Next Article in Journal
Systems Drug Design for Muscle Invasive Bladder Cancer and Advanced Bladder Cancer by Genome-Wide Microarray Data and Deep Learning Method with Drug Design Specifications
Next Article in Special Issue
Insights into the Biocompatibility and Biological Potential of a Chitosan Nanoencapsulated Textile Dye
Previous Article in Journal
Label-Free Surface-Enhanced Raman Spectroscopic Analysis of Proteins: Advances and Applications
Previous Article in Special Issue
Antibacterial Soy Protein Isolate Prepared by Quaternization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nanoparticles for Antimicrobial Agents Delivery—An Up-to-Date Review

by
Doina-Antonia Mercan
1,
Adelina-Gabriela Niculescu
1 and
Alexandru Mihai Grumezescu
1,2,3,*
1
Department of Science and Engineering of Oxide Materials and Nanomaterials, Politehnica University of Bucharest, 011061 Bucharest, Romania
2
Research Institute of the University of Bucharest—ICUB, University of Bucharest, 050657 Bucharest, Romania
3
Academy of Romanian Scientists, Ilfov No. 3, 050044 Bucharest, Romania
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(22), 13862; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232213862
Submission received: 24 October 2022 / Revised: 6 November 2022 / Accepted: 8 November 2022 / Published: 10 November 2022
(This article belongs to the Special Issue Antimicrobial Biomaterials: Recent Progress)

Abstract

:
Infectious diseases constitute an increasing threat to public health and medical systems worldwide. Particularly, the emergence of multidrug-resistant pathogens has left the pharmaceutical arsenal unarmed to fight against such severe microbial infections. Thus, the context has called for a paradigm shift in managing bacterial, fungal, viral, and parasitic infections, leading to the collision of medicine with nanotechnology. As a result, renewed research interest has been noted in utilizing various nanoparticles as drug delivery vehicles, aiming to overcome the limitations of current treatment options. In more detail, numerous studies have loaded natural and synthetic antimicrobial agents into different inorganic, lipid, and polymeric-based nanomaterials and tested them against clinically relevant pathogens. In this respect, this paper reviews the most recently reported successfully fabricated nanoformulations that demonstrated a great potential against bacteria, fungi, viruses, and parasites of interest for human medicine.

1. Introduction

Even though humans and pathogens have always dynamically interacted, this relationship became unbalanced. Human activities have caused pathogenic microbes, such as bacteria, fungi, viruses, and parasites, to appear and spread at a progressively distressing rate, rendering infectious diseases a common and burdensome health issue worldwide [1,2,3,4]. Numerous antimicrobial agents can be employed to fight against infections, yet they often face several limitations [5,6,7]. Specifically, the efficacy of conventional antimicrobial drugs is affected by their poor oral bioavailability and stability, low water solubility, low transportation rate across cellular membranes, lack of targeting, and systemic adverse effects [8,9,10,11]. Another important drawback of traditional drug-based therapeutic strategies is the inappropriate and inadequate administration of antimicrobial agents that have contributed to the emergence of drug-resistant pathogens and the formation of well-organized microbial communities called biofilms [12,13,14,15,16].
Unfortunately, biofilms have become a highly frequent problem in the clinical environment as microorganisms may adhere to and colonize the surfaces of biomedical devices. Thus, in order to avoid the acquiring of nosocomial infections, the burden of biofilms must be especially considered when using implantable and indwelling medical devices (e.g., catheters, stents, heart valves, pacemakers, prosthetic joints and implants, voice prostheses, contact lenses, internal and external fixation devices) [14,16,17,18,19,20,21,22,23]. Generally, the management of device-associated infections involves prolonged inpatient stay, surgical intervention, and long-term postoperative antibiotic therapy, all adding to healthcare costs and low patient compliance [10,13,24].
Despite existing therapies and medicines, infectious diseases in general, and biofilms in particular, remain difficult-to-eradicate problems [3]. Therefore, research efforts must be put into developing antimicrobial strategies able to surpass current challenges. In this context, the advances in nanotechnology represent promising opportunities for designing novel antimicrobial systems. A variety of nanoparticles (NPs) can be employed in developing performant delivery vehicles for natural and synthetic medicines capable of enhancing the activity of carried freight, ensuring a sustained drug release, and offering a chance of biofilm penetration and internalization into pathogenic microorganisms [8,25,26]. NPs loaded with antimicrobial agents can be used as effective therapeutics administered on different routes, but they can also be incorporated further into biomaterials for modifying surface nanotopography or coating biomedical devices, intending to potentiate or induce anti-infective properties [9,11,27,28,29,30,31].
In this respect, the present paper aims to overview the most recently developed antimicrobial nanoformulations, fabricated mainly between 2018 and 2022, that showed promising results when tested against clinically relevant pathogens, emphasizing their utility and versatility. Even though the topic has been previously addressed by several works [32,33,34,35,36], this review proposes a comprehensive path, correlating materials, fabrication methods, and delivered antimicrobial agents with targeted microorganisms, focusing on the applicability of drug delivery systems in treating and preventing bacterial, fungal, viral, and parasitic infections, and updating the literature with the newest developments in the field.

