [CrossRef]
  185. Sankar, S.; Parvathi, K.; Ramesan, M. Structural characterization, electrical properties and gas sensing applications of polypyrrole/Cu-Al2O3 hybrid nanocomposites. High Perform. Polym. 2020, 32, 719–728. [Google Scholar] [CrossRef]
  186. Nangai, E.K.; Saravanan, S. Synthesis, fabrication and testing of polymer nanocomposites: A review. Mater. Today: Proc. 2021. [Google Scholar] [CrossRef]
  187. Sedighi, F.; Esmaeili-Zare, M.; Sobhani-Nasab, A.; Behpour, M. Synthesis and characterization of CuWO4 nanoparticle and CuWO4/NiO nanocomposite using co-precipitation method; application in photodegradation of organic dye in water. J. Mater. Sci. Mater. Electron. 2018, 29, 13737–13745. [Google Scholar] [CrossRef]
  188. Jazi, F.S.; Parvin, N.; Rabiei, M.; Tahriri, M.; Shabestari, Z.M.; Azadmehr, A.R. Effect of the synthesis route on the grain size and morphology of ZnO/Ag nanocomposite. J. Ceram. Process. Res. 2012, 13, 523–526. [Google Scholar]
  189. Somraksa, W.; Suwanboon, S.; Amornpitoksuk, P.; Randorn, C. Physical and photocatalytic properties of CeO2/ZnO/ZnAl2O4 ternary nanocomposite prepared by Co-precipitation method. Mater. Res. 2020, 23, e20190627. [Google Scholar] [CrossRef]
  190. Eslami, H.; Ehrampoush, M.H.; Esmaeili, A.; Ebrahimi, A.A.; Ghaneian, M.T.; Falahzadeh, H.; Salmani, M.H. Synthesis of mesoporous Fe-Mn bimetal oxide nanocomposite by aeration co-precipitation method: Physicochemical, structural, and optical properties. Mater. Chem. Phys. 2019, 224, 65–72. [Google Scholar] [CrossRef]
  191. Egizbek, K.; Kozlovskiy, A.; Ludzik, K.; Zdorovets, M.; Korolkov, I.; Marciniak, B.; Jazdzewska, M.; Chudoba, D.; Nazarova, A.; Kontek, R. Stability and cytotoxicity study of NiFe2O4 nanocomposites synthesized by co-precipitation and subsequent thermal annealing. Ceram. Int. 2020, 46, 16548–16555. [Google Scholar] [CrossRef]
  192. Parashar, M.; Shukla, V.K.; Singh, R. Metal oxides nanoparticles via sol–gel method: A review on synthesis, characterization and applications. J. Mater. Sci. Mater. Electron. 2020, 31, 3729–3749. [Google Scholar] [CrossRef]
  193. Tseng, T.K.; Lin, Y.S.; Chen, Y.J.; Chu, H. A review of photocatalysts prepared by sol-gel method for VOCs removal. Int. J. Mol. Sci. 2010, 11, 2336–2361. [Google Scholar] [CrossRef]
  194. Ward, D.A.; Ko, E.I. Preparing catalytic materials by the sol-gel method. Ind. Eng. Chem. Res. 1995, 34, 421–433. [Google Scholar] [CrossRef]
  195. Hench, L.L.; West, J.K. The sol-gel process. Chem. Rev. 1990, 90, 33–72. [Google Scholar] [CrossRef]
  196. Taufik, A.; Albert, A.; Saleh, R. Sol-gel synthesis of ternary CuO/TiO2/ZnO nanocomposites for enhanced photocatalytic performance under UV and visible light irradiation. J. Photochem. Photobiol. A Chem. 2017, 344, 149–162. [Google Scholar] [CrossRef]
  197. Ahmed, M. Synthesis and structural features of mesoporous NiO/TiO2 nanocomposites prepared by sol–gel method for photodegradation of methylene blue dye. J. Photochem. Photobiol. A Chem. 2012, 238, 63–70. [Google Scholar] [CrossRef]
  198. Lavin, A.; Sivasamy, R.; Mosquera, E.; Morel, M.J. High proportion ZnO/CuO nanocomposites: Synthesis, structural and optical properties, and their photocatalytic behavior. Surf. Interfaces 2019, 17, 100367. [Google Scholar] [CrossRef]
  199. Amin, S.A.; Pazouki, M.; Hosseinnia, A. Synthesis of TiO2–Ag nanocomposite with sol–gel method and investigation of its antibacterial activity against E. coli. Powder Technol. 2009, 196, 241–245. [Google Scholar] [CrossRef]
  200. Flores-Rojas, G.; López-Saucedo, F.; Bucio, E. Gamma-irradiation applied in the synthesis of metallic and organic nanoparticles: A short review. Radiat. Phys. Chem. 2020, 169, 107962. [Google Scholar] [CrossRef]
