Next Article in Journal
Self-Medication with Antibiotics: Prevalence, Practices and Related Factors among the Pakistani Public
Next Article in Special Issue
Ultra-Small Silver Nanoparticles: A Sustainable Green Synthesis Approach for Antibacterial Activity
Previous Article in Journal
The Role of Gut Microbiota in the Skeletal Muscle Development and Fat Deposition in Pigs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nanomaterials-Based Combinatorial Therapy as a Strategy to Combat Antibiotic Resistance

by
Angel León-Buitimea
1,2,
Cesar R. Garza-Cárdenas
1,2,
María Fernanda Román-García
3,
César Agustín Ramírez-Díaz
4,
Martha Ulloa-Ramírez
5 and
José Rubén Morones-Ramírez
1,2,*
1
Facultad de Ciencias Químicas, Universidad Autónoma de Nuevo León (UANL), San Nicolás de los Garza 66455, Mexico
2
Centro de Investigación en Biotecnología y Nanotecnología, Facultad de Ciencias Químicas, Universidad Autónoma de Nuevo León, Parque de Investigación e Innovación Tecnológica, Apodaca 66628, Mexico
3
Centro de Estudios Científicos y Tecnológicos No. 9 “Juan de Dios Bátiz”, Instituto Politécnico Nacional, Miguel Hidalgo, Ciudad de México 11400, Mexico
4
Facultad de Ciencias Biológicas, Benemérita Universidad Autónoma de Puebla (BUAP), 4 Sur 104 Centro Histórico, Puebla 72000, Mexico
5
Centro Universitario de Tonalá, Universidad de Guadalajara (UDG), Guadalajara 45425, Mexico
*
Author to whom correspondence should be addressed.
Antibiotics 2022, 11(6), 794; https://doi.org/10.3390/antibiotics11060794
Submission received: 27 April 2022 / Revised: 3 June 2022 / Accepted: 7 June 2022 / Published: 12 June 2022
(This article belongs to the Special Issue Nanocarriers for Antibacterial Delivery)

Abstract

:
Since the discovery of antibiotics, humanity has been able to cope with the battle against bacterial infections. However, the inappropriate use of antibiotics, the lack of innovation in therapeutic agents, and other factors have allowed the emergence of new bacterial strains resistant to multiple antibiotic treatments, causing a crisis in the health sector. Furthermore, the World Health Organization has listed a series of pathogens (ESKAPE group) that have acquired new and varied resistance to different antibiotics families. Therefore, the scientific community has prioritized designing and developing novel treatments to combat these ESKAPE pathogens and other emergent multidrug-resistant bacteria. One of the solutions is the use of combinatorial therapies. Combinatorial therapies seek to enhance the effects of individual treatments at lower doses, bringing the advantage of being, in most cases, much less harmful to patients. Among the new developments in combinatorial therapies, nanomaterials have gained significant interest. Some of the most promising nanotherapeutics include polymers, inorganic nanoparticles, and antimicrobial peptides due to their bactericidal and nanocarrier properties. Therefore, this review focuses on discussing the state-of-the-art of the most significant advances and concludes with a perspective on the future developments of nanotherapeutic combinatorial treatments that target bacterial infections.

1. Introduction

Since the discovery of penicillin in 1928, researchers started a promising era of antibiotic discovery and development against human pathogens; this period was referred to as the golden age of antibiotics [1,2]. During that period, medicinal chemistry was a successful strategy to produce semi-synthetic antibiotics and improve their physicochemical and pharmacokinetic properties [3]. Unfortunately, the irrational use of antibiotics, inappropriate prescriptions, extensive application in agriculture, the lack of innovation, and obstacles in regulatory approvals, have favored the emergence of antibiotic-resistant bacteria [4,5]. The resistance mechanisms developed have favored the spread of multidrug-resistant (MDR), extensively drug resistant (XDR), and pan-drug resistant (PDR) bacteria, being the latter, the most dangerous because they possess resistance to most kinds of antibiotics [6]. The most common resistance mechanisms are the enzymatic modification of the antibiotic, alterations in membrane permeability, such as efflux transporters and porins, and modification of the binding site [1,7,8]. Bacteria produce enzymes that irreparably modify antibiotics; this mechanism (enzymatic modification) is one of the most important to combat. Some of these enzymes are β-lactamases, nucleotidyltransferases, and acetyltransferases [7,8]. Another mechanism bacteria use is to reduce the internal accumulation of drugs through membrane alterations, such as the development of porins or efflux pumps. Porins are proteins found in the membrane that act as channels that allow the passive diffusion of hydrophilic molecules in Gram-negative bacteria. A modification in their expression can confer resistance to different antibiotics [7,9,10]. Efflux pumps push the antibiotic out of the cell at a high rate, which leads to conditions where drug concentrations are never high enough to cause an antibacterial effect [11]. Finally, some bacteria can develop antibiotic resistance by changing the antibiotic target site [9,10]. One of the most important and well-known examples is the methicillin resistance present in some strains of S. aureus. This resistance is given by the mecA gene, which consists of creating low-affinity binding proteins to penicillins (PBP 2a). These modified proteins prevent the main antibiotic effect of methicillin, which is to inhibit the synthesis of peptidoglycans for the cell wall; moreover, it confers immunity to other families of antibiotics such as penicillins, cephalosporins, carbapenems, and other β-lactamics [9,12]. In 2017, the World Health Organization (WHO) published a list of high-priority pathogens highlighting Enterococcus faecium, Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii, Pseudomonas aeruginosa, and Enterobacter spp., (ESKAPE). Therefore, it is urgent to find new therapies against these bacteria since they have evolved to escape the effects of different treatments and transfer their resistance to other organisms [1,9,13]. The need to treat infections, especially those caused by multidrug-resistant bacteria, implies new antibiotics development [5]. However, bacteria invariably possess the ability to evolve and develop resistance to antibiotics soon after these treatments are introduced to the clinic [14].
Various nanomaterials such as polymers (natural and synthetic), inorganic nanoparticles, and antimicrobial peptides offer an innovative platform to develop new treatment strategies [15]. In addition to their scalability, low costs, and versatility, these nanomaterials offer advantages such as lower toxicity due to lower doses, lower resistance development, and an increased antibacterial effect due to the conjunction of individual mechanisms of action [16]. One of the strategies to prevent or delay the emergence of resistance is the design of combinatorial treatments, which in many cases, have been proven to increase the antibacterial effect of the individual therapeutics [13]. Moreover, combining two or more antimicrobial agents seems to be a reasonable alternative since they may act either by inhibiting multiple targets in different pathways, inhibiting other targets in the same pathway, or inhibiting the same target in different modalities [17].
Therefore, this review summarizes, discusses, and concludes on the use and applications of nanomaterial-based combinatorial therapies as novel therapeutic strategies to combat infections caused by antibiotic-resistant bacteria.

2. New Therapeutic Strategies: Combinatorial Treatments

The discovery and preclinical development of novel antibacterial agents must involve new approaches that allow an effective and sustainable combat against antibacterial resistance [18]. There is a strong trend toward the rational design of drug combinations using direct targeting of small molecules and finding new targets or new mechanisms of antimicrobial action [19]. Different antimicrobial agents, such as polymers (synthetic and natural) [20], inorganic nanoparticles [21], and antimicrobial peptides [22], are being developed to prevent, reduce or reverse antimicrobial resistance. Therefore, we provide insights into how polymers, metal nanoparticles, and antimicrobial peptides can be applied with other antimicrobial agents as combinatorial therapies to combat antibiotic-resistant bacteria (Figure 1).

2.1. General Mechanisms of Antibiotic Action of Nanomaterials

Combinatorial treatments, a combination of two or more therapeutic agents, are a potential strategy to address antibiotic resistance. The application of this strategy results in a reduction in the drug dose, lower toxicity, and a decrease in the development of bacterial resistance [23]. Polymers, inorganic nanoparticles, and antimicrobial peptides, which have well-known antimicrobial properties, are used as therapeutic agents to treat bacterial infectious diseases [20,24,25,26]. Various pathogenic bacteria have been treated with combinatorial treatments. The combined effects (synergy, potentiation, or additivity) of these nanomaterials, either among themselves or with conventional antibiotics, have been reported by multiple authors [27]. In many cases, the increased antibacterial effect is due to the combination of the individual drug effects [28].
For example, AMPy enhances the therapeutic effectiveness of antibiotics due to their well-defined globular structure and their ability to attach drugs onto their surface. They also reduce β-lactamase enzyme activity and disrupt cell walls. AMPy can complex antibiotics (bioconjugates) by forming a stable ion-pairing. Thus, they promote damage to the cell walls allowing the release of complexed antibiotics [29]. Moreover, AMPy can serve as a polymeric drug delivery system. They enhance the safety and efficacy of the other component by regulating the rate, time, and place of release in the body [30]. Finally, AMPy is utilized as a capping agent capable of inducing subtle changes in the nanomaterial, enhancing its therapeutic effects [31].
Regarding the inorganic nanoparticles, there are several possible modes of action on the bacteria (Figure 2): (a) alteration of the bacterial cell wall and membrane, (b) induction of oxidative stress due to excessive intracellular production of reactive oxygen species (ROS), (c) inhibition of crucial proteins/enzymes and DNA, (d) disruption of metabolic pathways, and (e) intracellular accumulation of metal ions released from INP [32]. Therefore, these nanomaterials have been proved to be most effective against different strains of pathogenic bacteria.
Antimicrobial peptides are cationic (positively charged) and amphiphilic (hydrophilic and hydrophobic) molecules, and these properties are related to their ability to interact with the bacterial cell membranes [33,34]. In general, AMPs may act through two main mechanisms; in the first one, the AMPs induce membrane disruption and cell lysis, causing bacterial death. The second one is when the AMPs enter cells (without membrane disruption) and interfere with an intracellular pathway or bind to nucleic acids. Thus, AMPs are potential candidates for developing novel antimicrobial therapeutics.

2.2. Nanomaterials-Based Combinatorial Treatments

2.2.1. Polymers

Thanks to advances in chemistry research, there is the possibility of designing and synthesizing antimicrobial polymers (AMPy), materials capable of inhibiting or killing bacteria [20]. AMPy can display antibacterial activities through their chemical structures, such as chitosan compounds with quaternary nitrogen groups, halamines, and poly-ε-lysine (ε-PL).
In this review, we summarize several contributions of combinatorial treatments that include synthetic and natural polymers as the main component of effective antimicrobial agents and their effectivity on bacteria. Table 1 shows the combinatorial treatments that incorporate synthetic and natural polymers as the main component of effective antimicrobial agents.

Synthetic

Although several investigations have demonstrated the antimicrobial effect of individual synthetic polymers, only a few reports include their combination with antibiotics or other molecules as combinatorial treatments [35]. Some of the most critical literature include:

Poly (Lactide-Co-Glycolide)