2. Nanoparticles for Antimicrobial Applications

NPs represent a key component in developing innovative anti-infectious strategies merging therapeutics and new materials toward enhancing antimicrobial potential. Their advantageous intrinsic properties, such as the high specific surface area in relation to volume and increased particle surface energy, render these materials more reactive and effective than their bulk counterparts [37,38,39]. Moreover, their small size is very suitable for antimicrobial biological operations, allowing NPs to interact with biological systems at the molecular level, permitting the targeted delivery of drugs and genes, and ensuring passage through biological barriers [39,40,41,42,43,44].
Moreover, the efficacy of NPs as antimicrobial agents’ carriers can be improved for specific goals (e.g., increased cellular uptake, selective recognition, non-cytotoxicity, better payload binding capacity) through various surface-functionalization approaches. Specifically, stimuli-responsive ligands or functional groups can be used for modifying the surface layer of NPs through different physical, chemical, or biological methods toward achieving optimal antimicrobial activity [31]. In this respect, two main stimulation approaches have been actively researched for delivery nanosystems: locally stimulated or externally stimulated (Figure 1). The first category assumes cargo release in response to chemical and biochemical stimuli at intracellular (e.g., enzymatic activities, hydrolysis, pH, etc.) or tissue level (i.e., specific microenvironmental changes associated with pathological conditions). In contrast, the second category of delivery vehicles supposes the activated targeting and sustained release under the influence of external stimuli, including magnetic fields, electric stimulation, ultrasound, light, and temperature [11,45].
Taking into account the inherent beneficial properties of NPs and the variety of surface engineering possibilities, numerous studies have developed a wide range of antimicrobial nanoformulations aiming to bring effective solutions against relevant infections (Figure 2).
In this respect, the following subsections emphasize the recent progress in the fabrication of nanosystems for the delivery of antibacterial, antifungal, antiviral, and antiparasitic agents.

2.1. Brief Overview of Nanoparticles Synthesis Methods

When discussing nanomaterials fabrication, two main approaches can be distinguished: top-down and bottom-up. The top-down approach implies starting from larger structures and reducing their size by means of mechanical force and the aid of finer and finer tools until reaching dimensions in the nano range [46,47]. These methods are preferred in industrial settings, as they can be easily scaled-up and produce fine particles with fine particle-producing capacity and reproducibility. Nonetheless, expensive equipment and intensive energy are required in such processes without guaranteeing control over particle growth and products’ purity [46,48]. In opposition to top-down techniques, bottom-up processes assume the fabrication of nanoparticles through the growth and self-assembly of smaller components of atomic or molecular dimensions, conforming to a natural physical principle or an externally applied driving force [47]. Such methods are simple, rapid, energy-efficient, and cost-effective, being ideal options for laboratory-scale production of amorphous particles with reduced dimensions, narrow particle size distribution, increased solubility, and enhanced bioavailability. However, NPs obtained in this manner tend to agglomerate and might also present stability issues, while there are also several drawbacks associated with the fabrication processes (e.g., low yield, interbatch variability, scaling up challenges) [46,48].
A variety of physical, chemical, and biological fabrication methods are available for the synthesis of nanostructures (Table 1), including co-precipitation, hydrothermal synthesis, inert gas condensation, sputtering, microemulsion, microwave-assisted, laser ablation, sol-gel, ultrasound, spark discharge, template synthesis, and biological synthesis [39,46]. Depending on the chemical nature of the NPs, desired properties of the final product, and the cost-effectiveness of fabrication steps, one may prefer one method over the others. Some of the most employed techniques for the fabrication of magnetic NPs, polymeric NPs, and lipid-based NPs, have been gathered in Figure 3, Figure 4 and Figure 5, respectively, offering a visual perspective on nanoconstructs’ synthesis.
Moreover, for producing NPs with tailored structures, mixed approaches can be employed, generally involving a preprocessing step followed by a high-energy step. As each method has its advantages and limitations, thoughtful consideration is required when choosing the synthesis method so that the final nanostructures would have physicochemical stability, low polydispersity, reproducible size, high purity, and optimum morphology for antimicrobial drug delivery purposes [48].