  201. Kokulnathan, T.; Wang, T.-J.; Ahmed, F.; Kumar, S. Deep eutectic solvents-assisted synthesis of NiFe-LDH/Mo2C nanocomposite for electrochemical determination of nitrite. J. Mol. Liq. 2023, 369, 120785. [Google Scholar] [CrossRef]
  202. Hashim, A.; Habeeb, M.; Jebur, Q. Structural, Dielectric and Optical properties for (Polyvinyl Alcohol–Polyethylene Oxide-Manganese Oxide) Nanocomposites. Egypt. J. Chem. 2019, 62, 735–749. [Google Scholar] [CrossRef]
  203. Madou, M.J. Manufacturing Techniques for Microfabrication and Nanotechnology; CRC Press: Boca Raton, FL, USA, 2011. [Google Scholar]
  204. Brust, M.; Walker, M.; Bethell, D.; Schiffrin, D.J.; Whyman, R. Synthesis of thiol-derivatised gold nanoparticles in a two-phase liquid–liquid system. J. Chem. Soc. Chem. Commun. 1994, 801–802. [Google Scholar] [CrossRef]
  205. Guisbiers, G.; Abudukelimu, G.; Hourlier, D. Size-dependent catalytic and melting properties of platinum-palladium nanoparticles. Nanoscale Res. Lett. 2011, 6, 1–5. [Google Scholar] [CrossRef] [PubMed]
  206. Vimala, K.; Samba Sivudu, K.; Murali Mohan, Y.; Sreedhar, B.; Mohana Raju, K. Controlled silver nanoparticles synthesis in semi-hydrogel networks of poly(acrylamide) and carbohydrates: A rational methodology for antibacterial application. Carbohydr. Polym. 2009, 75, 463–471. [Google Scholar] [CrossRef]
  207. Liu, H.L.; Dai, S.A.; Fu, K.Y.; Hsu, S.H. Antibacterial properties of silver nanoparticles in three different sizes and their nanocomposites with a new waterborne polyurethane. Int. J. Nanomed. 2010, 5, 1017–1028. [Google Scholar] [CrossRef]
  208. Liu, B.S.; Huang, T.B. Nanocomposites of genipin-crosslinked chitosan/silver nanoparticles--structural reinforcement and antimicrobial properties. Macromol. Biosci. 2008, 8, 932–941. [Google Scholar] [CrossRef]
  209. Fortunati, E.; Latterini, L.; Rinaldi, S.; Kenny, J.M.; Armentano, I. PLGA/Ag nanocomposites: In vitro degradation study and silver ion release. J. Mater. Sci. Mater. Med. 2011, 22, 2735–2744. [Google Scholar] [CrossRef]
  210. Cioffi, N.; Torsi, L.; Ditaranto, N.; Tantillo, G.; Ghibelli, L.; Sabbatini, L.; Bleve-Zacheo, T.; D’Alessio, M.; Zambonin, P.G.; Traversa, E. Copper Nanoparticle/Polymer Composites with Antifungal and Bacteriostatic Properties. Chem. Mater. 2005, 17, 5255–5262. [Google Scholar] [CrossRef]
  211. Shevtsova, T.; Cavallaro, G.; Lazzara, G.; Milioto, S.; Donchak, V.; Harhay, K.; Korolko, S.; Budkowski, A.; Stetsyshyn, Y. Temperature-responsive hybrid nanomaterials based on modified halloysite nanotubes uploaded with silver nanoparticles. Colloids Surf. A: Physicochem. Eng. Asp. 2022, 641, 128525. [Google Scholar] [CrossRef]
  212. Liu, H.F.; Liu, Z.L.; Xie, C.S.; Yu, J.; Zhu, C.H. The antifertility effectiveness of copper/low-density polyethylene nanocomposite and its influence on the endometrial environment in rats. Contraception 2007, 75, 157–161. [Google Scholar] [CrossRef] [PubMed]
  213. Choi, J.; Wang, N.S. Nanoparticles in Biomedical Applications and Their Safety Concerns; InTech Open: London, UK, 2011. [Google Scholar]
  214. Colino, C.I.; Lanao, J.M.; Gutierrez-Millan, C. Recent advances in functionalized nanomaterials for the diagnosis and treatment of bacterial infections. Mater. Sci. Eng. C Mater. Biol. Appl. 2021, 121, 111843. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Percentage of antibacterial studies of metal nanoparticles. Source: Scopus Database. Keywords: metal (silver, zinc, copper, gold, iron, and other metals), nanoparticles, antibacterial.