Poly (lactic-co-glycolic acid) (PLGA) is a synthetic functional polymer commonly used in the biomedical field due to its advantages, including biodegradability, biocompatibility, and non-toxicity [36]. Besides being one of the polymers approved by the FDA for use in humans, PLGA is widely used in drug delivery systems such as microspheres and nanoparticles, proven to be very efficient [37]. Several studies have shown that grafting an AMP (magainin II) by covalent immobilization to an electrospinning PLGA membrane reduced the number of adhered bacteria (E. coli and S. aureus) by more than 50% [38]. Another strategy that has sparked the interest of several authors is the incorporation of silver nanoparticles (AgNPs) into polymeric matrices. Authors showed that the integration of AgNPs within the PLGA matrix was carried out by creating nanofiber scaffolds, which inhibited bacterial growth (P. aeruginosa, K. pneumoniae, S. saprophyticus, and E. coli) depending on the concentration of AgNPs incorporated within the scaffold [39]. Furthermore, Zhu et al. developed an anti-biofilm system, drug-free, using cationic nanoparticles (CNPs). The bioactivity of the CNPs against Streptococcus mutans was examined, and the results showed a concentration-dependent activity against bacteria, an excellent inhibition in biofilm formation, and an interruption of mature biofilms [40].
Unfortunately, some studies showed cytotoxic effects on several cell lines due to the presence of AgNPs in polymeric matrices [41]. For example, Mohiti-Asli et al. showed that human epidermal keratinocytes and human dermal fibroblasts’ viability were decreased by an increase in silver concentration within the coating solution [42]. Moreover, some authors have explored the coupling of antibiotics to polymeric matrices to improve drug delivery [43]; however, not all combinations were successful. For example, roxithromycin (ROX) delivery was enhanced by anchoring to cyclodextrins (ROX-CD) and encapsulating them in PLGA NPs (ROX-CD/PLGA). Results showed that ROX/PLGA NPs dual combination was more potent in inhibiting the growth of selected MDR bacterial strains than the ROX-CD/PLG triple combination. This difference was basically due to the formation of the strong electrostatic interactions among the individual components, which interfere with the release of the drug from these formulations [44].
One strategy to incorporate antimicrobial properties into biomaterials such as PLGA is described by Qian et al. They showed that the PLGA/PCL (polycaprolactone) electrospun scaffold coated with collagen, and modified with silver, had enhanced biocompatibility, osteogenic, and antibacterial properties (against S. aureus and Streptococcus mutans) [45]. Another exciting strategy involved synthesizing organic-metallic hybrid systems, which contain AgNPs integrated into biodegradable polymer nanofibers such as PLGA [20]. These systems coupled benefits from both individual components, including a greater reaction surface area and higher permeability from PLGA with the antimicrobial properties of AgNPs [46]. Similarly, TiO2 nanoparticles have also been incorporated into polymeric systems. PLGA is one of the main polymers used in different techniques, and its antibacterial properties against a variety of microorganisms have been detected [47]. These studies evaluated the viability of the biofilms composed of PLGA-TiO2-NP under ultraviolet light irradiation, showing that biofilms for artificial deposit applications that contained one-tenth of TiO2-NP were effective against E. coli and S. aureus [47]. Despite the various applications of PLGA-incorporated TiO2-NPs, it has been shown that the pure PLGA solution, subjected to a moderate solvent removal process, can generate a tightly arranged thin film with fewer defects than when it is mixed with TiO2-NPs. The structure of pure PLGA is disrupted, leading to structural defects and increased permeability of the biofilm. Moreover, a composite biofilm with a high concentration of TiO2 had a correspondingly higher number of structural defects and is more fragile than composite biofilms with a low concentration of TiO2.
Table 1. Combinatorial treatments that include synthetic and natural polymers as the main component of effective antimicrobial agents.
Table 1. Combinatorial treatments that include synthetic and natural polymers as the main component of effective antimicrobial agents.
NanomaterialCombined with (Rate/Ratio)FormSizeTargeted BacteriaAntimicrobial EffectsReferences
Synthetic
Poly (lactide-co-glycolide)Magainin II (0.2 ± 0.05 μg/cm2)PLGA nanofibers
Magainin II covalently immobilized
PLGA nanofibers diameter 715 ± 45 nmEscherichia coli
Staphylococcus aureus
Reducing the number of adhered bacteria[38]
AgNPs (3% wt/v)Nanofibers of PLGA
AgNPs within the scaffold
Nanofiber diameters between 487 and 781 nm
AgNPs diameter < 100 nm
P. aeruginosa, K. pneumoniae, S. saprophyticus, and E. coliInhibition of bacterial growth[39]
Poloxamer 188
(0.1% w/v)
Nanospheres of PLGA
Poloxamer 188 coating
Nanospheres diameter 217.7 nmStreptococcus mutansInhibition of planktonic bacterial growth and biofilm formation, and disrupted ∼70% mature biofilm[40]
Polycaprolactone (PCL, nd)
Type I collagen (2% w/v)
AgNPs (nd)
Nanofibers of PLGA/PCL
AgNPs reduced in situ with nanofibers
Collagen coating
Nanofibers diameter 477 ± 186 nm,S. aureus and Streptococcus mutansAntibacterial properties[45]
TiO2 NPs (10% w/w)TiO2/PLGA composite biofilmsTiO2 NPs diameter 20 nm E. coli and S. aureusAntibacterial properties[47]
Poly (glycolic acid)ε-caprolactone (14%) trimethylene carbonate (14%)
Oxygen plasma treated
Monofilament suturendE. coli K12Antibacterial properties[48]
N-halamines polymersPGA sutures
N-halamines coating via layer-by-layer
ndS. aureus and E. coliEffective bactericide properties[49]
PLGA (30:70 PGA/PLGA)
AgNPs (3% wt)
PGA: PLGA fibers
AgNPs within the scaffold
Nanofibers diameter 1170 ± 166.98 nm
AgNPs diameter 22 nm
S. aureus and E. coliAntibacterial activity[50]
PLGA (50:50 PLGA/PGA)
AuAg core/shell NPs (600 mg/kg of stent)
PGA/PLGA ureteral stent
AuAg core/shell nanospheres
Au core diameter 10.94 nm
Ag shell thickness 6.98 nm
E. coli and S. aureusLong-lasting inhibitory activity and remarkable antibiofilm properties.[51]
Propylene fumarate (PPF, co-polymer)
Graphene oxide (GO, 5% wt)
Hydroxyapatite (HA, 20% wt).
PGA/PPF nanofibers
HA nanorods and GO within the scaffold
Nanofibers diameter 469 nm
HA nanorods diameter 18 nm and length 50–80 nm
S. aureus and E. coliExtensive biocidal activity.[52]
Natural
Chitosan Multiwalled carbon nanotubes (5 × 10−3% wt)Chitosan/MWCNT
biocomposites
-ESKAPE group bacteriaImproved antimicrobial activity[53]
Chlorhexidine (3% v/v)Chitosan nanoparticles
Chlorhexidine functionalization
Chitosan nanoparticles diameter 70.6 ± 14.8 nmEnterococcus faecalisImproved antibacterial activity[54]
Nisin (0.625 g/L)
Tea polyphenols (0.313 g/L)
Chitosan, Nisin and Tea polyphenols in dissolution -Gram-negative and Gram-positive bacteriaImproved antimicrobial activity[55]
Zinc-EDTA chelateChitosan solution
Zinc-EDTA chelate solution
-Penicillium italicumBetter inhibitory activity[56]
InulinModified (amphiphilic amino inulin)Chemical modification of inulinAmphiphilic amino inulin in solution S. aureusAntibacterial activity[57]
Chitosan (1% w/v)Inulin was glycated to chitosan in solution-Staphylococcus aureus ATCC 25923, Escherichia coli ATCC 25922, Pseudomonas aeruginosa ATCC 27853, Bacillus subtilis ATCC 23857, Candida albicans PTCC 5027 and Aspergillus niger ATCC 23857Significant antimicrobial activity[58]
Chitosan (nd)Covalent conjugation of inulin to chitosan in solution -S. aureusSignificantly improved [59]
Polyvinyl alcohol (PVA) (15% w/v)Composite nanofibers of crosslinked Inulin and PVANanofiber diameter widely dispersedE. coli and S. aureusIncreased antibacterial activity[60]
Carboxymethylcellulose (CMC films)
Celullose nanofiber (CNF, 2.5%)
L. plantarum (109 CFU/mL)
CMC films incorporated with inulin, CNF
L. plantarum inoculated on the film
CMC films
CNF diameter
35 nm, length 5 μm
S. aureus, E. coli, and K. pneumoniaeAntibacterial activity[61]
AlginateZnO NPs (nd)
Cellulose fibers
Cellulose cotton fibers impregnated with sodium alginate-ZnO NPs
ZnO NPs “rod-shape”
ZnO NPs diameter 25 ± 5 nmE. coliSignificant antibacterial activity[62]
Copper (Cu, ~100 µmol/g of microbed)Cu-alginate spherical microbedsCu-alginate microbeds diameter ~550 μmE. coli and S. aureusBactericidal effects[63]
Hydroxyapatite nanoparticles (HA NPs, 5% w/w)Alginate-HA NPs nanocomposite film
-Spherical HA-NPs
Alginate-HA NPs film thickness 0.036 ±0.002 mm
HA NPs diameter 25 ± 2 mm
Listeria monocytogenesShowed the highest antibacterial effect[64]
AgNPs (nd)Alginate-AgNPs solution
Spherical AgNPs
AgNPs diameter < 50 nmS. aureus and E. coliIncreased membrane permeability and disruption of the bacterial wall[65]
Graphene oxide (GO, 1% w/w alginate)
Zinc (Zn, 12% w/w alginate)
Alginate-GO cross-linked films
Zn covering
-S. aureusHigh antibacterial activity[66]
Hydroxypropyl methylcellulose (HPMC, 1% w/w)
ε-polylysine (ε-PL, 1% w/w)
Alginate-HPMC-ε-PL filmsFilm thickness 18 ± 6 µmE. coli and S. aureus99.9% bacterial reduction[67]
Corona treated Polypropylene (CPP, nd)
Copper oxide nanoparticles (CuO NPs, nd)
CPP-alginate fiber nanocomposite
CuO NPs reduced in matrix
CuO NPs diameter 43 ± 15 nmE. coli, S. aureus, and Candida albicansExcellent antimicrobial activity[68]
nd: not described.
Poly (Glycolic Acid)
Poly (glycolic acid) (PGA) is a biopolymer used for biomedical applications due to its biodegradability and thermal and mechanical properties [69]. Another notable feature of PGA is its easy de-esterification to monomer units, leading to an enhanced metabolization and, therefore, a faster degradation rate [70].
Chemical modifications perpetrated on the surface of a polymeric material can significantly control its antimicrobial properties. An example can be drawn from the nano structuration induced by oxygen plasma treatment on modified PGA (PGA (72%), ε-caprolactone (14%), and trimethylene carbonate (14%) absorbable monofilament sutures; this treatment led to materials with antimicrobial properties against E. coli K12 [48]. Likewise, an N-halamine-modified PGA multifilament obtained by the layer-by-layer technique produced effective sutures that successfully inactivated S. aureus and E. coli strains within the first 30 min of contact [49]. In addition, PGA-PLGA electrospun nanofiber sutures added with 3% AgNPs have shown antibacterial activity against Gram-positive (S. aureus) and Gram-negative (E. coli) bacteria. On the other hand, the results of the in vitro degradation test showed that the mechanical properties of these sutures decreased rapidly due to the presence of AgNPs. [50]. Moreover, PGA-PLGA membranes with embedded gold and silver nanoparticles have shown remarkable antibiofilm properties and long-lasting inhibitory properties of 99%, with removal times of between 5 and 10 min for E. coli and S. aureus [51]. Lately, electrospun fibers of a novel biodegradable PGA and propylene fumarate (PGA-co-PPF) copolymer with graphene oxide (GO) and hydroxyapatite nanorods (HA) have been reported to display extensive biocidal activity against Gram-positive (S. aureus) and Gram-negative (E. coli) bacteria [52].
Natural
Also known as biopolymers, natural polymers are materials biologically derived from living organisms and are primarily studied due to their biodegradable, bioactive, or in some cases, antimicrobial properties [71,72]. An essential characteristic of natural polymers is that they can be modified to design their properties, turning them into semi-synthetic polymers with active functional groups [71,73]. Additionally, they are affordable, environmentally friendly, and less expensive than synthetic polymers [73].
Chitosan
Chitosan, derived from the alkaline deacetylation of chitin, is a linear high molecular weight compound consisting of two monosaccharides, N-acetyl-D-glucosamine and D-glucosamine, linked by glucosidic β-1-4 bonds [74]. Among the most important biological properties, its biodegradability through hydrophilic enzymes stands out, in addition to its bactericidal activity [75]. Chitosan’s antimicrobial activity is attributed to the presence of amino groups in its composition, allowing it to interrupt the normal functions of the bacterial membrane, causing both the leakage of intracellular components and the breakdown of nutrient transport to the cell [76]. However, a disadvantage of chitosan is that its antibacterial activity occurs at pH < 6, so it must be synthetically modified [77] to study it under neutral and physiological conditions.
The combination of chitosan derivatives with other compounds has exhibited significant bactericidal activity. For example, chitosan-based biocomposites with multi-walled carbon nanotubes (MWCNT) are effective antimicrobial agents against some bacteria from the ESKAPE group [53]. Furthermore, the antibacterial activity of chitosan against Enterococcus faecalis has been significantly improved by combining it with chlorhexidine in endodontic sealants [54]. Moreover, two effective combinatorial strategies against resistant bacteria have been studied; one determined that the optimal combination of nisin, tea polyphenols (TP), and chitosan (0.625, 0.313, and 3.752 g/L, respectively) has inhibitory properties against Gram-negative and Gram-positive bacteria [55]; the second, demonstrated the synergy of chitosan hydrolysates combined with divalent metal ion-EDTA compounds. Finally, active molecular chitosan (AMC) was found to have better inhibitory activity against Penicillium italicum in the presence of the zinc-EDTA chelate compared to other combinations [56].
Inulin
Inulin is a natural polysaccharide and oligomer widely used in the food industry and generally composed of between 12 and 15 units of fructose [78]. Besides its fructose units, it usually contains a reducing end of glucopyranose units (GFn). Inulin is mainly extracted from low-requirement cultures such as Helianthus tuberosus, chicory, and yacon [79]. Chemical modifications are often introduced to enhance the properties of inulin before use. For example, the “click chemistry” method has been used to modify inulin at its primary hydroxyl groups, leading to the synthesis of amphiphilic amino inulin. Amphiphilic amino inulin has exhibited antibacterial activity against S. aureus, with an inhibition index of 58% at 1 mg/mL [57].
An interesting combination is coupling inulin with probiotics Lactobacillus sp. and Lactococcus sp. These systems have been proven symbiotic with regulatory properties for the growth of beneficial bacteria and antibacterial activity [80]. On the other hand, chitosan-inulin conjugates were obtained through the Maillard reaction with different pHs, corroborating that the conjugates with low pH values presented a more significant antimicrobial activity [58]. Another application of the inulin–chitosan conjugates is against bacterial biofilms. It has been shown that they significantly improved the activities against S. aureus biofilms and greatly inhibited their formation [59]. Wahbi et al. demonstrated that composite nanofibers (CNF) of inulin/polyvinyl alcohol (PVA) had an increased antibacterial activity against E. coli and S. aureus [60]. Finally, a probiotic nanocomposite film based on carboxymethylcellulose (containing cellulose nanofiber) and inulin was developed. It displayed antibacterial activity against nine pathogens, including S. aureus, E. coli, and K. pneumoniae [61].
Alginate
Alginate is a natural polysaccharide composed of mannuronic acid and guluronic acid. It has remarkable biocompatibility and biodegradability properties and is non-toxic to human cells [81]. This biopolymer is extracted mainly from brown algae (Phaeophyceae) through treatments with aqueous alkaline solutions. The extract is filtered and mixed with calcium chloride to obtain the alginate salt precipitate. The alginate salt is then converted to alginic acid by treating it with dilute hydrochloric acid (HCL) to purify it next to produce the water-soluble sodium alginate (Na-Alg or SA) [81].
Recently, the antimicrobial properties of sodium alginate (SA) have been studied in combination with other compounds; an example of this is zinc oxide-cellulose sodium alginate nanocomposite fibers (ZnO-SACNF) that have shown significant antibacterial activity against E. coli [62]. At the same time, copper alginate hydrogels in the form of microspheres produced immediate bactericidal effects against E. coli and S. aureus [63]. Moreover, it has been found that sodium alginate film with 5% hydroxyapatite nanoparticles (HA NPs) showed the highest antibacterial effect against Listeria monocytogenes [64]. Another study described the green synthesis of sodium alginate (SA)-silver nanoparticles (AgNPs), where SA acted as a stabilizer for AgNPs. These NPs induced cell death due to increased membrane permeability and disruption of the bacterial wall in S. aureus and E. coli [65].
Sodium alginate has also been used to synthesize zinc alginate films with and without 1% graphene oxide (GO), showing high antibacterial activity against S. aureus. Unfortunately, several mechanical properties were affected by incorporating 1% w/w of GO [66]. Similarly, crosslinked alginate/hydroxypropyl methylcellulose/ε-polylysine films with low content of plasticizers showed 99.9% bacterial reduction in both E. coli and S. aureus strains. It is noted that ε-PL should be used moderately because it competes for alginate chains, making the films more fragile [67]. Lastly, a novel antimicrobial nanocomposite was developed based on corona treated polypropylene (CPP), added with alginate and copper oxide nanoparticles; this nanocomposite exhibited excellent antimicrobial activity against E. coli, S. aureus, and Candida albicans [68].

Electrospinning and 3D Printing as a Novel Polymer Synthesis Technology

Among the most recent synthesis strategies and technologies to produce nanomaterials and their combinations, electrospinning and three-dimensional (3D) printing holds great promise as a production method for scaffolds, nanofibers, and supporting nanomaterials [82]. Electrospinning is a simple and versatile method for creating nanofiber materials and scaffolds with various structures and surface areas, which has a broad application field. Electrospinning involves converting a liquid polymer solution into solid nanofibers by applying an electrical force in the presence of a strong electric field [77]. Interestingly, electrospun nanofibers materials based on synthetic and natural antibacterial polymers have been studied due to their biocompatibility, biodegradability, and non-toxicity [83,84]. On the other hand, 3D printing is a relatively new, rapidly expanding method of manufacturing that allows the formation of 3D compact structures with desired and predefined architecture [85]. The use of nanomaterials (including a wide range of polymers, metals, composites, or natural products) in 3D printing is gaining attention due to the tremendous functionality this approach provides for those materials [86,87]. Nevertheless, further research is needed in this area, including using other polymeric materials, antimicrobial agents, and methods of producing filaments with selected properties.

2.2.2. Inorganic Nanoparticles

Another alternative to combat the ESKAPE bacteria is inorganic nanoparticles (INPs). Some of the most important INPs are gold nanoparticles (AuNPs), silver nanoparticles (AgNPs), magnetite NPs (MNPs), titanium oxide nanoparticles (TiO2NPs), zinc oxide nanoparticles (ZnONPs), and Ag-Au core-shell NPs [88,89,90]. INPs inhibit bacterial growth through different mechanisms [91,92]. However, combinatorial treatments of INPs with antibiotics, polymers, and antimicrobial peptides, can produce a synergistic effect and reduce the therapeutic doses [93,94].
The combination between INPs and polymers can be achieved through two strategies: using the polymer as a coating for the INPs or incorporating the nanoparticles into the polymer matrix [95,96]. There are different approaches to coating the surface of a nanoparticle with polymers; the most recognized is the modification of the surface through chemical treatments, ligand exchange techniques, and grafting techniques. Chemical surface treatments refer to a method of changing the structure and state of the surface of a nanoparticle by chemical reaction or chemisorption between the surface of the nanoparticle and the treatment agent [97,98]. The ligand exchange technique involves the substitution of one or more ligands in a complex ion with one or more different ligands. Usually, in nanoparticles, this technique is used to change the capping agent used in the original synthesis [97,99]. Grafting polymers to the surface of a nanoparticle enhances the chemical functionality and alters the surface of the INPs. Because monomers usually have a low molecular weight, they can penetrate nanoparticles and react with the activated sites on the nanoparticle surface, becoming partially filled with grafted macromolecular chains. Therefore, the aggregated nanoparticles become further separated [100,101].
On the other hand, two general approaches can be used to incorporate INPs into polymers: in situ and ex situ. In situ refers to using the polymer matrix as the reaction medium. Typically, the polymer is chemically modified to incorporate functional groups into its matrix that serve to reduce and form INPs on itself. Contrary to the in situ methodology, ex situ refers to the fact that the INPs particle is synthesized before incorporating into the polymer [95]. There is a recent interest on the part of the scientific community to create nanoscale polymeric structures to use them as nanocarriers. Some research groups use these nanovehicles to improve the delivery or transport of new therapies such as antimicrobial peptides [102]. Meanwhile, green synthesis has also led different research groups to create new synthesis routes for combinatorial therapies, facilitating the elimination of toxic compounds involved in conventional chemical synthesis and unwanted by-products by taking advantage of different organic residues [103,104].
Table 2 summarizes the combinatorial treatments that include inorganic nanoparticles as the main component of effective antimicrobial agents.