2.2. Antibacterial Nanoformulations

Bacterial infections represent one of the biggest global health problems, remaining a significant cause of morbidity and mortality despite the numerous available antibiotics [54]. This is primarily due to the appearance of multi-drug resistant bacterial strains that cannot be effectively treated with conventional therapeutics [40,55,56,57]. Improper prescription of antimicrobial drugs and overuse and/or misuse of antibiotics has led to the current antimicrobial resistance growing crisis, enhancing microbial virulence and allowing bacteria to evade the host’s immune response under the protection of a biofilm [40,56,58,59,60,61].
Therefore, a different approach had to be taken to enhance the antibacterial properties of existing drugs. Nanomaterials have become an attractive solution for transporting and releasing hydrophilic and lipophilic antibiotics or natural antimicrobial agents, as they can overpower bacterial resistance through several mechanisms. In more detail, NPs can ensure targeted delivery, allow passage through biological barriers, permeate and destroy the bacterial cell membrane, induce antimicrobial effects within cells, and impede biofilm formation [8,41,62,63,64,65,66,67].
Inorganic nanomaterials such as metal and metal oxide NPs have been of particular interest in creating antibacterial nanoformulations given their advantageous properties (e.g., low cost, long duration, safety, intrinsic antimicrobial activity) [59,68]. These materials act upon bacterial cells mainly through metal ions release, further increasing reactive oxygen species production and affecting bacterial metabolism. Nonetheless, repeated exposure can cause developing resistance even against these NPs. Thus, their antimicrobial efficacy is often enhanced through surface functionalization [69].
Some of the most commonly employed inorganic NPs for antibacterial applications are based on silver, iron oxide, zinc oxide, titanium oxide, magnesium oxide, and silica [69,70,71]. Drugs can be loaded into these systems either as coatings/shells on the NP surface [72], or they can be incorporated into the pores of the material [73]. Various synthetic antimicrobial agents have been used as cargos, including streptomycin [72], neomycin [72], vancomycin [74], cephalexin [75], and ciprofloxacin [76], leading to stronger effects against numerous pathogens. Specifically, antibiotic-loaded inorganic NPs have been tested with promising results against relevant bacterial strains, counting Staphylococcus aureus (S. aureus) [72,73,75,76], Pseudomonas aeruginosa (P. aeruginosa) [72], Bacillus subtilis (B. subtilis) [74], Bacillus cereus (B. cereus) [75], Streptococcus [74], Escherichia coli (E. coli) [73,74,75], and Salmonella typhimurium (S. typhimurium) [75].
Moreover, the nanosystems’ antimicrobial properties and safety can be enhanced by adding biocompatible coatings [72,75,77]. In addition, surface functionalization of NPs can be performed to bypass triggering host defense mechanisms until reaching the site of infection and avoid potential adverse reactions or inhibition of NPs bioactivity [66,78,79,80].
Polymers represent another class of highly convenient materials for fabricating antimicrobial drug-delivery NPs. The main advantages of polymeric materials reside in their variety, versatility, and ease of functionalization. These characteristics render polymers suitable for improving drug solubility, delivering the cargo to the desired site, and targeting bacterial pathogens [50,67,81].
Recent studies have focused on developing vehicles from natural polymers, as they are generally recognized to possess superior biocompatibility to synthetic materials. The most commonly employed natural polymers for antibacterial agents encapsulation are polysaccharides, chitosan and alginate being the choice of numerous studies [55,82,83,84,85,86,87,88,89,90,91,92,93]. Nonetheless, synthetic polymers such as polylactic acid (PLA) [94,95], poly(lactic-co-glycolic acid (PLGA) [96,97], and polyvinylpyrrolidone (PVP) [98] have also attracted research interest as nanocarriers.
In what concerns the freight, a broad range of antimicrobial agents has been reported in the literature as suitable for polymer encapsulation. Synthetic drugs (e.g., levofloxacin [82], gentamicin [83,92], N′-((5-nitrofuran-2-yl)methylen)-2-benzhydrazide [86], rifampicin [55,94], ascorbic acid [55], doxycycline [88], rifaximin [89], ampicillin [91], teicoplanin [98], camptothecin [99], and vancomycin [100]) and natural antimicrobials (e.g., oregano oil [90,101], Cinnamomum zeylanicum (C. zeylanicum) essential oil [84], nettle essential oil [85], Pistacia lentiscus (P. lentiscus) L. var. chia essential oil [95], and red propolis extract [96]) have been successfully incorporated into nanosized polymeric materials.
The synergic properties of engineered polymeric NPs and carried antibacterial moieties have led to the obtaining of promising candidates for anti-infective therapeutics against S. aureus [55,82,83,85,86,87,90,91,92,94,96,97,99,100], E. coli [83,84,85,89,90,92,95,99], P. aeruginosa [82,89,90,92,93,96], Erwinia carotovora (E. carotovora) [84], Pseudomonas fluorescens (P. fluorescens) [84], Enterococcus faecalis (E. faecalis) [88,90], Proteus mirabilis (P. mirabilis) [88], Bacillus haynesii (B. haynesii) [89], Streptococcus pyogenes (S. pyogenes) [90], Yersinia enterocolitica (Y. enerocolitica) [90], Listeria monocytogenes (L. monocytogenes) [92], B. subtilis [95], Streptococcus pneumoniae (S. pneumoniae) [98], Haemophilus influenzae (H. influenzae) [98], and Klebsiella pneumonia (K. pneumoniae) [99].
To emphasize the variety and versatility of recently developed antibacterial nanoformulations, Table 2 summarizes several studies that fabricated promising drug delivery nanosystems for fighting against clinically relevant strains. In addition, Figure 6 provides a visual perspective over some of the discussed nanoconstructs.