Figure 1. Percentage of antibacterial studies of metal nanoparticles. Source: Scopus Database. Keywords: metal (silver, zinc, copper, gold, iron, and other metals), nanoparticles, antibacterial.
Polymers 15 02167 g001
Figure 2. Mechanisms of action of silver, gold, zinc, and copper nanoparticles as antibacterial agents.
Figure 2. Mechanisms of action of silver, gold, zinc, and copper nanoparticles as antibacterial agents.
Polymers 15 02167 g002
Figure 3. Different synthetic approaches to produce metal–polymer nanocomposites.
Figure 3. Different synthetic approaches to produce metal–polymer nanocomposites.
Polymers 15 02167 g003
Figure 4. In situ polymerization techniques for the synthesis of nanocomposites.
Figure 4. In situ polymerization techniques for the synthesis of nanocomposites.
Polymers 15 02167 g004
Figure 5. Layer-by-layer film deposition method for the synthesis of nanocomposite with inorganic species.
Figure 5. Layer-by-layer film deposition method for the synthesis of nanocomposite with inorganic species.
Polymers 15 02167 g005
Figure 6. Hard template and soft approach for the synthesis of nanoscale architecture in the preparation of various polymeric nanocomposites containing inorganic nanoparticles.
Figure 6. Hard template and soft approach for the synthesis of nanoscale architecture in the preparation of various polymeric nanocomposites containing inorganic nanoparticles.
Polymers 15 02167 g006
Figure 7. Gamma-rays-catalyzed synthesis of nanoparticles and crosslinked polymers, nanocomposite with nanoparticles.
Figure 7. Gamma-rays-catalyzed synthesis of nanoparticles and crosslinked polymers, nanocomposite with nanoparticles.
Polymers 15 02167 g007
Table 1. Silver–polymer nanocomposites and their biomedical applications.
Table 1. Silver–polymer nanocomposites and their biomedical applications.
No.Polymer NameSourcePropertiesUsesRef.