Gold Nanoparticles

Gold nanoparticles (AuNPs) are of great interest due to their physicochemical, optical, biocompatibility, and low toxicity properties. Several authors have evaluated the antimicrobial activity of AuNPs; however, there is still no consensus on whether this nanomaterial has antimicrobial activity. Although several authors have shown that AuNPs do not have antibacterial activity [105,106,107,108,109,110,111,112,113], others have shown that they do [114,115,116,117]. Despite this, in this review, we analyzed the combinatorial therapies where AuNPs were functionalized with different protection agents and polymers or served as a nanocarrier for antibiotics and antimicrobial peptides, among others.
For example, Feng et al. evaluated the activity of AuNPs functionalized with N-heterocyclic compounds and assessed the antimicrobial effect against methicillin-resistant S. aureus (MRSA) and MDR P. aeruginosa. The results showed that 2-mercaptoimidazole-functionalized AuNPs had low cytotoxic activity in HUVEC cells (human umbilical vein endothelial cell) and excellent antibacterial effect in both strains [118]. Besides, Sun et al. synthesized AuNPs with 4,5-diamino-2 pyrimidiethiol (DAPT) and bovine serum albumin (BSA). This compound (AuNPs- DAPT-BSA) exhibited a bactericidal effect (killed up to 99%) in MDR E. coli and MRSA bacteria [119].
In recent years, the scientific community has focused on producing eco-friendly nanoparticles using compounds from extracts of plants, fungi, bacteria, algae, and actinomycetes, to reduce costs and mitigate the toxic effects of their precursors [120,121,122,123,124,125]. A study demonstrated that AuNPs synthesized from the extract of Garcinia mangostana L. combined with conventional antibiotics possess a synergistic antimicrobial effect against clinical isolates of Staphylococcus spp. and Pseudomonas spp. The results showed that AuNPs combined with azithromycin and streptomycin increased the antibacterial activity (34.8% and 33.3%, respectively) against Staphylococcus spp. compared to antibiotics alone. On the other hand, the combination of AuNPs with penicillin and azithromycin exhibited better antimicrobial activity (75% and 50%, respectively) against Pseudomonas spp. [126].
Combining AuNPs with polymers is another promising treatment strategy against resistant bacteria [127,128,129]. For example, Pradeepa et al. synthesized AuNPs with an exopolysaccharide (EPS) (extracted from Lactobacillus plantarum) and functionalized them with antibiotics (e.g., ciprofloxacin, levofloxacin, ceftriaxone, and cefotaxime). The treatments showed a synergistic antibacterial effect against MDR strains of K. pneumoniae, S. aureus, and E. coli. The combinations that showed better results were AuNP-ciprofloxacin against K. pneumoniae and E. coli; and AuNP-levofloxacin against S. aureus [130]. Furthermore, Li et al. showed that bacterial cellulose decorated by 4,6-diamino-2-pyrimidinotiol-modified AuNPs inhibited bacterial growth of E. coli, MDR E. coli, P. aeruginosa, and MDR P. aeruginosa [131]. Finally, a bioconjugate containing AuNPs, polycobaltocenium (PCo) homopolymer, and penicillin-G (Peni) exhibited remarkable antimicrobial efficiency (synergistic effect) against S. aureus and E. coli compared to the individual PCo-Peni and Peni treatments [132].
Table 2. Combinatorial treatments that include inorganic nanoparticles as the main component of effective antimicrobial agents.
Table 2. Combinatorial treatments that include inorganic nanoparticles as the main component of effective antimicrobial agents.
NanomaterialCombined with (Rate/Ratio)FormSizeTargeted BacteriaAntimicrobial EffectsReferences
Gold Nanoparticles (AuNPs)2-mercaptoimidazole
(MI, 10:1 AuNPs)
Spherical AuNPs
MI capping
AuNPs diameter ~ 3.5 nm MRSA
MDR P. aeruginosa
Excellent antimicrobial effects with low cytotoxic activity in HUVEC cells.[118]
4,5-diamino-2
pyrimidiethiol (DAPT)
Bovine serum albumin (BSA)
(21:1 BSA/DAPT)
Spherical AuNPs
DAPT and BSA
capping
AuNPs diameter 4.11
± 0.32 nm
MDR E. coliKilled up to 99% of bacteria[119]
Azithromycin (Azi, 3:1 Azi/AuNPs)
Streptomycin (Sty 1:1 Sty/AuNPs)
Spherical AuNPs
AuNPs disc
impregnated with antibiotic solution
AuNPs diameter between 20 to 40 nmClinical isolates
Staphylococcus spp.
Increased antibacterial activity
compared to antibiotics alone
[126]
Penicillin G (PeG, 1:5 PeG/AuNPs)
Azithromycin (3:1 Azi/AuNPs)
Clinical isolates
Pseudomonas spp.
Ciprofloxacin (4.3 µg
of antibiotic conjugated/mL)
Spherical AuNPs
Antibiotic
conjugation to AuNPs surface
Bare AuNPs diameter 10–20 nm
Functionalized AuNPs diameter 20–30 nm
MDR K. pneumoniae
MDR E. coli
Synergistic antibacterial effect[130]
Levofloxacin (3.87 µg
of antibiotic
conjugated/mL)
MDR S. aureus
Bacterial cellulose
(BC)
4,6-diamino-2-
pyrimidinotiol (Au-DAPT, 3.3 ± 0.3
µg/cm2)
BC membrane for wound dressing
decorated
Spherical AuNPs capped with DAPT
Au-DAPT NPs diameter
≈3 nm
E. coli,
MDR E. coli, P. aeruginosa and
MDR P. aeruginosa
Inhibited bacterial growth[131]
Polycobaltocenium
homopolymer (PCo,
38% w/w)
Penicillin G (PeG 27% w/w)
Spherical AuNPs capped with PCo and functionalized with PeGBare AuNPs diameter 2–3 nm
Functionalized AuNPs
diameter to 6 nm (Au@PCo)
S. aureus
E. coli
Synergistic effect compared with individual treatments[132]
Silver nanoparticles
(AgNPs)
Vancomycin, Oleandomycin, Ceftazidime, Penicillin G, Novobiocin, Carbenicillin, Lincomycin, and Erythromycin (15 µg/disc)Spherical AgNPs
Antibiotic disk impregnated with AgNPs (500 ppm)
AgNPs diameter 15–20 nmMDR P. aeruginosa
MDR E. coli
Synergistic effect compared with individual treatments[133]
Ampicillin (Amp) and amikacin (Amk) (1:1 Antibiotic/AgNPs)Spherical AgNPs functionalized with antibioticsBare AgNPs diameter 8.57 ± 1.17
AgNPs +Amp diameter 4.01 ± 0.80
AgNPs +Amk diameter 6.03 ± 0.87
Clinical isolates of
E. faecium, S. aureus,
A. baumannii, Enterobacter cloacae, E. coli, K. pneumoniae, and P. aeruginosa.
Synergistic, partial synergistic and additive antibacterial effects among the different combinations[134]
Bacteriocin extracted from Lactobacillus paracasei (nd)Spherical AgNPs conjugated with bacteriocinAgNPs diameter ~16 nmClinical MDR isolates
of S. aureus,
P. aeruginosa, K. pneumoniae, E. coli, and Staphylococcus pyogenes
Synergistic bactericidal effect compared to individual treatments[135]
Polyvinyl alcohol (PVA)
Chitosan (CS)
PVA-AgNPs and CS-AgNPs nanocomposite films
Spherical AgNPs
AgNPs diameter ~15 nmClinical isolates of
S. epidermis, S. aureus, K. pneumoniae, and
E. coli
Remarkable antimicrobial effect and inhibition of biofilm production[136]
Zinc Oxide nanoparticlesCefepime (0.0256 μg/mL)
Ampicillin (0.001 μg/mL)
Antibiotics in solution
Spherical ZnO NPs (80 μg/mL)
ZnO NPs diameter ~15 nmClinical isolates
of E. coli
Synergistic effect[137]
(ZnO NPs)Cephotaxime (0.032 μg/mL)
Ceftriaxone (0.1 μg/mL ceftriaxone)
Antibiotics and NPs in solution
Spherical ZnO NPs (60 μg/mL)
Clinical isolates of
K. pneumoniae
Ciprofloxacin (8 mg/mL)
Ceftazidime (32 mg/mL)
Antibiotics and NPs in solution
Spherical ZnO NPS
ZnO NPs diameter ~17.08 nmClinical isolates of
A. baumannii
Increased antimicrobial activity to overcome bacterial resistance[138]
Ciprofloxacin (nc)Ciprofloxacin conjugated to ZnO NPs
Multiple shapes of ZnO NPs
ZnO NPs diameter 20–24 nmKlebsiella spp. and
E. coli.
Increased antibacterial activity compared to individual treatments[139]
Colistin (1–4 μg/mL)Colistin and ZnO NPs in solution
ZnO NPs form n.d
ZnO NPs diameter 50 nm Clinical isolates of
P. aeruginosa
Synergistic effect[140]
Chitosan NPs (1:1 ZnO NPs/chitosan)Chitosan NPs and ZnO NPs in solution
ZnO NPs form n.d
ZnO NPs n.d.MDR E. coli
MDR E. faecium
Synergistic effect[141]
Lipid micelle (5:8 mass Lipid/ZnO NPs)
Chitosan (5:24 mass chitosan/ZnNPs
Lipid nanomicelles
Spherical ZnO NPs inside micelle
Chitosan capping micelles
ZnO NPs form n.d
Micelle diameter ~338.7 nm
ZnO NPs n.d.
MDR E. faecium50% reduction in bacterial biofilm formation[142]
EPS from Rhodotorula mucilaginosa UANL- 001L (2 mg/mL)EPS-ZnO NPs nanobiocomposite
ZnO NPS without defined shape
ZnO NPs diameter 8.32±1.99 nm.MDR P. aeruginosa
MDR S. aureus
Inhibition of bacterial growth (50–80%)[143,144]
No visible toxic effects in a Wistar rat model
Titanium dioxide nanoparticles
(TiO2 NPs)
Two geometric isomers ferrocene-carborane derivatives (FcSB, 0.5–1:4 FcSB/ TiO2 NPs)FcSB and TiO2 NPs in solution
Spherical TiO2 NPs
TiO2 NPs diameter 41 ± 12 nmClinical MDR isolates of A. baumannii100% inhibition of growth[145]
ZnO NPs (nd)TiO2 NPs and ZnO NPs in solution
Spherical TiO2 NPs and ZnO NPs
TiO2 NPs and ZnO NPs diameter between 20–50 nmClinical MDR isolates of A. baumannii, and K. pneumoniaeAdditive effects[146]
Silver ions (Ag+ 8% w/w)TiO2 anathase phase NPs shell with Ag+ incorporated
Spherical TiO2 NPs
TiO2 NPs diameter 200 ± 10 nm with a wall thickness of 20–30 nm MDR S. aureusStrong antibacterial activity[147]
Polytetrafluorethylene (PTFE, 2 g/L)TiO2 NPs- PTFE particles coated in a stainless-steel
surface
TiO2 NPs diameter < 25 nm,
PTFE particles 200–300 nm
E. coli
S. aureus
Antibacterial and anti-adhesion properties.[148]
ZnO NPs (1:3 TiO2 NPs/ZnO NPs)TiO2 NPs and ZnO NPs in solution
Shape n.d.
Size n.d.S. aureus ATCC29213, E. coli ATCC 25922, MRSA ATCC 38591, and K. pneumoniae ATCC 700603Bactericidal activity[149]
MRSA
MDR K. pneumoniae
50% reduction in biofilm
Cefepime
Ceftriaxone
Amikacin
Ciprofloxacin
(Sub-MIC values)
TiO2 NPs and antibiotics in solution
Irregulate shape TiO2 NPs
TiO2 NPs particle size 64.77 ± 0.14 nmMDR P. aeruginosaSynergistic effect[150]
Erythromycin (2–16 mg/L)TiO2 NPs and erythromycin in solution
“Round-shape” TiO2 NPs
TiO2 NPs size 15–18 nm.MRSASynergistic effect[151]
Silver (1.4% of nanoparticle)/
rifampicin, doxycycline, ceftriaxone, and cefotaxime (66.4, 60.3, 34.0 and 23.6 μg/mg of nanocomposite, respectively)
Antibiotics attached via electrostatic interactions Fe3O4/Ag NPs
Roundish shape Fe3O4/Ag NPs
Fe3O4/Ag NPs diameter 40–50 nm in sizeS. aureus and Bacillus pumilusAntibacterial properties[152,153,154]
Magnetite nanoparticles
(Fe3O4 NPs)
Cefepime (3.53 ± 0.1% w/w of NP
PLGA (film)
PLGA/Fe3O4-Ce NPs composite films -Spherical Fe3O4 NPs functionalized with cefepime (Fe3O4-Ce NPs)Fe3O4/Ce NPs diameter ~5 nm S. aureus and E. coliSuitable materials for the sterilization on implantable devices, biocompatible and efficient inhibition of bacterial biofilm[155]
EugenolFe3O4 NPs functionalized with Eugenol
“Quasi-spherical” shape Fe3O4 NPs
Fe3O4 NPs size < 10 nmS. aureus
P. aeruginosa
Excellent anti-adherence and anti-biofilm properties.
Low toxicity and an easily biodegradable material
[156]
Chitosan (1:5 Fe3O4 NPs/chitosan)Chitosan- Fe3O4NPs composites
Fe3O4 NPs coating
Fe3O4 NPs form nd
Fe3O4 NPs size ndE. coliAntibacterial properties
Dye absorbent
[157]
Cathelicidin LL-37 (128 μg/mL)
Ceragenin CSA-13 (0.5–8 μg/mL)
Fe3O4 NPsand peptides in solution
Spherical Fe3O4 NPs
Fe3O4 NPs diameter ~12 nmMRSA Xen 30, and P. aeruginosa Xen 5Antibacterial properties[158]

Silver Nanoparticles

Silver nanoparticles (AgNPs) are another promising nanomaterial to fight against infections caused by MDR bacteria. From ancient times, silver has been used due to its antimicrobial capacity. At present, silver is used in nanoparticles with applications in the food and medical industry [159,160,161,162]. Despite the enormous potential of these nanoparticles, some studies have shown that AgNPs have cytotoxic effects in different cell lines depending on their size, shape, concentration, or capping agent [163,164,165]. For this reason, combinations with other antimicrobial treatments have been proposed to improve the antibacterial activity and reduce the AgNPs’ cytotoxicity.
Many authors have shown that AgNPs combined with antibiotics synergize against MDR strains [166,167,168,169]. Li et al. synthesized AgNPs combined with antibiotics (vancomycin, oleandomycin, ceftazidime, penicillin G, novobiocin, carbenicillin, lincomycin, or erythromycin) showed a synergistic effect against MDR P. aeruginosa and E. coli [133]. In another study, AgNPs were conjugated with ampicillin and amikacin (1:1 ratio) and tested against clinical isolates of E. faecium, S. aureus, A. baumannii, Enterobacter cloacae, E. coli, K. pneumoniae, and P. aeruginosa. The results showed synergy, partial synergy, and additive effects among the conjugates [134].
Other studies have shown that the combination of bacteriocins and AgNPs is an effective treatment against ESKAPE pathogens [170,171,172,173]. For example, Gomaa used a bacteriocin extracted from Lactobacillus paracasei that, in combination with AgNPs, exhibited a more potent synergistic bactericidal effect than individual treatments against clinical MDR isolates of S. aureus, P. aeruginosa, K. pneumoniae, E. coli, and Staphylococcus pyogenes [135].
One of the strategies to reduce the AgNPs cytotoxicity is using polymers as capping agents or for controlled release [174,175,176,177]. For example, Abdalla et al. used polyvinyl alcohol (PV) and chitosan (C) as capping agents in AgNPs and were tested against clinical isolates of S. epidermis, S. aureus, K. pneumoniae, and E. coli. Interestingly, this combination had a remarkable antimicrobial effect and inhibited biofilm production in all analyzed strains [136].
Furthermore, AgNPs can be produced by alternative forms. For example, Figueiredo et al. have carried out studies on fungal synthesis. In this study, the AgNPs were functionalized with simvastatin and then tested against reference and MDR bacterial strains. The combination showed antibacterial activity against MRSA strains and extended-spectrum beta-lactamase E. coli. These results demonstrated the great potential of this combinatorial treatment to combat bacterial infections [178]. Another example of an alternative synthesis of AgNP is photo-reduction. Courrol et al. functionalized AgNPs with tryptophan and evaluated their effect on antibiotic-resistant and susceptible pathogens. The results showed inhibition of ~100% (bactericidal effect) in Salmonella thipymurium, P. aeruginosa, S. aureus, MDR S. epidermis, MDR K. pneumoniae, and MDR E. coli. The inhibition of biofilm formation was also investigated, and data showed an inhibitory effect in E. coli, K. pneumoniae, C. freundii, P. aeruginosa, S. aureus, and S. epidermidis [179].
Finally, one of the most recent advances involving silver nanomaterials is silver phosphate nanoparticles (Ag3PO4 NPs). Among semiconductor nanomaterials, Ag3PO4 NPs have attracted considerable attention due to their photocatalytic activity, which has the potential to generate ROS for the degradation of organic dyes under visible light irradiation, making them a candidate for killing pathogenic bacteria [180,181,182]. Steckiewicz et al. evaluated the antimicrobial and antibiofilm properties of Ag3PO4 with different morphologies. The authors observed that the cubic morphologies had better antimicrobial activity, obtaining an 8 μg/mL MIC against S. aureus ATCC 25923 and MRSA ATCC 33591. In contrast, minimum biofilm eradication concentrations (MBECs) of 32 μg/mL and 16 μg/mL were observed for S. aureus ATCC 25923 and MRSA ATCC 33591, respectively. Unfortunately, these nanoparticles have cytotoxic effects on osteoblast line cells (hFOB1.19, MC3T3-E1, Saos-2, C2C12, and HD) starting at 5 μg/mL [183]. So far, there is limited literature on their antimicrobial properties and cytotoxicity. Therefore, more studies are needed to show the potential application of this promising nanomaterial.

Zinc Oxide Nanoparticles

Zinc oxide nanoparticles (ZnO NPs) are widely used in different areas, such as cosmetics, solar panel development, biosensors, and medicine [184,185,186]. Lately, ZnO NPs have received considerable attention due to their efficient bactericidal activity against multidrug-resistant bacteria [187,188]. Despite their excellent antimicrobial properties, some studies have shown the potential toxicity in different cell lines [189]. Some alternatives have been sought to counteract these effects, such as the combination with other antimicrobial compounds.
Several investigations have reported synergistic activity between ZnO-NPs and antibiotics. For example, Bhande et al. used ZnO NPs/β-lactam antibiotics-based combinatorial therapies to treat beta-lactamase-producing bacteria (clinical isolates from urinary tract infections). The combination of ZnO NPs with cefepime or ampicillin was synergistic in E. coli. In contrast, the combination of ZnO NPs with cefotaxime or ceftriaxone revealed a similar synergy pattern in K. pneumoniae [137]. In another study, the combination of ZnO NPs with ciprofloxacin and ceftazidime was tested against resistant A. baumannii. The results showed that the antimicrobial activity of ciprofloxacin and ceftazidime was increased in the presence of ZnO NPs. Thus, this combined strategy effectively overcame bacterial resistance [138]. Recently, Tyagi et al. synthesized ZnO-NPs conjugated with ciprofloxacin and evaluated its effect against Klebsiella spp. and E. coli. This conjugate increased the antibacterial activity against Klebsiella spp. and E. coli by 4.3- and 2.7-fold, respectively, compared to ZnO-NPs alone [139]. Moreover, a recent study investigated the effect of ZnO NPs in combination with meropenem, ciprofloxacin, and colistin against P. aeruginosa ATCC strain and clinical isolate strains of P. aeruginosa. Notably, the combination of ZnO NPs with colistin showed a synergistic effect, and it might be beneficial as a therapeutic alternative to P. aeruginosa infections [140].
During the last five years, different research groups have sought to use natural polymers in combination with ZnO NPs as a possible treatment against multidrug-resistant bacteria [141,190,191,192,193]. A group of researchers evaluated the antibacterial effects of NPs of chitosan, ZnO alone, and a combination of chitosan and ZnO against MDR and wild-type strains. ZnO combined with chitosan showed synergistic effectiveness on MDR E. coli and MDR E. faecium [141]. A few years later, Mehta et al. tested biofilm activity against MDR E. faecium with the previously described composite (ZnO-chitosan composite) in a lipid micelle. Interestingly, results showed a 50% reduction in bacterial biofilm size than chitosan and ZnO alone [142]. Finally, our research group has synthesized a novel biocomposite that serves as an antimicrobial against MDR S. aureus and MDR P. aeruginosa. This biocomposite was composed of an exopolysaccharide (EPS) from Rhodotorula mucilaginosa UANL-001L as a capping agent that previously showed antibacterial properties [144] and ZnO-NPs. We showed that the biocomposite inhibited bacterial growth up to 50 and 80% in MDR P. aeruginosa and MDR S. aureus. Furthermore, we have determined the toxicological effects of this biocomposite in a Wistar rat model. Our results showed that a three-day oral regimen of 6 mg/mL did not produce any toxic adverse effects in the rats, so we concluded that this composite has potential applications as an antimicrobial agent against resistant strains [143].

Titanium Dioxide Nanoparticles

Along with ZnO NPs, titanium dioxide nanoparticles (TiO2 NPs) are among the most promising metal oxide nanoparticles for pathogen control. Due to their whitish color, these NPs are used as an additive in the paint, cosmetics, and food industry [194]. Moreover, they have excellent catalytic properties and are widely used for water and air treatment, organic synthesis, fertilizer production, and, finally, as a product for decontamination and as a food preservative due to their antimicrobial activity [195,196]. Despite its wide applications, various studies have shown that TiO2 NPs are quite toxic in animal models and even classified as possibly carcinogenic in humans by The International Agency for Research on Cancer (IARC) thus reducing their harmful effects is vital to continue its application [195,197,198,199]. As seen throughout this review, combinatorial treatments can reduce toxicity by decreasing individual concentrations. Thus, combinatorial strategies seem to be an option to improve therapies with TiO2-NPs.
Many research groups have been dedicated to improving the antibacterial activity of TiO2 NPs by combining them with other nanoparticles or nanomaterials [200,201,202]. In the first example of TiO2 NPs-based combinatorial treatments, we present the study by Li et al., where they observed synergy between TiO2 NPs and two geometric isomers carborane derivatives (named FcSB1 and FcSB2) as a treatment against clinical isolates of MDR A. baumannii. Results showed that by combining fractional MIC values, up to 100% of MDR A. baumannii growth could be inhibited [145]. Then, Masoumi et al. studied the antimicrobial properties of TiO2 NPs, ZnO NPs, and synthetic peptides against MDR clinical isolates of A. baumannii, K. pneumoniae, and P. aeruginosa. This combination showed additive effects against A. baumannii and K. pneumoniae. The combination of nanoparticles (TiO2 NPs and ZnO NPs) had an additive effect on A. baumannii and K. pneumoniae strains [146]. Recently, Hérault et al. tested the antibacterial effect of silver-containing titanium dioxide nanocapsules against E. coli, S. aureus, and MDR S. aureus. The authors demonstrated that these nanocapsules had great antibacterial activity against wild-type and MDR S. aureus. Thus, this result showed the potential biomedical use of this therapeutic strategy [147].
Other studies have proven that the combination of TiO2 NP with different polymers can be used against MDR bacteria [203,204]. For example, Zhang et al. synthesized a nanocomposite of polytetrafluorethylene (PTFE) and TiO2-NPs against E. coli and S. aureus. In this study, the nanocomposite showed antibacterial and anti-adhesion properties against both bacteria, and it also demonstrated biocompatibility in fibroblast cells. Thus, this nanocomposite is a promising candidate for biomedical implants [148]. Recently, Harun et al. used a TiO2/ZnO nanocomposite against non-MDR and MDR bacterial strains (S. aureus ATCC29213, E. coli ATCC 25922, MRSA ATCC 38591, and K. pneumoniae ATCC 700603). Results showed the bactericidal activity of heterogeneous TiO2/ZnO (25T75Z molar ratio) nanocomposite in Gram-positive and -negative bacteria. Interestingly, the authors also observed a 50% reduction in the biofilm formation in MRSA and MDR K. pneumoniae [149].
Similarly, different research groups have used TiO2-NPs to improve the antibacterial activity of different antibiotics in MDR strains [205,206]. A recent study evaluated the antibacterial effect of TiO2-NPs in combination with antibiotics against MDR P. aeruginosa strains. The combination of TiO2-NPs with cefepime, ceftriaxone, amikacin, and ciprofloxacin exhibited synergistic activity against all tested isolates [150]. Finally, a recent study demonstrated the enhanced antibacterial activity of combination erythromycin/TiO2-NPs against MRSA. The combination was found to be more potent than individual treatments. This therapeutic strategy seems to be a potential alternative to conventional antibiotics to treat MRSA strains [151].