2.3. Antifungal Nanoformulations

Fungal infections represent a significant health issue, being associated with high morbidity and mortality. Immunocompromised hosts are particularly susceptible to invasive infections, with the mortality rates in such patients going above 60% in certain situations [3,102,103]. The conventional approach in treating such infections assumes the administration of antifungal agents such as polyenes, azoles, and echinocandins [104]. Nevertheless, these drugs present a series of disadvantages that impede their therapeutic action. Conventional antifungals exhibit non-neglectable toxicity, adverse side effects, acquired resistance, and unclear effects in immunocompromised patients [105,106,107,108].
Therefore, recent research tried to solve these issues by orienting to safer strategies, including encapsulation into biocompatible NPs and replacement with natural alternatives. In this regard, scientists have explored a plethora of nanomaterials for designing performant delivery systems for antifungal agents. Promising results have been reported when using metal and metal oxide-based NPs [109,110], natural polymers [90,111,112], biocompatible synthetic polymers [113,114,115,116,117,118], and lipid-based nanocarriers [119,120]. Regarding the choice of antimicrobial agents, most studies elaborated nanoconstructs for the delivery of synthetic drugs (e.g., nystatin [109], fluconazole [109], amphotericin B [110,117], voriconazole [111], itraconazole [115,116], ketoconazole [118], miconazole nitrate [119], clotrimazole [112]), but several natural antifungal agents (e.g., seedless Vitis vinifera (V. vinifera) [112], oregano oil [90], Lippia sidoides (L. sidoides) essential oil [120], pterostilbene [113], farnesol [114]) have also been investigated and led to promising results.
In what concerns the species of interest, most studies have focused on developing antifungal nanoformulations targeting Candida albicans (C. albicans) [90,109,110,111,112,114,115,117,118,119], as candidiasis is among the most common invasive mycotic diseases, and C. albicans is recognized as the leading cause of invasive candidiasis [121,122]. Nonetheless, other fungal pathogens have also been considered, including Aspergillus brasiliensis (A. brasiliens) [109,113], Aspergilus niger (A. niger) [112], Cryptococcus neoformans (C. neoformans) [110], Histoplasma capsulatum (H. capsulatum) [116], Trichophyton rubrum (T. rubrum) [118], Trichophyton mentagrophytes (T. mentagrophytes) [118], Microsporum gypseum (M. gypseum) [118], Candida dubliniensis (C. dubliniensis) [118], Candida krusei (C. krusei) [118], Candida parapsilosis (C. parapsilosis) [118], Candida tropicalis (C. tropicalis) [118], and Candida auris (C. auris) [120].
For clarity, Table 3 correlates NP material, physicochemical properties of the delivery nanosystems, carried antifungal agents, and targeted pathogens, while Figure 7 schematically illustrates a few of these nanostructures.