1Polyethylene glycol PEGSynthetic5 to 20 nmAntimicrobial[40]
2ChitosanSynthetic20.3 ± 0.7 nm and 44.6 ± 0.3 nmAntifungal[41]
3Modified polypropyleneSynthetic26 and 41Antibacterial[42]
4Poly-acrylonitrile (PAN)Synthetic30–95 nmAntimicrobial activity, electrical conductivity, ultraviolet blocking, and catalytic activity[43]
5Poly (vinyl alcohol) (PVA)Synthetic40.49–44.77Antimicrobial[44]
6Cellulose nanowhiskersCotton fibers6 nm and 18 nmAntibacterial[45]
7Chitosan hydrogelsSynthetic4.45–9.22 nmAntibacterial[46]
8ChitosanSynthetic190 nmAntioxidant, antibacterial, hemolytic, and cutaneous wound healing.[47]
9Gelatin chitosanSynthetic3–6 nmAntimicrobial and wound healing[48]
10ChitosanSynthetic35 nmAntibacterial and wound healing[49]
11Poly (methyl methacrylate) with chitosanSynthetic37.2 ± 16.4 nmAntibacterial[50]
12ChitosanSynthetic5.0–9.9 nmDental barrier membrane applications[51]
13Tragacanth/hydroxy-propyl methyl cellulose/beeswax filmsNatural8–10 nmAntimicrobial, edible films[52]
14Polyacrylonitrile nanofibersSynthetic7–20 nm (AgNP); 450–700 nm (nanofiber)Antibacterial, antioxidant[53]
15Reduced graphene oxideSynthetic<10 nmAntibacterial[54]
16Chitosan-based carbon nitride-polydopamine silver compositeSynthetic-Wound healing antibacterial[55]
17Hemicellulose, chitosan/chitin, and glutaraldehydeSynthetic-Wound healing dressings[56]
18Polyurethane and cellulose acetateSynthetic-Wound healing[57]
19Arabinoxylan-co-acrylic acid, nanohydroxyapatite (nHAp), graphene oxideSynthetic20–100 nmBone tissue engineering, antibacterial, biocompatible[58]
20Carboxymethyl chitosan-polyamideamine alginate compositeSynthetic158.0 ± 2.3 nmAntibacterial, wound healing[59]
21P4VP, P (4VP-co-OEGMA246), and POEGMA246Synthetic20–60 nmOptical application[60]
Table 2. Copper–polymer nanocomposites and their biomedical applications.
Table 2. Copper–polymer nanocomposites and their biomedical applications.
No.Polymer NameSourcePropertiesUsesRef.
1ChitosanSynthetic160 nmAntimicrobial, photocatalyst[66]
2Chitosan: Pluronic F127Synthetic~8 ± 2 nmAntimicrobial[67]
3StarchNatural200 nmAntimicrobial, antioxidant, and anticancer[68]
4PolyacrylonitrileSynthetic100–200 nm long, 80 nm wideAntibacterial[69]
5Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV)Synthetic182.65 nmAntibacterial[70]
6PolyurethaneSynthetic47.51 nmAntibacterial[71]
7PolystyreneSynthetic18–25 nmAntimicrobial[72]
8Bacterial celluloseNatural25 and 35 nmAntimicrobial[73]
9Cotton fabrics modified with polycarboxylic acidsSynthetic Antibacterial[74]
10MontmorilloniteSynthetic15.29 nmAntibacterial[75]
11Reduced graphene oxideSynthetic5–40 nmAntibacterial[76]
12ChitosanNatural17 nmAntimicrobial, sporicidal, and biofilm-inhibitory activity[77]
13Polyethylene oxideSynthetic12 ± 6 nmPreservation system for fruit salads[78]
14ChitinSynthetic52.1 ± 20 nmAnticancer[79]
15ChitosanSynthetic11 ± 6 nmAntibacterial[80]
16CelluloseNatural21–30 nm and 31–40 nmAntibacterial[81]
17Carboxymethyl starchSynthetic30–50 nmAntimicrobial, wound healing[82]
18ChitosanNatural40–110Antimicrobial[83]
19NanochitosanNatural18 to 40 nmAntimicrobial[84]
20Cellulose gumNatural7 to 12 nmAntimicrobial against UTI[85]
21Cellulose acetateSynthetic143.2 to 157.1 nm,Antimicrobial, cell viability[86]
22Hypromellose polymerSynthetic50–60 nm, nanofibersAntibacterial[87]
23poly(vinyl alcohol)Synthetic300 nmAntimicrobial-air microfilter[88]
24Bacterial celluloseNatural10–100 nmAntibacterial[89]
25Hydroxypropyl methyl cellulose hydrogelSynthetic3–17 nmAntibacterial[90]
26Chitosan-capped NPNatural5–9 nmMicrobial-resistant nanocomposites[91]
27Poly(diallyldimethylammonium chloride) (PDADMAC), Poly(sodium 4-styrenesulfonate)(PSS)Synthetic50 and 70 nmAntimicrobial coating[92]
283-(Trimethoxysilyl)propyl methacrylate, mesoporous silica gelSynthetic~10 nmBactericidal, wound healing[93]
29Poly(piperizinamide)/copper composite membraneSynthetic1 nm pore sizeMitigation of hexaconazole from water and combat microbial contamination[94]
30Graphene oxide–chitosan compositeSynthetic2.4 to 257.6 nmAntibacterial and cytotoxic activities[95]
31Chitosan nanocompositeNaturalPolygonal 30–50 nmAntibacterial and electrical properties[96]
32Polyvinyl alcoholSynthetic1–100 nmAntibacterial[97]
33ChitosanBio-synthetic20–100 nmAntibacterial activity[98]
34Chitosan nanocompositeNatural, nanocomposite using Sida cordifolia extract~Antibacterial, anticancer activity against on breast and lung cancer cell lines[99]
35Multifunctional chitosan–gallic-acid-based nanocompositeNatural30 nmAntibacterial, wound dressing[100]
36PolyacryonitrileSyntheticNanofibers, nanoparticles, 100 nmAntibacterial against resistant strains, MRSA[101]
37Nanoderivatives-modified chitosan/hyaluronic acidSynthetic5–10nmWound healing[102]
Table 4. Gold–polymer nanocomposites and their biomedical applications.