Magnetite Nanoparticles

Magnetite can be used as a nanomaterial due to its magnetic and biological properties, such as thermal properties and chemical and colloidal stability [207]. Although it is mentioned in some works with antibacterial and biofilm activities, the doses to achieve these effects are quite high [208]. Therefore, an alternative is to use it in combinatorial treatments.
One of the unconventional methods of sterilizing medical equipment is developing multifunctional bioactive coatings capable of inhibiting microbial proliferation. Within this method, we can find that structures formed by a spherical core of magnetite functionalized with cefepime (Fe3O4 @ CEF) and a thin layer of PLGA have been satisfactorily synthesized, which proved to be suitable materials for the sterilization of the surface of implantable devices, in terms of improved biocompatibility and efficient and constant inhibition of bacterial biofilm (staphylococcal and Escherichia coli colonization) [155]. Likewise, water-dispersible nanostructures based on magnetite (Fe3O4) and eugenol (E) have been synthesized. These nanostructures have proven to be another successful alternative to control and prevent infections associated with microbial biofilms of S. aureus and P. aeruginosa. This nanomaterial showed excellent anti-adherence and anti-biofilm properties, besides using bioactive natural compounds, representing lower toxicity and an easily biodegradable material [156]. Later, chitosan and magnetite nanoparticles biocomposites (Cs/Fe3O4) were synthesized for efficient dye adsorption and as an antibacterial agent. Besides the excellent adsorbent efficiency, the composite showed an antibacterial efficacy against E. coli [157]. Finally, a study reported the bactericidal activity of MNPs combined with the human antibacterial peptide cathelicidin LL-37, ceragenins, or classical antibiotics (vancomycin and colistin) against MRSA Xen 30 and P. aeruginosa Xen 5. The combination of LL-37 peptide or ceragenin CSA-13 with MNPs showed antibacterial properties, suggesting an alternative to developing new methods to treat infections caused by MDR bacteria [158].
Last, a research group conducted magnetite/silver/antibiotic (rifampicin, doxycycline, ceftriaxone, and cefotaxime) nanocomposites for targeted antimicrobial therapy. Antimicrobial properties of magnetite/silver/rifampicin and magnetite/silver/doxycycline nanocomposites were confirmed in S. aureus and Bacillus pumilus [152,154]. Then, the cytotoxicity of these nanocomposites was evaluated in human embryonic kidney 293 (HEK293T). Unfortunately, the results showed that magnetite/silver/rifampicin and magnetite/silver/doxycycline NPs induced cytotoxicity in the tested cell line [153]. Thus, more cytotoxicity studies should be performed before the clinical application of these nanocomposites.

2.2.3. Antimicrobial Peptides

Antimicrobial peptides (AMPs) are a family of naturally occurring antimicrobial low-molecular-weight proteins with a broad spectrum of antimicrobial activity [209]. In recent years, research on antimicrobial peptides has increased. Several AMPs are in clinical development, and some others have undergone clinical trials; however, a few have been successfully commercialized [209,210,211]. AMPs include molecules such as defensins, cathelicidins, granulysin, S-100 proteins, and colistin, among others [212]. Colistin is one of the seven US Food and Drug Administration (FDA)-approved AMP [213]. The Gram-positive bacterium Paenibacillus polymyxa produces it and is currently being used as an antibiotic (last-resort drug) to treat multidrug-resistant Gram-negative bacteria [214,215].

AMPs-Based Combinatorial Treatments

Combinatory therapies mainly lead to synergism or additive antimicrobial effects; therefore, the combination has a stronger effect than a single drug. Moreover, the use of combinatory therapies represents a clinical improvement and significantly decreases the possibility of antibacterial resistance [216]. Several studies have shown the advantages of therapies based on AMPs and antibiotics against MDR and biofilm-forming bacteria [22]. A summary of the combinatorial treatments that include antimicrobial peptides as the main component of effective antimicrobial agents is presented in Table 3.
Human neutrophil peptide (HNP-1) is one of the most potent defensins produced by neutrophils [217]. A study showed that the combination of antituberculosis antibiotics isoniazid and rifampicin with HNP-1 had a better antimicrobial effect against Mycobacterium tuberculosis from infected lungs, liver, and spleen than they were employed alone [218]. The synthetic LL-37 is the only cathelicidin-derived antimicrobial peptide found in humans [219]. The evaluation of the antibacterial activity of amoxicillin with clavulanic acid and amikacin combined with the synthetic LL-37 peptide revealed a more significant killing effect against clinical isolates of S. aureus [220]. Another peptide is arenicin-1, which has been reported to exhibit broad-spectrum antimicrobial activity [221]. Combining this peptide with conventional antibiotics (ampicillin, erythromycin, and chloramphenicol) demonstrated synergistic activity and killed bacteria by interfering with the biosynthesis of DNS, proteins, or cell wall components [221].
A recent study has demonstrated that the synthetic cyclolipopeptide analog of polymyxin (AMP38) was tested in combination with carbapenems. The synergistic effect was observed to cause the killing of biofilm-forming and carbapenem-resistant P. aeruginosa [222]. On the other hand, the synergistic effects of antimicrobial peptide DP7 on some multidrug-resistant bacterial strains (S. aureus, P. aeruginosa, and E. coli were evaluated). The results showed that the combination of DP7 peptide with azithromycin or vancomycin was more effective, especially against highly antibiotic-resistant strains [223].
Table 3. Combinatorial treatments that include antimicrobial peptides as the main component of effective antimicrobial agents.
Table 3. Combinatorial treatments that include antimicrobial peptides as the main component of effective antimicrobial agents.
NanomaterialCombined withTargeted BacteriaAntimicrobial EffectsReferences
Human neutrophil peptideIsoniazid and rifampicinMycobacterium tuberculosisAntimicrobial effect[218]
Synthetic LL-37 (cathelicidin-derived peptide)Amoxicillin with clavulanic acid and amikacinClinical isolates of S. aureusSignificant killing effect[220]
Arenicin-1Ampicillin, erythromycin, and chloramphenicolStaphylococcus aureus (ATCC 25923), Enterococcus faecium (ATCC, 19434), Staphylococcus epidermidis (KCTC 1917), Pseudomonas aeruginosa (ATCC 27853), E. coli (ATCC 25922), and E. coli O-157 (ATCC 43895)Synergistic activity and kill bacteria by interfering with biosynthesis of DNS, proteins, or cell wall components[224]
Synthetic cyclolipopeptide analog of polymyxin (AMP38)CarbapenemsCarbapenem-resistant
P. aeruginosa
Synergistic effect[222]
Peptide DP7Azithromycin or vancomycinMDR strains (S. aureus, P. aeruginosa, and E. coli)Antimicrobial effect[223]
ASU014Oxacillin MRSAImproved the killing effect[225]
A broad set of AMPsCiprofloxacin, meropenem, erythromycin, and vancomycinEnterococcus faecium, S. aureus, K. pneumoniae, A. baumannii, P. aeruginosa, Enterobacter cloacaeSynergistic effects[226]
Recent studies have recently shown the effect of a broad set of AMPs combined with antibiotics against drug-resistant strains. Combinatorial treatment of ASU014, a bivalent branched peptide, with oxacillin, was very efficient against MRSA. The synergism between both meaningfully improved the killing effect compared to single drugs. Lower peptide concentrations and sub-MIC doses of the antibiotic were required for the complete eradication of the pathogen [225]. Another study evaluated the synergy between conventional antibiotics (ciprofloxacin, meropenem, erythromycin, and vancomycin) and a broad set of AMPs in a murine model induced by ESKAPE (Enterococcus faecium, S. aureus, K. pneumoniae, A. baumannii, P. aeruginosa, Enterobacter cloacae) pathogens. Combinatorial treatments showed synergistic effects that significantly reduced abscess sizes and/or improved clearance of bacterial isolates from the infection site, regardless of the antibiotic mode of action. Therefore, these results open new opportunities to develop alternative therapies [226].
As we mentioned, antimicrobial peptides are an up-and-coming alternative for the treatment of infections caused by multidrug-resistant bacteria; however, they present some disadvantages such as their reliability due to their susceptibility to proteases and changes in pH, short half-life, and rapid renal clearance [227]. Therefore, it is possible to use AMPs in combination with drug delivery systems, such as liposome encapsulation, inorganic (AuNPs or AgNPs), and polymeric (chitosan, or PLGA) nanoparticles [228].
All the previously described data strongly support the idea that combinatorial treatments have increased antibacterial activity. However, evaluating these treatments in animal models and clinical trials is necessary to assess their efficacy and safety. We have summarized in Table 4 some of the most relevant studies of nanomaterial-based combinational treatments where their antibacterial activity was evaluated using in vivo models.
Although promising results have been obtained in preclinical studies, additional clinical research is required to elucidate the efficacy of combination treatments for clinical practice. So far, a few clinical trials have ventured into the development of combinatorial therapies, including inorganic and polymeric nanoparticles and AMPs for antibacterial treatment. Clinical trials of nanomaterial-based combinational treatments are summarized in Table 5.
A phase III study is currently underway in the literature to test the efficacy of the new chitosan containing lesion sterilization therapy and tissue repair (LSTR) material in a dual antibiotic paste. This study has been published, but the registry is not yet open [231]. The incorporation of different nanomaterials has been proposed to improve the antimicrobial properties of bioceramic seals. For example, a phase IV study seeks to test the efficacy of a combination of AgNPs or chitosan with bioceramic sealants to improve its antimicrobial properties using an infection model. The results of this study are not published yet [232,233]. One of the basic requirements for a good treatment of caries is to find a way to seal these areas and isolate them from microorganisms so that they do not enter and colonize [234]. As an example, we have a phase III clinical trial that evaluates the antibacterial effect of a glass ionomer to seal caries when incorporated with chitosan and TiO2 NPs on carious dentin treated after partial removal of caries. Unfortunately, the results of this trial have not been published yet. On the other hand, a phase III clinical study evaluated the antibacterial effect of a new product called Nano Silver Fluoride ® (a solution of AgNPs, chitosan, and fluoride) in occlusal carious molars treated with the partial caries removal technique to prevent and treat caries cavities. In this study, 44 people were brought together to analyze the antimicrobial properties of the so-called Nano Silver Fluoride ®. Of the 44 people, 22 were treated with Nano Silver Fluoride ®, and the rest with a control treatment. Surprisingly, Nano Silver Fluoride ® manages to reduce the count of bacteria present in Molars Treated by up to 44%.
Regarding the AMPs, the results obtained in pre-clinical studies demonstrated their potential to be evaluated in clinical trials to assess their efficacy and safety. To date, 41 AMPs have been evaluated in clinical phases; some are still under study, and others have been completed, discontinued, or approved. Currently, five AMPs (nisin, gramicidin, polymyxins, daptomycin, and melittin) are in clinical use as an alternative to conventional treatments [235]. However, only a couple of studies have evaluated the combinatorial treatment as a strategy to protect or reduce the toxicity of AMPs. Nisin (nisin A) is naturally produced by lactic acid bacteria (Lactococcus lactis). Applications of nisin in humans include dental care and pharmaceutical products to treat stomach ulcers and colon infections. Nisin is degraded by enzymes in the gastrointestinal (GI) tract; therefore, it requires a delivery system to reach the site of action without being digested and absorbed during its passage through the GI tract. A strategy to improve this, nisin was encapsulated in pectin/ hydroxypropyl methylcellulose-coated tablets [236]. Polymyxins (A, B, C, D, and E) are a group of cyclic polypeptides which are used to treat eye (polymyxin B) and wound (polymyxin E) infections. These AMPs are last-resource treatment options because of their common side effects, including nephrotoxicity and neurotoxicity. To reduce the occurrence of adverse effects, polymyxin B has been administered as a pro-drug that, when hydrolyzed, produces the active AMPs. Moreover, polymyxin E has been successfully incorporated into hydrogels to treat burn wound infections [237].
Table 5. Clinical trials of nanomaterial-based combinational treatments.
Table 5. Clinical trials of nanomaterial-based combinational treatments.
NanomaterialApplied withTrial Description Clinical Trial Identifier
Chitosan nanoparticlesDouble antibiotic pasteTo evaluate the clinical double antibiotic, paste mixed with chitosan nanoparticles gel in lesion sterilization and tissue repair in non-vital primary molars.NCT05079802
Silver nanoparticles and chitosanBioceramic sealerTo assess the antibacterial efficacy and adaptability of bioceramic sealer when incorporated with nanosilver.NCT04481945
Titanium dioxide nanoparticles and chitosanGlass ionomer To study antibacterial effect on carious dentine of glass ionomer when modified with chitosan and titanium dioxide nanoparticles.NCT04365270
Silver nanoparticles and chitosan FluorideEvaluation of the antibacterial effect of nano silver fluoride on occlusal carious molars.NCT03186261
NisinPectin/hydroxypropyl methylcellulose-coated tabletsDelivery system to reach the site of action without being digested.[236]
Polymyxin EHydrogelsTo treat of burn wound infections.[237]

3. Conclusions

Even though antibiotics have been a great weapon against pathogenic bacteria, it is urgent to create new therapies against pathogens that have developed strategies to overcome their mechanisms. Novel materials, such as polymers, nanoparticles, or AMPs, offer clear advantages over conventional antibiotics to combat various infectious diseases. However, it was also noted that some therapies have unwanted toxic effects on human cell lines. Only a few combinatorial treatments have reached the clinic. Challenges toward clinical applications of combination strategies include cytotoxicity effects, production costs, bioavailability, and efficacy. Several strategies have been designed to overcome these challenges, such as chemical modifications, delivery systems, and new synthesis technology (electrospinning and three-dimensional printing). The development of combinatorial treatments is a multidisciplinary field; recently, intense research on the synthesis strategies and sophisticated techniques to produce complex nanomaterials has led to the development of combinatorial therapies involving nanomaterials. Given the significant research in the field, it may be expected that humankind will greatly benefit from nanomaterials and their combinations in the very near future, especially in the treatment of multidrug-resistant bacteria.

Author Contributions

Conceptualization, A.L.-B., C.R.G.-C. and J.R.M.-R.; writing—original draft preparation, A.L.-B., C.R.G.-C., M.F.R.-G., C.A.R.-D., M.U.-R. and J.R.M.-R.; writing—review and editing, A.L.-B., C.R.G.-C., M.F.R.-G., C.A.R.-D., M.U.-R. and J.R.M.-R.; supervision, A.L.-B., C.R.G.-C. and J.R.M.-R. All authors have read and agreed to the published version of the manuscript.

Funding

We would like to acknowledge Paicyt 2019–2020 and 2020–2021 Science Grant from the Universidad Autónoma de Nuevo León; CONACyT Grants for Basic Science grant 221332; Fronteras de la Ciencia grant 1502, Infraestructura Grant 279957 and Apoyos a la Ciencia de Frontera grant 316869.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

Angel Leon-Buitimea would like to thank the support from Beca de Posdoctorado Nacional 2018–2020 from CONACyT and Garza-Cárdenas CR would like to thank Beca Nacional de Posgrado 2020–2020 from CONACyT.