2.4. Antiviral Nanoformulations

Viruses are another class of dangerous pathogens, as they are responsible for around two million deaths per year [123]. Their small size allows viruses to enter the human body through various routes and internalize into living cells. Some of the most pathogenic viruses include human immunodeficiency virus (HIV), human papillomavirus (HPV), herpes simplex virus (HSV), norovirus, hepatitis viruses, and coronaviruses, leading to significant morbidity and mortality. Moreover, the occurrence of viral outbreaks was seen to have devastating effects from economic and social points of view [3,124].
In the fight against infections, NPs offer certain advantages for delivering antivirals to the target sites as they have the ability to surpass biological barriers thanks to their small size and tailored surface characteristics. Through their unique properties, NPs allow antivirals to be released at the infection site, followed by their attachment to viral receptors on the surface of host cells or internalization within the cell resulting in the disruption of the viral replication cycle [124].
Taking into account the benefits of NPs, several research studies explored the antimicrobial potential of a number of antiviral drugs encapsulated in different nanomaterials. Examples of investigated nanocarriers include silver NPs [125], titanium dioxide NPs [126], oligo- and polysaccharide-based NPs [127,128,129,130,131], solid lipid NPs (SLNs) [132,133], and large unilamellar vesicles [134]. The ingenious association with antivirals (e.g., docetaxel [125], flavonoids [126], zidovudine [127,128], dolutegravir sodium [129], efavirenz [130], acyclovir [131,132], ritonavir [133]) and functionalization agents has led to the obtaining of promising anti-infective therapeutic nanoformulations. Specifically, the proposed delivery systems targeted clinically relevant viruses, such as severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) [125,126], HIV [127,128,129,130,133,134,135], and HSV [131,132].
To better clarify the features of the newly developed antiviral drug delivery systems, Table 4 and Figure 8 summarize several examples of studies in the field.

2.5. Antiparasitic Nanoformulations

Aside from bacteria, fungi, and viruses, several parasites have also been recognized for their infective potential. One example is represented by Leishmania spp., which comprises a group of flagellated protozoans responsible for neglected tropical diseases known as leishmaniasis. Characterized by high mortality, disability, and morbidity rates, leishmaniasis represents a major global health concern, being endemic in 102 countries worldwide [136,137,138]. Consequently, scientific interest arose in finding antiparasitic solutions able to effectively and efficiently fight against Leishmania.
For instance, Badirzadeh et al. [137] proposed coating silver NPs with curcumin. In vitro and in vivo tests performed on mouse models demonstrated encouraging results, the nanoformulation significantly reducing the burden of promastigotes and amastigotes of Leishmania parasite in a single treatment. Alternatively, Snoussi et al. [136] prepared silver-loaded biochar that exhibited strong antiparasitic activity against the promastigotes stage of Leishmania donovani, Leishmania amazonensis, and epimastigotes of Trypanosoma cruzi. On a different note, Durak and colleagues [139] encapsulated two active ingredients with antibacterial and antiparasitic activities (i.e., caffeic acid phenethyl ester and juglone) into single polymeric NPs. These multifunctional nanoformulations proved synergistic activity, being promising candidates for antiparasitic therapy.
Several studies have also directed their efforts toward creating delivery nanosystems aimed at other pathogens. For example, Kanwal et al. [140] reported the fabrication of silver NPs conjugated with novel bisindole and thiazole derivatives as potential antiamoebic formulations with enhanced activity against Balamuthia mandrillaris and Naegleria fowleri. In contrast, Real et al. [141] developed a drug delivery system for treating fascioliasis. For this purpose, the researchers loaded triclabendazole into nanocapsules, enhancing its bioavailability and lowering its cytotoxic effects compared to the free drug. Differently, Wei et al. [142] created a nanocarrier for decoquinate, a drug known to have control effects on hematogeneous parasites. The authors encapsulated the antiparasitic agent into disodium glycyrrhizinate NPs with protamine and anionic hyaluronic acid layers (Figure 9), significantly increasing the drug’s bioavailability, ensuring a higher concentration in the blood and preferential liver tissue accumulation.