Table 4. Gold–polymer nanocomposites and their biomedical applications.
No.Polymer nameSourcePropertiesUsesRef.
1Chitosan/aminopropylsilaneSynthetic3.99 nmAntimicrobial[140]
2ChitosanSynthetic5–100 nmAntimicrobial and wound-healing activities[141]
3Imidazole-molecule-capped chitosanSynthetic12.16 nm, 10.87 nm, 9.56 nm, and 8.56 nmAntimicrobial[70]
4ChitosanNatural16.9 ± 2.0, 25.0 ± 4.0 and 34.1 ± 5.9 nmAntibacterial[142]
5ChitosanSynthetic14 nmAntibacterial[14]
6Carboxymethyl celluloseSynthetic10–90 nmAntimicrobial and anticancer[143]
7Heparin-polyvinyl alcoholSyntheticless than 100 nmWound healing[144]
8Highly reduced graphene oxideSynthetic3.27 ± 0.02 nmAntiproliferative[145]
9Black phosphorus/gold nanocompositesSyntheticNanoscaleAntibacterial activity due to photothermal effect, oxidative stress, and physical membrane damage[146]
10Chitosan-capped gold nanoparticles coupled with ampicillinNaturalellipsoidal particles, 50–100 nmAntimicrobial activity[147]
11Chitosan, (gold)/perlite (Au/Perl) nanocompositesNaturalMesoporous/spherical shape, 13–15 nmAntibacterial, wound healing against MRSA[148]
12Sodium alginate nanocomposite with gold nanoparticles/polyaniline boronic acidSynthetic15–20 nmAntibacterial, biocompatible[149]
13Chitosan,
imidazole-molecule-capped chitosan
Synthetic8.5–12.5 nmAntimicrobial activity[150]
14Dialyzed natural polymer, fibroin, (gold) nanocompositeNatural20–30 nmZika virus vector larvicidal activities[151]
15Polyoxoborate matrix/nanocomposite of gold nanoparticlesSyntheticNanoscaledAntibacterial[152]
16Poly(N- isopropylacrylamide) (PNIPAM) brushesSyntheticNanoscaledNanosensor[153]
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

Ghazzy, A.; Naik, R.R.; Shakya, A.K. Metal–Polymer Nanocomposites: A Promising Approach to Antibacterial Materials. Polymers 2023, 15, 2167. https://doi.org/10.3390/polym15092167

AMA Style

Ghazzy A, Naik RR, Shakya AK. Metal–Polymer Nanocomposites: A Promising Approach to Antibacterial Materials. Polymers. 2023; 15(9):2167. https://doi.org/10.3390/polym15092167

Chicago/Turabian Style

Ghazzy, Asma, Rajashri R. Naik, and Ashok K. Shakya. 2023. "Metal–Polymer Nanocomposites: A Promising Approach to Antibacterial Materials" Polymers 15, no. 9: 2167. https://doi.org/10.3390/polym15092167

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