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Mulani, M.S.; Kamble, E.E.; Kumkar, S.N.; Tawre, M.S.; Pardesi, K.R. Emerging strategies to combat ESKAPE pathogens in the era of antimicrobial resistance: A review. Front. Microbiol. 2019, 10, 539. [Google Scholar] [CrossRef] [PubMed]
  2. Petchiappan, A.; Chatterji, D. Antibiotic Resistance: Current Perspectives. ACS Omega 2017, 2, 7400–7409. [Google Scholar] [CrossRef] [PubMed]
  3. Brown, E.D.; Wright, G.D. Antibacterial drug discovery in the resistance era. Nature 2016, 529, 336–343. [Google Scholar] [CrossRef]
  4. Aslam, B.; Wang, W.; Arshad, M.I.; Khurshid, M.; Muzammil, S.; Rasool, M.H.; Nisar, M.A.; Alvi, R.F.; Aslam, M.A.; Qamar, M.U.; et al. Antibiotic resistance: A rundown of a global crisis. Infect. Drug Resist. 2018, 11, 1645–1658. [Google Scholar] [CrossRef] [PubMed]
  5. Ventola, C.L. The antibiotic resistance crisis: Part 1: Causes and threats. P T 2015, 40, 277–283. [Google Scholar]
  6. Basak, S.; Singh, P.; Rajurkar, M. Multidrug Resistant and Extensively Drug Resistant Bacteria: A Study. J. Pathog. 2016, 2016, 4065603. [Google Scholar] [CrossRef] [PubMed]
  7. Schaenzer, A.J.; Wright, G.D. Antibiotic Resistance by Enzymatic Modification of Antibiotic Targets. Trends Mol. Med. 2020, 26, 768–782. [Google Scholar] [CrossRef]
  8. Santajit, S.; Indrawattana, N. Mechanisms of Antimicrobial Resistance in ESKAPE Pathogens. BioMed Res. Int. 2016, 2016, 2475067. [Google Scholar] [CrossRef]
  9. Munita, J.M.; Arias, C.A. Mechanisms of Antibiotic Resistance. Microbiol. Spectr. 2016, 4, 10–15. [Google Scholar] [CrossRef]
  10. Reygaert, W.C. An overview of the antimicrobial resistance mechanisms of bacteria. AIMS Microbiol. 2018, 4, 482–501. [Google Scholar] [CrossRef]
  11. Li, X.; Plésiat, P. The challenge of efflux-mediated antibiotic resistance in Gram-negative bacteria. Clin. Microbiol. Rev. 2015, 28, 337–418. [Google Scholar] [CrossRef] [PubMed]
  12. Sauvage, E.; Kerff, F.; Terrak, M.; Ayala, J.A.; Charlier, P. The penicillin-binding proteins: Structure and role in peptidoglycan biosynthesis. FEMS Microbiol. Rev. 2008, 32, 234–258. [Google Scholar] [CrossRef]
  13. Worthington, R.J.; Melander, C. Combination approaches to combat multidrug-resistant bacteria. Trends Biotechnol. 2013, 31, 177–184. [Google Scholar] [CrossRef] [PubMed]
  14. Walsh, C. Molecular mechanisms that confer antibacterial drug resistance. Nature 2000, 406, 775–781. [Google Scholar] [CrossRef] [PubMed]
  15. Vallet-Regí, M.; González, B.; Izquierdo-Barba, I. Nanomaterials as Promising Alternative in the Infection Treatment. Int. J. Mol. Sci. 2019, 20, 3806. [Google Scholar] [CrossRef] [PubMed]
  16. Malmsten, M. Nanomaterials as Antimicrobial Agents. In Handbook of Nanomaterials Properties; Springer: Berlin/Heidelberg, Germany, 2014; pp. 1053–1075. [Google Scholar]
  17. Fischbach, M.A. Combination therapies for combating antimicrobial resistance. Curr. Opin. Microbiol. 2011, 14, 519–523. [Google Scholar] [CrossRef]
  18. Hughes, D.; Karlén, A. Discovery and preclinical development of new antibiotics. Ups. J. Med. Sci. 2014, 119, 162–169. [Google Scholar] [CrossRef]
  19. Ramsay, R.R.; Popovic-Nikolic, M.R.; Nikolic, K.; Uliassi, E.; Bolognesi, M.L. A perspective on multi-target drug discovery and design for complex diseases. Clin. Transl. Med. 2018, 7, 3. [Google Scholar] [CrossRef]
  20. Kamaruzzaman, N.F.; Tan, L.P.; Hamdan, R.H.; Choong, S.S.; Wong, W.K.; Gibson, A.J.; Chivu, A.; Pina, M.d.F. Antimicrobial Polymers: The Potential Replacement of Existing Antibiotics? Int. J. Mol. Sci. 2019, 20, 2747. [Google Scholar] [CrossRef]
  21. Baptista, P.V.; McCusker, M.P.; Carvalho, A.; Ferreira, D.A.; Mohan, N.M.; Martins, M.; Fernandes, A.R. Nano-strategies to fight multidrug resistant bacteria-"A Battle of the Titans". Front. Microbiol. 2018, 9, 1441. [Google Scholar] [CrossRef]
  22. Lewies, A.; Du Plessis, L.H.; Wentzel, J.F. Antimicrobial Peptides: The Achilles’ Heel of Antibiotic Resistance? Probiotics Antimicrob. Proteins 2019, 11, 370–381. [Google Scholar] [CrossRef] [PubMed]
  23. León-Buitimea, A.; Garza-Cárdenas, C.R.; Garza-Cervantes, J.A.; Lerma-Escalera, J.A.; Morones-Ramírez, J.R. The Demand for New Antibiotics: Antimicrobial Peptides, Nanoparticles, and Combinatorial Therapies as Future Strategies in Antibacterial Agent Design. Front. Microbiol. 2020, 11, 1669. [Google Scholar] [CrossRef] [PubMed]
  24. Negut, I.; Bita, B.; Groza, A. Polymeric Coatings and Antimicrobial Peptides as Efficient Systems for Treating Implantable Medical Devices Associated-Infections. Polymers 2022, 14, 1611. [Google Scholar] [CrossRef] [PubMed]
  25. Spirescu, V.A.; Chircov, C.; Grumezescu, A.M.; Andronescu, E. Polymeric Nanoparticles for Antimicrobial Therapies: An up-to-date Overview. Polymers 2021, 13, 724. [Google Scholar] [CrossRef]
  26. Gera, S.; Kankuri, E.; Kogermann, K. Antimicrobial peptides – Unleashing their therapeutic potential using nanotechnology. Pharmacol. Ther. 2022, 232, 107990. [Google Scholar] [CrossRef]
  27. Hemeg, H.A. Nanomaterials for alternative antibacterial therapy. Int. J. Nanomed. 2017, 12, 8211–8225. [Google Scholar] [CrossRef]
  28. Kavčič, B.; Tkačik, G.; Bollenbach, T. Mechanisms of drug interactions between translation-inhibiting antibiotics. Nat. Commun. 2020, 11, 4013. [Google Scholar] [CrossRef]
  29. Yañez-Macías, R.; Muñoz-Bonilla, A.; De Jesús-Tellez, M.A.; Maldonado-Textle, H.; Guerrero-Sánchez, C.; Schubert, U.S.; Guerrero-Santos, R. Combinations of Antimicrobial Polymers with Nanomaterials and Bioactives to Improve Biocidal Therapies. Polymers 2019, 11, 1789. [Google Scholar] [CrossRef]
  30. Sung, Y.K.; Kim, S.W. Recent advances in polymeric drug delivery systems. Biomater. Res. 2020, 24, 12. [Google Scholar] [CrossRef]
  31. Sidhu, A.K.; Verma, N.; Kaushal, P. Role of Biogenic Capping Agents in the Synthesis of Metallic Nanoparticles and Evaluation of Their Therapeutic Potential. Front. Nanotechnol. 2022, 3, 105. [Google Scholar] [CrossRef]
  32. Spirescu, V.A.; Chircov, C.; Grumezescu, A.M.; Vasile, B.Ș.; Andronescu, E. Inorganic Nanoparticles and Composite Films for Antimicrobial Therapies. Int. J. Mol. Sci. 2021, 22, 4595. [Google Scholar] [CrossRef]
  33. Lei, J.; Sun, L.; Huang, S.; Zhu, C.; Li, P.; He, J.; Mackey, V.; Coy, D.H.; He, Q. The antimicrobial peptides and their potential clinical applications. Am. J. Transl. Res. 2019, 11, 3919–3931. [Google Scholar] [PubMed]
  34. Benfield, A.H.; Henriques, S.T. Mode-of-Action of Antimicrobial Peptides: Membrane Disruption vs. Intracellular Mechanisms. Front. Med. Technol. 2020, 2, 20. [Google Scholar] [CrossRef] [PubMed]
  35. Namivandi-Zangeneh, R.; Sadrearhami, Z.; Dutta, D.; Willcox, M.; Wong, E.H.H.; Boyer, C. Synergy between Synthetic Antimicrobial Polymer and Antibiotics: A Promising Platform to Combat Multidrug-Resistant Bacteria. ACS Infect. Dis. 2019, 5, 1357–1365. [Google Scholar] [CrossRef] [PubMed]
  36. Song, X.; You, J.; Xu, C.; Zhu, A.; Yan, C.; Guo, R. Enhanced Anticancer Cells Effects of Optimized Suspension Stable As 2 O 3 -Loaded Poly (lactic-co-glycolic acid) Nanocapsules. Chin. J. Chem. 2015, 7, 777–784. [Google Scholar] [CrossRef]
  37. Reinbold, J.; Hierlemann, T.; Hinkel, H.; Müller, I.; Maier, M.E.; Weindl, T.; Schlensak, C.; Wendel, H.P.; Krajewski, S. Development and in vitro characterization of poly (lactide-co-glycolide) microspheres loaded with an antibacterial natural drug for the treatment of long-term bacterial infections. Drug Des. Dev. Ther. 2016, 10, 2823–2832. [Google Scholar]
  38. Yüksel, E.; Karakeçili, A. Antibacterial activity on electrospun poly(lactide-co-glycolide) based membranes via Magainin II grafting. Mater. Sci. Eng. C 2014, 45, 510–518. [Google Scholar] [CrossRef]
  39. Góra, A.; Prabhakaran, M.P.; Eunice, G.T.L.; Lakshminarayanan, R.; Ramakrishna, S. Silver nanoparticle incorporated poly( l -lactide- co -glycolide) nanofibers: Evaluation of their biocompatibility and antibacterial properties. J. Appl. Polym. Sci. 2015, 132, 42. [Google Scholar] [CrossRef]
  40. Zhu, G.; Lu, B.; Zhang, T.-X.; Zhang, T.; Zhang, C.; Li, Y.; Peng, Q. Antibiofilm effect of drug-free and cationic poly(D,L-lactide-co-glycolide) nanoparticles via nano–bacteria interactions. Nanomedicine 2018, 13, 1093–1106. [Google Scholar] [CrossRef]
  41. Chaubey, N.; Sahoo, A.K.; Chattopadhyay, A.; Ghosh, S.S. Silver nanoparticle loaded PLGA composite nanoparticles for improving therapeutic efficacy of recombinant IFNγ by targeting the cell surface. Biomater. Sci. 2014, 2, 1080–1089. [Google Scholar] [CrossRef]
  42. Mohiti-Asli, M.; Pourdeyhimi, B.; Loboa, E.G. Novel, silver-ion-releasing nanofibrous scaffolds exhibit excellent antibacterial efficacy without the use of silver nanoparticles. Acta Biomater. 2014, 10, 2096–2104. [Google Scholar] [CrossRef] [PubMed]
  43. Siedenbiedel, F.; Tiller, J.C. Antimicrobial Polymers in Solution and on Surfaces: Overview and Functional Principles. Polymer 2012, 4, 46–71. [Google Scholar] [CrossRef]
  44. Masood, F.; Yasin, T.; Bukhari, H.; Mujahid, M. Characterization and application of roxithromycin loaded cyclodextrin based nanoparticles for treatment of multidrug resistant bacteria. Mater. Sci. Eng. C 2016, 61, 1–7. [Google Scholar] [CrossRef]
  45. Qian, Y.; Zhou, X.; Zhang, F.; Diekwisch, T.G.H.; Luan, X.; Yang, J. Triple PLGA / PCL Scaffold Modification Including Silver Impregnation, Collagen Coating, and Electrospinning Significantly Improve Biocompatibility, Antimicrobial, and Osteogenic Properties for Orofacial Tissue Regeneration. ACS Appl. Mater. Interfaces 2019, 11, 37381–37396. [Google Scholar] [CrossRef] [PubMed]
  46. Faria, A.F.D.; Shaulsky, E.; Chavez, L.H.A.; Elimelech, M. Antimicrobial Electrospun Biopolymer Nano fi ber Mats Functionalized with Graphene Oxide—Silver Nanocomposites. ACS Appl. Mater. Interfaces 2015, 23, 12751–12759. [Google Scholar] [CrossRef] [PubMed]
  47. Wu, J.; Li, C.; Tsai, C.; Chou, C. Synthesis of antibacterial TiO 2 / PLGA composite biofilms. Nanomed. Nanotechnol. Biol. Med. 2014, 5, e1097–e1107. [Google Scholar] [CrossRef] [PubMed]
  48. Serrano, C.; García-Fernández, L.; Fernández-Blázquez, J.P.; Barbeck, M.; Ghanaati, S.; Unger, R.; Kirkpatrick, J.; Arzt, E.; Funk, L.; Turón, P.; et al. Nanostructured medical sutures with antibacterial properties. Biomaterials 2015, 52, 291–300. [Google Scholar] [CrossRef]
  49. Umair, M.M.; Jiang, Z.; Safdar, W.; Xie, Z.; Ren, X. N-Halamine-modified polyglycolide ( PGA ) multifilament as a potential bactericidal surgical suture: In vitro study. J. Appl. Polym. Sci. 2015, 42483, 1–9. [Google Scholar] [CrossRef]
  50. Rouhollahi, F.; Hosseini, S.A.; Alihosseini, F. Investigation on the Biodegradability and Antibacterial Properties of Nanohybrid Suture Based on Silver Incorporated PGA-PLGA Nanofibers. Fibers Polym. 2018, 19, 2056–2065. [Google Scholar] [CrossRef]
  51. Gao, L.; Wang, Y.; Li, Y.; Xu, M.; Sun, G. Biomimetic biodegradable Ag @ Au nanoparticle-embedded ureteral stent with a constantly renewable contact-killing antimicrobial surface and antibiofilm and extraction-free properties. Acta Biomater. 2020, 114, 117–132. [Google Scholar] [CrossRef]
  52. Díez-Pascual, A.M.; Díez-Vicente, A.L. Multifunctional poly(glycolic acid-co-propylene fumarate) electrospun fibers reinforced with graphene oxide and hydroxyapatite nanorods. J. Mater. Chem. B 2017, 5, 4084–4096. [Google Scholar] [CrossRef] [PubMed]
  53. Mohamed, N.A.; Abd El-Ghany, N.A. Synthesis, characterization and antimicrobial activity of novel aminosalicylhydrazide cross linked chitosan modified with multi-walled carbon nanotubes. Cellulose 2019, 26, 1141–1156. [Google Scholar] [CrossRef]
  54. Loyola-Rodríguez, J.P.; Torres-Méndez, F.; Espinosa-Cristobal, L.F.; García-Cortes, J.O.; Loyola-Leyva, A.; González, F.J.; Soto-Barreras, U.; Nieto-Aguilar, R.; Contreras-Palma, G. Antimicrobial activity of endodontic sealers and medications containing chitosan and silver nanoparticles against Enterococcus faecalis. J. Appl. Biomater. Funct. Mater. 2019, 17, 228080001985177. [Google Scholar] [CrossRef] [PubMed]
  55. He, L.; Zou, L.; Yang, Q.; Xia, J.; Zhou, K.; Zhu, Y.; Han, X.; Pu, B.; Hu, B.; Deng, W.; et al. Antimicrobial Activities of Nisin, Tea Polyphenols, and Chitosan and their Combinations in Chilled Mutton. J. Food Sci. 2016, 81, M1466–M1471. [Google Scholar] [CrossRef]
  56. Kim, J.W.; Park, J.K. Synergistic antimicrobial properties of active molecular chitosan with EDTA-divalent metal ion compounds. J. Phytopathol. 2017, 165, 641–651. [Google Scholar] [CrossRef]
  57. Dong, F.; Zhang, J.; Yu, C.; Li, Q.; Ren, J.; Wang, G.; Gu, G.; Guo, Z. Synthesis of amphiphilic aminated inulin via ‘click chemistry’ and evaluation for its antibacterial activity. Bioorg. Med. Chem. Lett. 2014, 24, 4590–4593. [Google Scholar] [CrossRef]
  58. Nooshkam, M.; Falah, F.; Zareie, Z.; Tabatabaei Yazdi, F.; Shahidi, F.; Mortazavi, S.A. Antioxidant potential and antimicrobial activity of chitosan–inulin conjugates obtained through the Maillard reaction. Food Sci. Biotechnol. 2019, 28, 1861–1869. [Google Scholar] [CrossRef]
  59. Zhang, G.; Liu, J.; Li, R.; Jiao, S.; Feng, C.; Wang, Z.; Du, Y. Conjugation of Inulin Improves Anti-Biofilm Activity of Chitosan. Mar. Drugs 2018, 16, 151. [Google Scholar] [CrossRef]
  60. Wahbi, W.; Siam, R.; Kegere, J.; El-Mehalmey, W.A.; Mamdouh, W. Novel Inulin Electrospun Composite Nanofibers: Prebiotic and Antibacterial Activities. ACS Omega 2020, 5, 3006–3015. [Google Scholar] [CrossRef]
  61. Zabihollahi, N.; Alizadeh, A.; Almasi, H.; Hanifian, S.; Hamishekar, H. Development and characterization of carboxymethyl cellulose based probiotic nanocomposite film containing cellulose nanofiber and inulin for chicken fillet shelf life extension. Int. J. Biol. Macromol. 2020, 160, 409–417. [Google Scholar] [CrossRef]
  62. Varaprasad, K.; Raghavendra, G.M.; Jayaramudu, T.; Seo, J. Nano zinc oxide–sodium alginate antibacterial cellulose fibres. Carbohydr. Polym. 2016, 135, 349–355. [Google Scholar] [CrossRef] [PubMed]
  63. Madzovska-Malagurski, I.; Vukasinovic-Sekulic, M.; Kostic, D.; Levic, S. Towards antimicrobial yet bioactive Cu-alginate hydrogels. Biomed. Mater. 2016, 11, 035015. [Google Scholar] [CrossRef]
  64. Gholizadeh, B.S.; Buazar, F.; Hosseini, S.M.; Mousavi, S.M. Enhanced antibacterial activity, mechanical and physical properties of alginate/hydroxyapatite bionanocomposite film. Int. J. Biol. Macromol. 2018, 116, 786–792. [Google Scholar] [CrossRef] [PubMed]
  65. Shao, Y.; Wu, C.; Wu, T.; Yuan, C.; Chen, S.; Ding, T.; Ye, X.; Hu, Y. Green synthesis of sodium alginate-silver nanoparticles and their antibacterial activity. Int. J. Biol. Macromol. 2018, 111, 1281–1292. [Google Scholar] [CrossRef] [PubMed]
  66. Frígols, B.; Martí, M.; Hernández-Oliver, C.; Aarstad, O.; Teialeret Ulset, A.-S.; Inger Sætrom, G.; Lillelund Aachmann, F.; Serrano-Aroca, Á. Graphene oxide in zinc alginate films: Antibacterial activity, water sorption, wettability and opacity. Carbohydr. Polym. 2019, 14, e0212819. [Google Scholar]
  67. Blachechen, T.S.; Petri, D.F.S. Physicochemical and antimicrobial properties of in situ crosslinked alginate/hydroxypropyl methylcellulose/ε-polylysine films. J. Appl. Polym. Sci. 2020, 137, 48832. [Google Scholar] [CrossRef]
  68. Marković, D.; Tseng, H.-H.; Nunney, T.; Radoičić, M.; Ilic-Tomic, T.; Radetić, M. Novel antimicrobial nanocomposite based on polypropylene non-woven fabric, biopolymer alginate and copper oxides nanoparticles. Appl. Surf. Sci. 2020, 527, 146829. [Google Scholar] [CrossRef]
  69. Budak, K.; Sogut, O.; Aydemir Sezer, U. A review on synthesis and biomedical applications of polyglycolic acid. J. Polym. Res. 2020, 27, 208. [Google Scholar] [CrossRef]
  70. Mozumder, M.S.; Mairpady, A.; Mourad, A.-H.I. Polymeric nanobiocomposites for biomedical applications. J. Biomed. Mater. Res. Part B Appl. Biomater. 2017, 105, 1241–1259. [Google Scholar] [CrossRef]
  71. Sayed, S.; Jardine, M.A. Antimicrobial Biopolymers. In Advanced Functional Materials; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2015; pp. 493–533. [Google Scholar]
  72. Francis, R. Relevance of Natural Degradable Polymers in the Biomedical Field. Biomed. Appl. Polym. Mater. Compos. 2017, 303–360. [Google Scholar]
  73. Mogoşanu, G.D.; Grumezescu, A.M. Pharmaceutical Natural Polymers: Structure and Chemistry. Handb. Polym. Pharm. Technol. 2015, 1, 477–519. [Google Scholar]
  74. Edwards-jones, V. Antimicrobial assessment of a chitosan microfibre dressing: A natural antimicrobial. J. Wound Care 1859, 11, 716–721. [Google Scholar] [CrossRef] [PubMed]
  75. Martínez-Robles, Á.M.; Loyola-Rodríguez, J.P.; Zavala-Alonso, N.V.; Martinez-Martinez, R.E.; Ruiz, F.; Lara-Castro, R.H.; Donohué-Cornejo, A.; Reyes-López, S.Y.; Espinosa-Cristóbal, L.F. Antimicrobial properties of biofunctionalized silver nanoparticles on clinical isolates of Streptococcus mutans and its serotypes. Nanomaterials 2016, 6, 136. [Google Scholar] [CrossRef] [PubMed]
  76. Chien, R.; Yen, M.; Mau, J. Antimicrobial and antitumor activities of chitosan from shiitake stipes, compared to commercial chitosan from crab shells. Carbohydr. Polym. 2016, 138, 259–264. [Google Scholar] [CrossRef] [PubMed]
  77. Chen, Y.; Shafiq, M.; Liu, M.; Morsi, Y.; Mo, X. Advanced fabrication for electrospun three-dimensional nanofiber aerogels and scaffolds. Bioact. Mater. 2020, 5, 963–979. [Google Scholar] [CrossRef] [PubMed]
  78. Lopes, S.M.S.; Krausov, G.; Gonçalves, R.A.C.; Oliveira, A.J.B. De Isolation and characterization of inulin with a high degree of polymerization from roots of Stevia rebaudiana (Bert) Bertoni. Carbohydr. Res. 2015, 411, 15–21. [Google Scholar] [CrossRef] [PubMed]
  79. Guo, Z.; Li, Q.; Wang, G.; Dong, F.; Zhou, H.; Zhang, J. Synthesis, characterization, and antifungal activity of novel inulin derivatives with chlorinated benzene. Carbohydr. Polym. 2014, 99, 469–473. [Google Scholar] [CrossRef] [PubMed]
  80. Pranckutė, R.; Kaunietis, A.; Kuisienė, N.; Čitavičius, D. Development of Synbiotics with Inulin, Palatinose, α-Cyclodextrin and Probiotic Bacteria. Polish, J. Microbiol. 2014, 63, 33–41. [Google Scholar] [CrossRef]
  81. Venkatesan, J.; Bhatnagar, I.; Manivasagan, P.; Kang, K.-H.; Kim, S.-K. Alginate composites for bone tissue engineering: A review. Int. J. Biol. Macromol. 2015, 72, 269–281. [Google Scholar] [CrossRef]
  82. Smith, J.A.; Mele, E. Electrospinning and Additive Manufacturing: Adding Three-Dimensionality to Electrospun Scaffolds for Tissue Engineering. Front. Bioeng. Biotechnol. 2021, 9, 1238. [Google Scholar] [CrossRef]
  83. Li, H.; Chen, X.; Lu, W.; Wang, J.; Xu, Y.; Guo, Y. Application of Electrospinning in Antibacterial Field. Nanomaterials 2021, 11, 1822. [Google Scholar] [CrossRef] [PubMed]
  84. Maliszewska, I.; Czapka, T. Electrospun Polymer Nanofibers with Antimicrobial Activity. Polymers 2022, 14, 1661. [Google Scholar] [CrossRef] [PubMed]
  85. Podstawczyk, D.; Skrzypczak, D.; Połomska, X.; Stargała, A.; Witek-Krowiak, A.; Guiseppi-Elie, A.; Galewski, Z. Preparation of antimicrobial 3D printing filament: In situ thermal formation of silver nanoparticles during the material extrusion. Polym. Compos. 2020, 41, 4692–4705. [Google Scholar] [CrossRef]
  86. Ligon, S.C.; Liska, R.; Stampfl, J.; Gurr, M.; Mülhaupt, R. Polymers for 3D Printing and Customized Additive Manufacturing. Chem. Rev. 2017, 117, 10212–10290. [Google Scholar] [CrossRef]
  87. Ngo, T.D.; Kashani, A.; Imbalzano, G.; Nguyen, K.T.Q.; Hui, D. Additive manufacturing (3D printing): A review of materials, methods, applications and challenges. Compos. Part B Eng. 2018, 143, 172–196. [Google Scholar] [CrossRef]
  88. Dizaj, S.M.; Lotfipour, F.; Barzegar-Jalali, M.; Zarrintan, M.H.; Adibkia, K. Antimicrobial activity of the metals and metal oxide nanoparticles. Mater. Sci. Eng. C 2014, 44, 278–284. [Google Scholar] [CrossRef]
  89. Kurtjak, M.; Aničić, N.; Vukomanovicć, M. Inorganic Nanoparticles: Innovative Tools for Antimicrobial Agents. In Antibacterial Agents; InTech: Rijeka, Croatia, 2017. [Google Scholar]
  90. Wang, L.; Hu, C.; Shao, L. The antimicrobial activity of nanoparticles: Present situation and prospects for the future. Int. J. Nanomedicine 2017, 12, 1227–1249. [Google Scholar] [CrossRef]
  91. Mechanistic Insights into the Antimicrobial Actions of Metallic Nanoparticles and Their Implications for Multidrug Resistance. Available online: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6566786/ (accessed on 10 April 2020).
  92. Slavin, Y.N.; Asnis, J.; Häfeli, U.O.; Bach, H. Metal nanoparticles: Understanding the mechanisms behind antibacterial activity. J. Nanobiotechnol. 2017, 15, 1–20. [Google Scholar] [CrossRef]
  93. Krychowiak, M.; Kawiak, A.; Narajczyk, M.; Borowik, A.; Królicka, A. Silver Nanoparticles Combined With Naphthoquinones as an Effective Synergistic Strategy Against Staphylococcus aureus. Front. Pharmacol. 2018, 9, 816. [Google Scholar] [CrossRef]
  94. Ma, L.; Kohli, M.; Smith, A. Nanoparticles for combination drug therapy. ACS Nano 2013, 7, 9518–9525. [Google Scholar] [CrossRef]
  95. Palza, H. Antimicrobial Polymers with Metal Nanoparticles. Int. J. Mol. Sci. 2015, 16, 2099–2116. [Google Scholar] [CrossRef] [PubMed]
  96. Quarta, A.; Curcio, A.; Kakwere, H.; Pellegrino, T. Polymer coated inorganic nanoparticles: Tailoring the nanocrystal surface for designing nanoprobes with biological implications. Nanoscale 2012, 4, 3319–3334. [Google Scholar] [CrossRef] [PubMed]
  97. Chen, Y.; Xianyu, Y.; Jiang, X. Surface Modification of Gold Nanoparticles with Small Molecules for Biochemical Analysis. Acc. Chem. Res. 2017, 50, 310–319. [Google Scholar] [CrossRef] [PubMed]
  98. Hood, M.A.; Mari, M.; Muñoz-Espí, R. Synthetic Strategies in the Preparation of Polymer/Inorganic Hybrid Nanoparticles. Materials 2014, 7, 4057–4087. [Google Scholar] [CrossRef]
  99. Zhang, F.; Lees, E.; Amin, F.; Rivera_Gil, P.; Yang, F.; Mulvaney, P.; Parak, W.J. Polymer-Coated Nanoparticles: A Universal Tool for Biolabelling Experiments. Small 2011, 7, 3113–3127. [Google Scholar] [CrossRef]
  100. Kango, S.; Kalia, S.; Celli, A.; Njuguna, J.; Habibi, Y.; Kumar, R. Surface modification of inorganic nanoparticles for development of organic–inorganic nanocomposites—A review. Prog. Polym. Sci. 2013, 38, 1232–1261. [Google Scholar] [CrossRef]
  101. Fam, S.Y.; Chee, C.F.; Yong, C.Y.; Ho, K.L.; Mariatulqabtiah, A.R.; Tan, W.S. Stealth Coating of Nanoparticles in Drug-Delivery Systems. Nanomaterials 2020, 10, 787. [Google Scholar] [CrossRef]
  102. Teixeira, M.C.; Carbone, C.; Sousa, M.C.; Espina, M.; Garcia, M.L.; Sanchez-Lopez, E.; Souto, E.B. Nanomedicines for the Delivery of Antimicrobial Peptides (AMPs). Nanomaterials 2020, 10, 560. [Google Scholar] [CrossRef]
  103. Suárez, L.; Castellano, J.; Díaz, S.; Tcharkhtchi, A.; Ortega, Z. Are Natural-Based Composites Sustainable? Polymers 2021, 13, 2326. [Google Scholar] [CrossRef]
  104. Gour, A.; Jain, N.K. Advances in green synthesis of nanoparticles. Artif. Cells Nanomed. Biotechnol. 2019, 47, 844–851. [Google Scholar] [CrossRef]
  105. Zhang, Y.; Shareena Dasari, T.P.; Deng, H.; Yu, H. Antimicrobial Activity of Gold Nanoparticles and Ionic Gold. J. Environ. Sci. Heal. Part C 2015, 33, 286–327. [Google Scholar] [CrossRef] [PubMed]
  106. Tiwari, P.; Vig, K.; Dennis, V.; Singh, S. Functionalized Gold Nanoparticles and Their Biomedical Applications. Nanomaterials 2011, 1, 31–63. [Google Scholar] [CrossRef] [PubMed]
  107. Zhao, X.; Jia, Y.; Li, J.; Dong, R.; Zhang, J.; Ma, C.; Wang, H.; Rui, Y.; Jiang, X. Indole Derivative-Capped Gold Nanoparticles as an Effective Bactericide in Vivo. ACS Appl. Mater. Interfaces 2018, 10, 29398–29406. [Google Scholar] [CrossRef] [PubMed]
  108. Zheng, Y.; Liu, W.; Qin, Z.; Chen, Y.; Jiang, H.; Wang, X. Mercaptopyrimidine-Conjugated Gold Nanoclusters as Nanoantibiotics for Combating Multidrug-Resistant Superbugs. Bioconjug. Chem. 2018, 29, 3094–3103. [Google Scholar] [CrossRef] [PubMed]
  109. Yang, X.; Wei, Q.; Shao, H.; Jiang, X. Multivalent Aminosaccharide-Based Gold Nanoparticles as Narrow-Spectrum Antibiotics in Vivo. ACS Appl. Mater. Interfaces 2019, 11, 7725–7730. [Google Scholar] [CrossRef] [PubMed]
  110. Payne, J.N.; Waghwani, H.K.; Connor, M.G.; Hamilton, W.; Tockstein, S.; Moolani, H.; Chavda, F.; Badwaik, V.; Lawrenz, M.B.; Dakshinamurthy, R. Novel Synthesis of Kanamycin Conjugated Gold Nanoparticles with Potent Antibacterial Activity. Front. Microbiol. 2016, 7, 607. [Google Scholar] [CrossRef]
  111. Bagga, P.; Siddiqui, H.H.; Akhtar, J.; Mahmood, T.; Khan, M.Z. and M.S. Gold Nanoparticles Conjugated Levofloxacin: For Improved Antibacterial Activity Over Levofloxacin Alone. Curr. Drug Deliv. 2017, 14, 1114–1119. [Google Scholar] [CrossRef]
  112. Shamaila, S.; Zafar, N.; Riaz, S.; Sharif, R.; Nazir, J.; Naseem, S. Gold Nanoparticles: An Efficient Antimicrobial Agent against Enteric Bacterial Human Pathogen. Nanomaterials 2016, 6, 71. [Google Scholar] [CrossRef]
  113. Elahi, N.; Kamali, M.; Baghersad, M.H. Recent biomedical applications of gold nanoparticles: A review. Talanta 2018, 184, 537–556. [Google Scholar] [CrossRef]
  114. Bindhu, M.R.; Umadevi, M. Antibacterial activities of green synthesized gold nanoparticles. Mater. Lett. 2014, 120, 122–125. [Google Scholar] [CrossRef]
  115. Abdel-Raouf, N.; Al-Enazi, N.M.; Ibraheem, I.B.M. Green biosynthesis of gold nanoparticles using Galaxaura elongata and characterization of their antibacterial activity. Arab. J. Chem. 2017, 10, S3029–S3039. [Google Scholar] [CrossRef]
  116. Muthuvel, A.; Adavallan, K.; Balamurugan, K.; Krishnakumar, N. Biosynthesis of gold nanoparticles using Solanum nigrum leaf extract and screening their free radical scavenging and antibacterial properties. Biomed. Prev. Nutr. 2014, 4, 325–332. [Google Scholar] [CrossRef]
  117. Gu, X.; Xu, Z.; Gu, L.; Xu, H.; Han, F.; Chen, B.; Pan, X. Preparation and antibacterial properties of gold nanoparticles: A review. Environ. Chem. Lett. 2021, 19, 167–187. [Google Scholar] [CrossRef]
  118. Feng, Y.; Chen, W.; Jia, Y.; Tian, Y.; Zhao, Y.; Long, F.; Rui, Y.; Jiang, X. N-Heterocyclic molecule-capped gold nanoparticles as effective antibiotics against multi-drug resistant bacteria. Nanoscale 2016, 8, 13223–13227. [Google Scholar] [CrossRef] [PubMed]
  119. Sun, Z.; Zheng, W.; Zhu, G.; Lian, J.; Wang, J.; Hui, P.; He, S.; Chen, W.; Jiang, X. Albumin Broadens the Antibacterial Capabilities of Nonantibiotic Small Molecule-Capped Gold Nanoparticles. ACS Appl. Mater. Interfaces 2019, 11, 45381–45389. [Google Scholar] [CrossRef]
  120. Boroumand Moghaddam, A.; Namvar, F.; Moniri, M.; Tahir, P.M.; Azizi, S.; Mohamad, R. Nanoparticles Biosynthesized by Fungi and Yeast: A Review of Their Preparation, Properties, and Medical Applications. Molecules 2015, 20, 16540–16565. [Google Scholar] [CrossRef]
  121. Abdelghany, T.M.; Al-Rajhi, A.M.H.; Al Abboud, M.A.; Alawlaqi, M.M.; Ganash Magdah, A.; Helmy, E.A.M.; Mabrouk, A.S. Recent Advances in Green Synthesis of Silver Nanoparticles and Their Applications: About Future Directions. A Review. Bionanoscience 2018, 8, 5–16. [Google Scholar] [CrossRef]
  122. Qais, F.A.; Samreen; Ahmad, I. Green Synthesis of Metal Nanoparticles: Characterization and their Antibacterial Efficacy BT-Antibacterial Drug Discovery to Combat MDR: Natural Compounds, Nanotechnology and Novel Synthetic Sources. In Antibacterial Drug Discovery to Combat MDR; Ahmad, I., Ahmad, S., Rumbaugh, K.P., Eds.; Springer Singapore: Singapore, 2019; pp. 635–680. ISBN 978-981-13-9871-1. [Google Scholar]
  123. Ruddaraju, L.K.; Pammi, S.V.N.; Guntuku, G.S.; Padavala, V.S.; Kolapalli, V.R.M. A review on anti-bacterials to combat resistance: From ancient era of plants and metals to present and future perspectives of green nano technological combinations. Asian J. Pharm. Sci. 2020, 15, 42–59. [Google Scholar] [CrossRef]
  124. Asariha, M.; Chahardoli, A.; Karimi, N.; Gholamhosseinpour, M.; Khoshroo, A.; Nemati, H.; Shokoohinia, Y.; Fattahi, A. Green synthesis and structural characterization of gold nanoparticles from Achillea wilhelmsii leaf infusion and in vitro evaluation. Bull. Mater. Sci. 2020, 43, 57. [Google Scholar] [CrossRef]
  125. Brown, A.N.; Smith, K.; Samuels, T.A.; Lu, J.; Obare, S.O.; Scott, M.E. Nanoparticles Functionalized with Ampicillin Destroy Multiple-Antibiotic-Resistant Isolates of Pseudomonas aeruginosa and Enterobacter aerogenes and Methicillin-Resistant Staphylococcus aureus. Appl. Environ. Microbiol. 2012, 78, 2768–2774. [Google Scholar] [CrossRef]
  126. Nishanthi, R.; Malathi, S.; Paul, S.J.; Palani, P. Green synthesis and characterization of bioinspired silver, gold and platinum nanoparticles and evaluation of their synergistic antibacterial activity after combining with different classes of antibiotics. Mater. Sci. Eng. C 2019, 96, 693–707. [Google Scholar] [CrossRef]
  127. Kumari, A.; Yadav, S.K.; Yadav, S.C. Biodegradable polymeric nanoparticles based drug delivery systems. Colloids Surf. B Biointerfaces 2010, 75, 1–18. [Google Scholar] [CrossRef] [PubMed]
  128. Yang, X.; Yang, J.; Wang, L.; Ran, B.; Jia, Y.; Zhang, L.; Yang, G.; Shao, H.; Jiang, X. Pharmaceutical Intermediate-Modified Gold Nanoparticles: Against Multidrug-Resistant Bacteria and Wound-Healing Application via an Electrospun Scaffold. ACS Nano 2017, 11, 5737–5745. [Google Scholar] [CrossRef] [PubMed]
  129. Alavi, M.; Rai, M. Topical delivery of growth factors and metal/metal oxide nanoparticles to infected wounds by polymeric nanoparticles: An overview. Expert Rev. Anti. Infect. Ther. 2020, 18, 1021–1032. [Google Scholar] [CrossRef] [PubMed]
  130. Pradeepa; Vidya, S.M.; Mutalik, S.; Udaya Bhat, K.; Huilgol, P.; Avadhani, K. Preparation of gold nanoparticles by novel bacterial exopolysaccharide for antibiotic delivery. Life Sci. 2016, 153, 171–179. [Google Scholar] [CrossRef] [PubMed]
  131. Li, Y.; Tian, Y.; Zheng, W.; Feng, Y.; Huang, R.; Shao, J.; Tang, R.; Wang, P.; Jia, Y.; Zhang, J.; et al. Composites of Bacterial Cellulose and Small Molecule-Decorated Gold Nanoparticles for Treating Gram-Negative Bacteria-Infected Wounds. Small 2017, 13, 1700130. [Google Scholar] [CrossRef]
  132. Yang, P.; Pageni, P.; Rahman, M.A.; Bam, M.; Zhu, T.; Chen, Y.P.; Nagarkatti, M.; Decho, A.W.; Tang, C. Gold Nanoparticles with Antibiotic-Metallopolymers toward Broad-Spectrum Antibacterial Effects. Adv. Healthc. Mater. 2019, 8, 1800854. [Google Scholar] [CrossRef]
  133. Lin, P.; Wang, F.-Q.; Li, C.-T.; Yan, Z.-F. An Enhancement of Antibacterial Activity and Synergistic Effect of Biosynthesized Silver Nanoparticles by Eurotium cristatum with Various Antibiotics. Biotechnol. Bioprocess Eng. 2020, 25, 450–458. [Google Scholar] [CrossRef]
  134. Lopez-Carrizales, M.; Velasco, K.I.; Castillo, C.; Flores, A.; Magaña, M.; Martinez-Castanon, G.A.; Martinez-Gutierrez, F. In vitro synergism of silver nanoparticles with antibiotics as an alternative treatment in multiresistant uropathogens. Antibiotics 2018, 7, 50. [Google Scholar] [CrossRef]
  135. Gomaa, E.Z. Synergistic Antibacterial Efficiency of Bacteriocin and Silver Nanoparticles Produced by Probiotic Lactobacillus paracasei Against Multidrug Resistant Bacteria. Int. J. Pept. Res. Ther. 2019, 25, 1113–1125. [Google Scholar] [CrossRef]
  136. Abdallah, O.M.; EL-Baghdady, K.Z.; Khalil, M.M.H.; El Borhamy, M.I.; Meligi, G.A. Antibacterial, antibiofilm and cytotoxic activities of biogenic polyvinyl alcohol-silver and chitosan-silver nanocomposites. J. Polym. Res. 2020, 27, 74. [Google Scholar] [CrossRef]
  137. Bhande, R.M.; Khobragade, C.N.; Mane, R.S.; Bhande, S. Enhanced synergism of antibiotics with zinc oxide nanoparticles against extended spectrum β-lactamase producers implicated in urinary tract infections. J. Nanoparticle Res. 2013, 15, 1413. [Google Scholar] [CrossRef]
  138. Ghasemi, F.; Jalal, R. Antimicrobial action of zinc oxide nanoparticles in combination with ciprofloxacin and ceftazidime against multidrug-resistant Acinetobacter baumannii. J. Glob. Antimicrob. Resist. 2016, 6, 118–122. [Google Scholar] [CrossRef] [PubMed]
  139. Tyagi, P.K.; Gola, D.; Tyagi, S.; Mishra, A.K.; Kumar, A.; Chauhan, N.; Ahuja, A.; Sirohi, S. Synthesis of zinc oxide nanoparticles and its conjugation with antibiotic: Antibacterial and morphological characterization. Environ. Nanotechnol. Monit. Manag. 2020, 14, 100391. [Google Scholar] [CrossRef]
  140. Fadwa, A.O.; Alkoblan, D.K.; Mateen, A.; Albarag, A.M. Synergistic effects of zinc oxide nanoparticles and various antibiotics combination against Pseudomonas aeruginosa clinically isolated bacterial strains. Saudi J. Biol. Sci. 2020, 28, 928–935. [Google Scholar] [CrossRef]
  141. Limayem, A.; Micciche, A.; Haller, E.; Zhang, C.; Mohapatra, S. Nanotherapeutics for mutating multi-drug resistant fecal bacteria. J. NanotecNanosci 2015, 1, 100–106. [Google Scholar]
  142. Mehta, M.; Allen-Gipson, D.; Mohapatra, S.; Kindy, M.; Limayem, A. Study on the therapeutic index and synergistic effect of Chitosan-zinc oxide nanomicellar composites for drug-resistant bacterial biofilm inhibition. Int. J. Pharm. 2019, 565, 472–480. [Google Scholar] [CrossRef]
  143. Garza-Cervantes, J.A.; Escárcega-González, C.E.; Barriga Castro, E.D.; Mendiola-Garza, G.; Marichal-Cancino, B.A.; López-Vázquez, M.A.; Morones-Ramirez, J.R. Antimicrobial and antibiofilm activity of biopolymer-Ni, Zn nanoparticle biocomposites synthesized using R. mucilaginosa UANL-001L exopolysaccharide as a capping agent. Int. J. Nanomed. 2019, 14, 2557–2571. [Google Scholar] [CrossRef]
  144. Vazquez-Rodriguez, A.; Vasto-Anzaldo, X.G.; Barboza Perez, D.; Vázquez-Garza, E.; Chapoy-Villanueva, H.; García-Rivas, G.; Garza-Cervantes, J.A.; Gómez-Lugo, J.J.; Gomez-Loredo, A.E.; Garza Gonzalez, M.T.; et al. Microbial Competition of Rhodotorula mucilaginosa UANL-001L and E. Coli increase biosynthesis of Non-Toxic Exopolysaccharide with Applications as a Wide-Spectrum Antimicrobial. Sci. Rep. 2018, 8, 1–14. [Google Scholar] [CrossRef]
  145. Li, S.; Wu, C.; Zhao, X.; Jiang, H.; Yan, H.; Wang, X. Synergistic antibacterial activity of new isomeric carborane derivatives through combination with nanoscaled titania. J. Biomed. Nanotechnol. 2013, 9, 393–402. [Google Scholar] [CrossRef]
  146. Masoumi, S.; Shakibaie, M.R.; Gholamrezazadeh, M.; Monirzadeh, F. Evaluation Synergistic Effect of TiO2, ZnO Nanoparticles and Amphiphilic Peptides (Mastoparan-B, Indolicidin) Against Drug-Resistant Pseudomonas aeruginosa, Klebsiella pneumoniae and Acinetobacter baumannii. Arch. Pediatr. Infect. Dis. 2018, 6, e57920. [Google Scholar] [CrossRef]
  147. Hérault, N.; Wagner, J.; Abram, S.-L.; Widmer, J.; Horvath, L.; Vanhecke, D.; Bourquin, C.; Fromm, K.M. Silver-Containing Titanium Dioxide Nanocapsules for Combating Multidrug-Resistant Bacteria. Int. J. Nanomed. 2020, 15, 1267–1281. [Google Scholar] [CrossRef] [PubMed]
  148. Zhang, S.; Liang, X.; Gadd, G.M.; Zhao, Q. Advanced titanium dioxide-polytetrafluorethylene (TiO2-PTFE) nanocomposite coatings on stainless steel surfaces with antibacterial and anti-corrosion properties. Appl. Surf. Sci. 2019, 490, 231–241. [Google Scholar] [CrossRef]
  149. Harun, N.H.; Mydin, R.B.S.M.N.; Sreekantan, S.; Saharudin, K.A.; Basiron, N.; Aris, F.; Wan Mohd Zain, W.N.; Seeni, A. Bactericidal Capacity of a Heterogeneous TiO2/ZnO Nanocomposite against Multidrug-Resistant and Non-Multidrug-Resistant Bacterial Strains Associated with Nosocomial Infections. ACS Omega 2020, 5, 12027–12034. [Google Scholar] [CrossRef]
  150. Ahmed, F.Y.; Farghaly Aly, U.; Abd El-Baky, R.M.; Waly, N.G.F.M. Comparative Study of Antibacterial Effects of Titanium Dioxide Nanoparticles Alone and in Combination with Antibiotics on MDR Pseudomonas aeruginosa Strains. Int. J. Nanomed. 2020, 15, 3393–3404. [Google Scholar] [CrossRef]
  151. Ullah, K.; Khan, S.A.; Mannan, A.; Khan, R.; Murtaza, G.; Yameen, M.A. Enhancing the Antibacterial Activity of Erythromycin with Titanium Dioxide Nanoparticles against MRSA. Curr. Pharm. Biotechnol. 2020, 21, 948–954. [Google Scholar] [CrossRef]
  152. Ivashchenko, O.; Lewandowski, M.; Peplińska, B.; Jarek, M.; Nowaczyk, G.; Wiesner, M.; Załęski, K.; Babutina, T.; Warowicka, A.; Jurga, S. Synthesis and characterization of magnetite/silver/antibiotic nanocomposites for targeted antimicrobial therapy. Mater. Sci. Eng. C 2015, 55, 343–359. [Google Scholar] [CrossRef]
  153. Ivashchenko, O.; Woźniak, A.; Coy, E.; Peplinska, B.; Gapinski, J.; Jurga, S. Release and cytotoxicity studies of magnetite/Ag/antibiotic nanoparticles: An interdependent relationship. Colloids Surf. B Biointerfaces 2017, 152, 85–94. [Google Scholar] [CrossRef]
  154. Ivashchenko, O.; Jurga-Stopa, J.; Coy, E.; Peplinska, B.; Pietralik, Z.; Jurga, S. Fourier transform infrared and Raman spectroscopy studies on magnetite/Ag/antibiotic nanocomposites. Appl. Surf. Sci. 2016, 364, 400–409. [Google Scholar] [CrossRef]
  155. Ficai, D.; Grumezescu, V.; Mariana, O.; Popescu, R.C.; Holban, A.M.; Ficai, A.; Grumezescu, A.M.; Mogoanta, L.; Dan, G.; Id, M.; et al. Antibiofilm Coatings Based on PLGA and Nanostructured Cefepime-Functionalized Magnetite. Nanomaterials 2018, 9, 633. [Google Scholar] [CrossRef]
  156. Grumezescu, V.; Maria, A.; Iordache, F.; Socol, G.; Mogos, G.D.; Mihai, A.; Ficai, A. Applied Surface Science MAPLE fabricated magnetite @ eugenol and (3-hidroxybutyric acid-co-3-hidroxyvaleric acid)– polyvinyl alcohol microspheres coated surfaces with anti-microbial properties. Appl. Surf. Sci. 2014, 306, 16–22. [Google Scholar] [CrossRef]
  157. Haldorai, Y.; Kharismadewi, D.; Tuma, D.; Shim, J.-J. Properties of chitosan/magnetite nanoparticles composites for efficient dye adsorption and antibacterial agent. Korean J. Chem. Eng. 2015, 32, 1688–1693. [Google Scholar] [CrossRef]
  158. Niemirowicz, K.; Piktel, E.; Wilczewska, A.; Markiewicz, K.; Durnaś, B.; Wątek, M.; Puszkarz, I.; Wróblewska, M.; Niklińska, W.; Savage, P.; et al. Core–shell magnetic nanoparticles display synergistic antibacterial effects against Pseudomonas aeruginosa and Staphylococcus aureus when combined with cathelicidin LL-37 or selected ceragenins. Int. J. Nanomed. 2016, 11, 5443–5455. [Google Scholar] [CrossRef] [PubMed]
  159. Barillo, D.J.; Marx, D.E. Silver in medicine: A brief history BC 335 to present. Burns 2014, 40, S3–S8. [Google Scholar] [CrossRef]
  160. Politano, A.D.; Campbell, K.T.; Rosenberger, L.H.; Sawyer, R.G. Use of silver in the prevention and treatment of infections: Silver review. Surg. Infect. (Larchmt) 2013, 14, 8–20. [Google Scholar] [CrossRef]
  161. Syafiuddin, A.; Salmiati; Salim, M.R.; Beng Hong Kueh, A.; Hadibarata, T.; Nur, H. A Review of Silver Nanoparticles: Research Trends, Global Consumption, Synthesis, Properties, and Future Challenges. J. Chin. Chem. Soc. 2017, 64, 732–756. [Google Scholar] [CrossRef]
  162. Yaqoob, A.A.; Umar, K.; Ibrahim, M.N.M. Silver nanoparticles: Various methods of synthesis, size affecting factors and their potential applications–a review. Appl. Nanosci. 2020, 10, 1369–1378. [Google Scholar] [CrossRef]
  163. Gaillet, S.; Rouanet, J.-M. Silver nanoparticles: Their potential toxic effects after oral exposure and underlying mechanisms—A review. Food Chem. Toxicol. an Int. J. Publ. Br. Ind. Biol. Res. Assoc. 2015, 77, 58–63. [Google Scholar] [CrossRef]
  164. Akter, M.; Sikder, M.T.; Rahman, M.M.; Ullah, A.K.M.A.; Hossain, K.F.B.; Banik, S.; Hosokawa, T.; Saito, T.; Kurasaki, M. A systematic review on silver nanoparticles-induced cytotoxicity: Physicochemical properties and perspectives. J. Adv. Res. 2018, 9, 1–16. [Google Scholar] [CrossRef]
  165. Liao, C.; Li, Y.; Tjong, S.C. Bactericidal and Cytotoxic Properties of Silver Nanoparticles. Int. J. Mol. Sci. 2019, 20, 499. [Google Scholar] [CrossRef]
  166. Singh, R.; Wagh, P.; Wadhwani, S.; Gaidhani, S.; Kumbhar, A.; Bellare, J.; Chopade, B.A. Synthesis, optimization, and characterization of silver nanoparticles from Acinetobacter calcoaceticus and their enhanced antibacterial activity when combined with antibiotics. Int. J. Nanomed. 2013, 8, 4277. [Google Scholar] [CrossRef]
  167. Shahverdi, A.R.; Fakhimi, A.; Shahverdi, H.R.; Minaian, S. Synthesis and effect of silver nanoparticles on the antibacterial activity of different antibiotics against Staphylococcus aureus and Escherichia coli. Nanomed. Nanotechnol. Biol. Med. 2007, 3, 168–171. [Google Scholar] [CrossRef] [PubMed]
  168. Naqvi, S.Z.H.; Kiran, U.; Ali, M.I.; Jamal, A.; Hameed, A.; Ahmed, S.; Ali, N. Combined efficacy of biologically synthesized silver nanoparticles and different antibiotics against multidrug-resistant bacteria. Int. J. Nanomed. 2013, 8, 3187–3195. [Google Scholar] [CrossRef] [PubMed]
  169. Vazquez-Muñoz, R.; Meza-Villezcas, A.; Fournier, P.G.J.; Soria-Castro, E.; Juarez-Moreno, K.; Gallego-Hernández, A.L.; Bogdanchikova, N.; Vazquez-Duhalt, R.; Huerta-Saquero, A. Enhancement of antibiotics antimicrobial activity due to the silver nanoparticles impact on the cell membrane. PLoS One 2019, 14, e0224904. [Google Scholar] [CrossRef] [PubMed]
  170. Pandit, R.; Rai, M.; Santos, C.A. Enhanced antimicrobial activity of the food-protecting nisin peptide by bioconjugation with silver nanoparticles. Environ. Chem. Lett. 2017, 15, 443–452. [Google Scholar] [CrossRef]
  171. Sidhu, P.K.; Nehra, K. Bacteriocin-capped silver nanoparticles for enhanced antimicrobial efficacy against food pathogens. IET Nanobiotechnol. 2020, 14, 245–252. [Google Scholar] [CrossRef]
  172. Ansari, A.; Pervez, S.; Javed, U.; Abro, M.I.; Nawaz, M.A.; Qader, S.A.U.; Aman, A. Characterization and interplay of bacteriocin and exopolysaccharide-mediated silver nanoparticles as an antibacterial agent. Int. J. Biol. Macromol. 2018, 115, 643–650. [Google Scholar] [CrossRef]
  173. Fahim, H.A.; Khairalla, A.S.; El-Gendy, A.O. Nanotechnology: A Valuable Strategy to Improve Bacteriocin Formulations. Front. Microbiol. 2016, 7, 1385. [Google Scholar] [CrossRef]
  174. Sanyasi, S.; Majhi, R.K.; Kumar, S.; Mishra, M.; Ghosh, A.; Suar, M.; Satyam, P.V.; Mohapatra, H.; Goswami, C.; Goswami, L. Polysaccharide-capped silver Nanoparticles inhibit biofilm formation and eliminate multi-drug-resistant bacteria by disrupting bacterial cytoskeleton with reduced cytotoxicity towards mammalian cells. Sci. Rep. 2016, 6, 24929. [Google Scholar] [CrossRef]
  175. Alavi, M.; Rai, M. Recent progress in nanoformulations of silver nanoparticles with cellulose, chitosan, and alginic acid biopolymers for antibacterial applications. Appl. Microbiol. Biotechnol. 2019, 103, 8669–8676. [Google Scholar] [CrossRef]
  176. Kaur, A.; Kumar, R. Enhanced bactericidal efficacy of polymer stabilized silver nanoparticles in conjugation with different classes of antibiotics. RSC Adv. 2019, 9, 1095–1105. [Google Scholar] [CrossRef] [PubMed]
  177. Lyu, Y.; Yu, M.; Liu, Q.; Zhang, Q.; Liu, Z.; Tian, Y.; Li, D.; Changdao, M. Synthesis of silver nanoparticles using oxidized amylose and combination with curcumin for enhanced antibacterial activity. Carbohydr. Polym. 2020, 230, 115573. [Google Scholar] [CrossRef]
  178. Figueiredo, E.; Ribeiro, J.; Nishio, E.; Scandorieiro, S.; Costa, A.; Cardozo, V.; de Oliveira, A.; Durán, N.; Panagio, L.; Kobayashi, R.; et al. New Approach For Simvastatin As An Antibacterial: Synergistic Effect With Bio-Synthesized Silver Nanoparticles Against Multidrug-Resistant Bacteria. Int. J. Nanomed. 2019, 14, 7975–7985. [Google Scholar] [CrossRef] [PubMed]
  179. dos Santos Courrol, D.; Lopes, C.R.B.; Pereira, C.B.P.; Franzolin, M.R.; de Oliveira Silva, F.R.; Courrol, L.C. Tryptophan Silver Nanoparticles Synthesized by Photoreduction Method: Characterization and Determination of Bactericidal and Anti-Biofilm Activities on Resistant and Susceptible Bacteria. Int. J. Tryptophan Res. 2019, 12, 1178646919831677. [Google Scholar] [CrossRef] [PubMed]
  180. Panthi, G.; Hassan, M.M.; Kuk, Y.-S.; Kim, J.Y.; Chung, H.-J.; Hong, S.-T.; Park, M. Enhanced Antibacterial Property of Sulfate-Doped Ag3PO4 Nanoparticles Supported on PAN Electrospun Nanofibers. Molecules 2020, 25, 1411. [Google Scholar] [CrossRef] [PubMed]
  181. Piccirillo, C.; Pinto, R.A.; Tobaldi, D.M.; Pullar, R.C.; Labrincha, J.A.; Pintado, M.M.E.; Castro, P.M.L. Light induced antibacterial activity and photocatalytic properties of Ag/Ag3PO4 -based material of marine origin. J. Photochem. Photobiol. A Chem. 2015, 296, 40–47. [Google Scholar] [CrossRef]
  182. Zhang, C.; Wang, J.; Chi, R.; Shi, J.; Yang, Y.; Zhang, X. Reduced graphene oxide loaded with MoS2 and Ag3PO4 nanoparticles/PVA interpenetrating hydrogels for improved mechanical and antibacterial properties. Mater. Des. 2019, 183, 108166. [Google Scholar] [CrossRef]
  183. Steckiewicz, K.P.; Zwara, J.; Jaskiewicz, M.; Kowalski, S.; Kamysz, W.; Zaleska-Medynska, A.; Inkielewicz-Stepniak, I. Shape-Depended Biological Properties of Ag3PO4 Microparticles: Evaluation of Antimicrobial Properties and Cytotoxicity in In Vitro Model—Safety Assessment of Potential Clinical Usage. Oxid. Med. Cell. Longev. 2019, 2019, 6740325. [Google Scholar] [CrossRef]
  184. Xie, Y.; He, Y.; Irwin, P.L.; Jin, T.; Shi, X. Antibacterial activity and mechanism of action of zinc oxide nanoparticles against Campylobacter jejuni. Appl. Environ. Microbiol. 2011, 77, 2325–2331. [Google Scholar] [CrossRef]
  185. Sabir, S.; Arshad, M.; Chaudhari, S.K. Zinc Oxide Nanoparticles for Revolutionizing Agriculture: Synthesis and Applications. Sci. World J. 2014, 2014, 925494. [Google Scholar] [CrossRef]
  186. Espitia, P.J.P.; Otoni, C.G.; Soares, N.F.F. Chapter 34-Zinc Oxide Nanoparticles for Food Packaging Applications. In; Barros-Velázquez, J.B.T.-A.F.P., Ed.; Academic Press: San Diego, CA, USA, 2016; pp. 425–431. ISBN 978-0-12-800723-5. [Google Scholar]
  187. Sirelkhatim, A.; Mahmud, S.; Seeni, A.; Kaus, N.H.M.; Ann, L.C.; Bakhori, S.K.M.; Hasan, H.; Mohamad, D. Review on Zinc Oxide Nanoparticles: Antibacterial Activity and Toxicity Mechanism. Nano-Micro Lett. 2015, 7, 219–242. [Google Scholar] [CrossRef] [PubMed]
  188. Kadiyala, U.; Kotov, N.A.; VanEpps, J.S. Antibacterial Metal Oxide Nanoparticles: Challenges in Interpreting the Literature. Curr. Pharm. Des. 2018, 24, 896–903. [Google Scholar] [CrossRef] [PubMed]
  189. Vandebriel, R.J.; De Jong, W.H. A review of mammalian toxicity of ZnO nanoparticles. Nanotechnol. Sci. Appl. 2012, 5, 61–71. [Google Scholar] [CrossRef] [PubMed]
  190. Perelshtein, I.; Ruderman, E.; Perkas, N.; Tzanov, T.; Beddow, J.; Joyce, E.; Mason, T.J.; Blanes, M.; Mollá, K.; Patlolla, A.; et al. Chitosan and chitosan–ZnO-based complex nanoparticles: Formation, characterization, and antibacterial activity. J. Mater. Chem. B 2013, 1, 1968–1976. [Google Scholar] [CrossRef]
  191. Sathiya, S.M.; Okram, G.S.; Maria Dhivya, S.; Manivannan, G.; Jothi Rajan, M.A. Interaction of Chitosan/Zinc Oxide Nanocomposites and their Antibacterial Activities with Escherichia coli. Mater. Today Proc. 2016, 3, 3855–3860. [Google Scholar] [CrossRef]
  192. Baek, S.; Joo, S.H.; Toborek, M. Treatment of antibiotic-resistant bacteria by encapsulation of ZnO nanoparticles in an alginate biopolymer: Insights into treatment mechanisms. J. Hazard. Mater. 2019, 373, 122–130. [Google Scholar] [CrossRef]
  193. Darbasizadeh, B.; Fatahi, Y.; Feyzi-barnaji, B.; Arabi, M.; Motasadizadeh, H.; Farhadnejad, H.; Moraffah, F.; Rabiee, N. Crosslinked-polyvinyl alcohol-carboxymethyl cellulose/ZnO nanocomposite fibrous mats containing erythromycin (PVA-CMC/ZnO-EM): Fabrication, characterization and in-vitro release and anti-bacterial properties. Int. J. Biol. Macromol. 2019, 141, 1137–1146. [Google Scholar] [CrossRef]
  194. Weir, A.; Westerhoff, P.; Fabricius, L.; Hristovski, K.; von Goetz, N. Titanium Dioxide Nanoparticles in Food and Personal Care Products. Environ. Sci. Technol. 2012, 46, 2242–2250. [Google Scholar] [CrossRef]
  195. Shi, H.; Magaye, R.; Castranova, V.; Zhao, J. Titanium dioxide nanoparticles: A review of current toxicological data. Part. Fibre Toxicol. 2013, 10, 15. [Google Scholar] [CrossRef]
  196. López de Dicastillo, C.; Guerrero Correa, M.; Martínez, F.B.; Streitt, C.; José Galotto, M. Antimicrobial Effect of Titanium Dioxide Nanoparticles. In Antimicrobial Resistance [Working Title]; IntechOpen: London, UK, 2020. [Google Scholar]
  197. Baranowska-Wójcik, E.; Szwajgier, D.; Oleszczuk, P.; Winiarska-Mieczan, A. Effects of Titanium Dioxide Nanoparticles Exposure on Human Health—A Review. Biol. Trace Elem. Res. 2020, 193, 118–129. [Google Scholar] [CrossRef]
  198. Ling, C.; An, H.; Li, L.; Wang, J.; Lu, T.; Wang, H.; Hu, Y.; Song, G.; Liu, S. Genotoxicity Evaluation of Titanium Dioxide Nanoparticles In Vitro: A Systematic Review of the Literature and Meta-analysis. Biol. Trace Elem. Res. 2020, 199, 2057–2076. [Google Scholar] [CrossRef]
  199. Wani, M.R.; Shadab, G. Titanium dioxide nanoparticle genotoxicity: A review of recent in vivo and in vitro studies. Toxicol. Ind. Health 2020, 36, 514–530. [Google Scholar] [CrossRef] [PubMed]
  200. Sophee, S.S.; Prasad, R.G.S.V.; Srinivas, J.V.; Aparna, R.S.L.; Phani, A.R. Antibacterial Activity of TiO2 and ZnO Microparticles Combination on Water Polluting Bacteria. J. Green Sci. Technol. 2013, 1, 20–26. [Google Scholar] [CrossRef]
  201. Soo, J.Z.; Chai, L.C.; Ang, B.C.; Ong, B.H. Enhancing the Antibacterial Performance of Titanium Dioxide Nanofibers by Coating with Silver Nanoparticles. ACS Appl. Nano Mater. 2020, 3, 5743–5751. [Google Scholar] [CrossRef]
  202. Li, W.; Li, B.; Meng, M.; Cui, Y.; Wu, Y.; Zhang, Y.; Dong, H.; Feng, Y. Bimetallic Au/Ag decorated TiO2 nanocomposite membrane for enhanced photocatalytic degradation of tetracycline and bactericidal efficiency. Appl. Surf. Sci. 2019, 487, 1008–1017. [Google Scholar] [CrossRef]
  203. Harun, N.H.; Mydin, R.B.S.M.N.; Sreekantan, S.; Saharudin, K.A.; Basiron, N.; Seeni, A. The bactericidal potential of LLDPE with TiO2/ZnO nanocomposites against multidrug resistant pathogens associated with hospital acquired infections. J. Biomater. Sci. Polym. Ed. 2020, 31, 1757–1769. [Google Scholar] [CrossRef]
  204. Saharudin, K.A.; Sreekantan, S.; Basiron, N.; Khor, Y.L.; Harun, N.H.; Mydin, R.B.S.M.N.; Md Akil, H.; Seeni, A.; Vignesh, K. Bacteriostatic Activity of LLDPE Nanocomposite Embedded with Sol–Gel Synthesized TiO2/ZnO Coupled Oxides at Various Ratios. Polymers 2018, 10, 878. [Google Scholar] [CrossRef] [PubMed]
  205. Roy, A.S.; Parveen, A.; Koppalkar, A.R.; Prasad, M.V.N.A. Effect of nano-titanium dioxide with different antibiotics against methicillin-resistant Staphylococcus aureus. J. Biomater. Nanobiotechnol. 2010, 1, 37. [Google Scholar] [CrossRef]
  206. Arora, B.; Murar, M.; Dhumale, V. Antimicrobial potential of TiO2 nanoparticles against MDR Pseudomonas aeruginosa. J. Exp. Nanosci. 2015, 10, 819–827. [Google Scholar] [CrossRef]
  207. Niemirowicz, K.; Markiewicz, K.; Wilczewska, A.; Car, H. Magnetic nanoparticles as new diagnostic tools in medicine. Adv. Med. Sci. 2012, 57, 196–207. [Google Scholar] [CrossRef]
  208. Zhang, C.; Du, C.; Liao, J.-Y.; Gu, Y.; Gong, Y.; Pei, J.; Gu, H.; Yin, D.; Gao, L.; Pan, Y. Synthesis of magnetite hybrid nanocomplexes to eliminate bacteria and enhance biofilm disruption. Biomater. Sci. 2019, 7, 2833–2840. [Google Scholar] [CrossRef] [PubMed]
  209. Fox, J.L. Antimicrobial peptides stage a comeback. Nat. Biotechnol. 2013, 31, 379–382. [Google Scholar] [CrossRef] [PubMed]
  210. Hancock, R.E.W.; Sahl, H.G. Antimicrobial and host-defense peptides as new anti-infective therapeutic strategies. Nat. Biotechnol. 2006, 24, 1551–1557. [Google Scholar] [CrossRef] [PubMed]
  211. Gordon, Y.J.; Romanowski, E.G.; McDermott, A.M. Mini review: A review of antimicrobial peptides and their therapeutic potential as anti-infective drugs. Curr. Eye Res. 2005, 30, 505–515. [Google Scholar] [CrossRef] [PubMed]
  212. Jalian, H.R.; Kim, J. Antimicrobial Peptides. In Clinical and Basic Immunodermatology; Springer London: London, UK, 2008; pp. 131–145. ISBN 9781848001640. [Google Scholar]
  213. Chen, C.H.; Lu, T.K. Development and Challenges of Antimicrobial Peptides for Therapeutic Applications. Antibiotics 2020, 9, 24. [Google Scholar] [CrossRef] [PubMed]
  214. Van Epps, H.L. René dubos: Unearthing antibiotics. J. Exp. Med. 2006, 203, 259. [Google Scholar] [CrossRef]
  215. Li, J.; Nation, R.L.; Turnidge, J.D.; Milne, R.W.; Coulthard, K.; Rayner, C.R.; Paterson, D.L. Colistin: The re-emerging antibiotic for multidrug-resistant Gram-negative bacterial infections. Lancet Infect. Dis. 2006, 6, 589–601. [Google Scholar] [CrossRef]
  216. Pizzolato-Cezar, L.R.; Okuda-Shinagawa, N.M.; Machini, M.T. Combinatory Therapy Antimicrobial Peptide-Antibiotic to Minimize the Ongoing Rise of Resistance. Front. Microbiol. 2019, 10, 1703. [Google Scholar] [CrossRef]
  217. Dabirian, S.; Taslimi, Y.; Zahedifard, F.; Gholami, E.; Doustdari, F.; Motamedirad, M.; Khatami, S.; Azadmanesh, K.; Nylen, S.; Rafati, S. Human Neutrophil Peptide-1 (HNP-1): A New Anti-Leishmanial Drug Candidate. PLoS Negl. Trop. Dis. 2013, 7, e2491. [Google Scholar] [CrossRef]
  218. Kalita, A.; Verma, I.; Khuller, G.K. Role of Human Neutrophil Peptide–1 as a Possible Adjunct to Antituberculosis Chemotherapy. J. Infect. Dis. 2004, 190, 1476–1480. [Google Scholar] [CrossRef]
  219. Dürr, U.H.N.; Sudheendra, U.S.; Ramamoorthy, A. LL-37, the only human member of the cathelicidin family of antimicrobial peptides. Biochim. Biophys. Acta Biomembr. 2006, 1758, 1408–1425. [Google Scholar] [CrossRef] [PubMed]
  220. Leszczyńska, K.; Namiot, A.; Janmey, P.A.; Bucki, R. Modulation of exogenous antibiotic activity by host cathelicidin LL-37. APMIS 2010, 118, 830–836. [Google Scholar] [CrossRef] [PubMed]
  221. Choi, H.; Lee, D.G. Synergistic effect of antimicrobial peptide arenicin-1 in combination with antibiotics against pathogenic bacteria. Res. Microbiol. 2012, 163, 479–486. [Google Scholar] [CrossRef] [PubMed]
  222. Rudilla, H.; Fusté, E.; Cajal, Y.; Rabanal, F.; Vinuesa, T.; Viñas, M. Synergistic antipseudomonal effects of synthetic peptide AMP38 and carbapenems. Molecules 2016, 21, 1223. [Google Scholar] [CrossRef] [PubMed]
  223. Wu, X.; Li, Z.; Li, X.; Tian, Y.; Fan, Y.; Yu, C.; Zhou, B.; Liu, Y.; Xiang, R.; Yang, L. Synergistic effects of antimicrobial peptide DP7 combined with antibiotics against multidrug-resistant bacteria. Drug Des. Devel. Ther. 2017, 11, 939–946. [Google Scholar] [CrossRef]
  224. Choi, H.; Lee, D.G. Antimicrobial peptide pleurocidin synergizes with antibiotics through hydroxyl radical formation and membrane damage, and exerts antibiofilm activity. Biochim. Biophys. Acta Gen. Subj. 2012, 1820, 1831–1838. [Google Scholar] [CrossRef]
  225. Lainson, J.C.; Daly, S.M.; Triplett, K.; Johnston, S.A.; Hall, P.R.; Diehnelt, C.W. Synthetic Antibacterial Peptide Exhibits Synergy with Oxacillin against MRSA. ACS Med. Chem. Lett. 2017, 8, 853–857. [Google Scholar] [CrossRef]
  226. Pletzer, D.; Mansour, S.C.; Hancock, R.E.W. Synergy between conventional antibiotics and anti-biofilm peptides in a murine, sub-cutaneous abscess model caused by recalcitrant ESKAPE pathogens. PLoS Pathog. 2018, 14, e1007084. [Google Scholar] [CrossRef]
  227. Bueno, J.; Demirci, F.; Baser, K.H.C. Antimicrobial Strategies in Novel Drug Delivery Systems. In The Microbiology of Skin, Soft Tissue, Bone and Joint Infections; Elsevier: Amsterdam, The Netherlands, 2017; pp. 271–286. [Google Scholar]
  228. Fadaka, A.O.; Sibuyi, N.R.S.; Madiehe, A.M.; Meyer, M. Nanotechnology-Based Delivery Systems for Antimicrobial Peptides. Pharmaceutics 2021, 13, 1795. [Google Scholar] [CrossRef]
  229. Feng, Q.; Huang, Y.; Chen, M.; Li, G.; Chen, Y. Functional synergy of α-helical antimicrobial peptides and traditional antibiotics against Gram-negative and Gram-positive bacteria in vitro and in vivo. Eur. J. Clin. Microbiol. Infect. Dis. 2014, 34, 197–204. [Google Scholar] [CrossRef]
  230. Bolatchiev, A.; Baturin, V.; Bazikov, I.; Maltsev, A.; Kunitsina, E. Effect of antimicrobial peptides HNP-1 and hBD-1 on Staphylococcus aureus strains in vitro and in vivo. Fundam. Clin. Pharmacol. 2020, 34, 102–108. [Google Scholar] [CrossRef] [PubMed]
  231. Sain, S.; Reshmi, J.; Anandaraj, S.; George, S.; Issac, J.S.; John, S. Lesion Sterilization and Tissue Repair-Current Concepts and Practices. Int. J. Clin. Pediatr. Dent. 2018, 11, 446–450. [Google Scholar] [CrossRef] [PubMed]
  232. Al-Haddad, A.; Che Ab Aziz, Z.A. Bioceramic-Based Root Canal Sealers: A Review. Int. J. Biomater. 2016, 2016, 9753210. [Google Scholar] [CrossRef] [PubMed]
  233. Wang, Z. Bioceramic materials in endodontics. Endod. Top. 2015, 32, 3–30. [Google Scholar] [CrossRef]
  234. Sicca, C.; Bobbio, E.; Quartuccio, N.; Nicolò, G.; Cistaro, A. Prevention of dental caries: A review of effective treatments. J. Clin. Exp. Dent. 2016, 8, e604–e610. [Google Scholar] [CrossRef]
  235. Dijksteel, G.S.; Ulrich, M.M.W.; Middelkoop, E.; Boekema, B.K.H.L. Review: Lessons Learned From Clinical Trials Using Antimicrobial Peptides (AMPs). Front. Microbiol. 2021, 12, 287. [Google Scholar] [CrossRef]
  236. Ugurlu, T.; Turkoglu, M.; Gurer, U.S.; Akarsu, B.G. Colonic delivery of compression coated nisin tablets using pectin/HPMC polymer mixture. Eur. J. Pharm. Biopharm. 2007, 67, 202–210. [Google Scholar] [CrossRef]
  237. Zhu, C.; Zhao, J.; Kempe, K.; Wilson, P.; Wang, J.; Velkov, T.; Li, J.; Davis, T.P.; Whittaker, M.R.; Haddleton, D.M. A Hydrogel-Based Localized Release of Colistin for Antimicrobial Treatment of Burn Wound Infection. Macromol. Biosci. 2017, 17, 1600320. [Google Scholar] [CrossRef]
Figure 1. Combinatorial therapies used in the treatment of antibiotic-resistant bacteria.
Figure 1. Combinatorial therapies used in the treatment of antibiotic-resistant bacteria.
Antibiotics 11 00794 g001
Figure 2. Schematic representation of antimicrobial mechanisms of inorganic nanoparticles (INP). (a) alteration of the bacterial cell wall and membrane, (b) induction of oxidative stress due to excessive intracellular production of reactive oxygen species (ROS), (c) inhibition of crucial proteins/enzymes and DNA, (d) disruption of metabolic pathways, and (e) intracellular accumulation of metal ions released from INP.
Figure 2. Schematic representation of antimicrobial mechanisms of inorganic nanoparticles (INP). (a) alteration of the bacterial cell wall and membrane, (b) induction of oxidative stress due to excessive intracellular production of reactive oxygen species (ROS), (c) inhibition of crucial proteins/enzymes and DNA, (d) disruption of metabolic pathways, and (e) intracellular accumulation of metal ions released from INP.
Antibiotics 11 00794 g002
Table 4. Antibacterial activity of nanomaterial-based combinational treatments using in vivo models.
Table 4. Antibacterial activity of nanomaterial-based combinational treatments using in vivo models.
NanomaterialCombined withIn Vivo ModelObservationsReferences
Polymers
Poly (lactide-co-glycolide)Polycaprolactone
Type I collagen
AgNPs
Mouse periodontitis modelNovel silver-modified/collagen-coated PLGA/PCL scaffold features biocompatible, osteogenic, and antibacterial properties.[45]
Poly (glycolic acid)PLGA
AuAg core/shell NPs
Farm pigs with stents implantedThe stent exhibited remarkable antibiofilm property and reduced the level of inflammatory and necrotic cells.
The stent maintained structural integrity without the presence of large fragments in the urinary system
[51]
Nanoparticles
Gold nanoparticles (AuNPs)Bacterial cellulose
(BC) 4,6-diamino-
2-pyrimidinotiol (DAPT)
Rat Wound Infection ModelThe BC-Au-DAPT nanocomposites applied as wound dressings showed excellent anti-MDR bacteria activity and high biocompatibility. [131]
Silver nanoparticles
(AgNPs)
Bacteriocin (BC) extracted from
Lactobacillus paracasei
A. salina modelBC/AgNPs bioconjugate was compatible to the biological system.[135]
Zinc oxide nanoparticles
(ZnO NPs)
EPS from Rhodotorula mucilaginosa UANL- 001L Wistar rat renal model EPS-capped ZnO NPs showed no toxic effect in vivo[143,144]
Titanium dioxide nanoparticles
(TiO2 NPs)
Silver ions Female mice (Charles Rivers) macrophages and dendritic cells modelDespite uptake into macrophages, no proinflammatory response nor cytotoxicity in these cells were detected for our nanocapsules[147]
Antimicrobial peptides (AMPs)
PL-5LevofloxacinMouse wound infection modelThe synergistic application of PL-5 and levofloxacin inhibited bacteria, with a bacteriostatic rate of 99.9%[229]
HNP-1Silica nanoparticles Rats wound infection modelGels containing HNP-1 and showed a significantly faster wound healing in comparison with control.[230]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

León-Buitimea, A.; Garza-Cárdenas, C.R.; Román-García, M.F.; Ramírez-Díaz, C.A.; Ulloa-Ramírez, M.; Morones-Ramírez, J.R. Nanomaterials-Based Combinatorial Therapy as a Strategy to Combat Antibiotic Resistance. Antibiotics 2022, 11, 794. https://doi.org/10.3390/antibiotics11060794

AMA Style

León-Buitimea A, Garza-Cárdenas CR, Román-García MF, Ramírez-Díaz CA, Ulloa-Ramírez M, Morones-Ramírez JR. Nanomaterials-Based Combinatorial Therapy as a Strategy to Combat Antibiotic Resistance. Antibiotics. 2022; 11(6):794. https://doi.org/10.3390/antibiotics11060794

Chicago/Turabian Style

León-Buitimea, Angel, Cesar R. Garza-Cárdenas, María Fernanda Román-García, César Agustín Ramírez-Díaz, Martha Ulloa-Ramírez, and José Rubén Morones-Ramírez. 2022. "Nanomaterials-Based Combinatorial Therapy as a Strategy to Combat Antibiotic Resistance" Antibiotics 11, no. 6: 794. https://doi.org/10.3390/antibiotics11060794

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