3. Discussion

Humans are exposed to numerous pathogens that can trigger burdensome bacterial, fungal, viral, and parasitic infections. Conventional treatment approaches revolve around the systemic administration of synthetic drugs that, due to the emergence of drug-resistant microbial strains, exhibit low efficacy, in addition to the disadvantages of poor solubility, toxicity, and adverse effects. In this context, nanotechnology started being increasingly explored for designing improved antimicrobial agents
NPs of many sorts (Figure 10) have been recently developed as performant carriers of numerous antimicrobial agents, holding promise for improved strategies to combat a wide range of infectious diseases. Nanodimensional materials, such as metallic NPs, metal-oxide NPs, lipid-based NPs, and polymeric NPs, have attracted considerable interest in recent years for fabricating delivery vehicles. Specifically, the variety and versatility of nanomaterials have been extensively explored by researchers for creating innovative therapeutic formulations that can be administrated on different routes, including oral [2,141,142], ocular [82,83,143], intranasal [125,129,130], intratracheal [55,144], intravaginal [145,146], intravesical [147,148], and transdermal [149,150] routes. Compared to free drugs, NP-loaded antimicrobial agents can be administered in so many ways due to their increased safety, reduced systemic adverse effects, enhanced solubility, and improved bioavailability. Moreover, the various natural and synthetic antimicrobial cargos can be released in a targeted manner by adding various functionalization agents onto the nanocarriers’ surface. Functionalization agents can also work in synergy with the core delivery system, increasing therapeutic efficacy and reducing drug resistance [11,31].
In addition to their stand-alone utility, NPs may further be incorporated in different other materials to create bionanocomposites with enhanced antimicrobial properties. In this respect, researchers propose the use of various nanostructured gels [82,83,98,100,151], patches [149,150,152], wound dressings [153,154,155,156], and scaffolds [157,158,159] as alternative solutions for treating and preventing microbial infections. A particularly exploited application of NPs is the fabrication of coatings for creating surfaces with antimicrobial and antibiofilm properties [12,19,27,29,77,160,161,162,163,164]. Even though aimed mainly at the modification of biomedical devices, such as catheters, implants, and prostheses, applying antimicrobial coatings can also be of high utility in covering other contact surfaces. For instance, they can be used to prevent pathogens from spreading from day-to-day objects, including doorknobs, packaging, and handrails [37].

4. Conclusions and Future Perspectives

To summarize, various NPs have been investigated as drug delivery vehicles to surpass traditional drugs’ limitations. Numerous studies have successfully loaded natural and synthetic drugs into inorganic, lipid, and polymeric-based nanosystems, obtaining promising results against a broad range of pathogens, but mostly bacterial strains.
To conclude, there is an increased research interest in developing alternative antimicrobial agents, and current progress demonstrates the great potential of nanostructured materials in preventing and treating infectious diseases. Nonetheless, there is still room for improvement in the field, especially concerning the expansion of antiviral, antifungal, and antiparasitic applications of drug delivery nanosystems. Further studies should also focus on managing complex and mixed biofilms, an understudied and challenging niche of microbial infections. Moreover, being so new, most of the discussed nano-therapeutic options have not yet advanced beyond preclinical testing. Thus, rigorous additional studies are required before they become clinically and commercially available solutions.

Author Contributions

D.-A.M., A.-G.N., and A.M.G. participated in reviewing, writing, and revision. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Antabe, R.; Ziegler, B.R. Diseases, emerging and infectious. In International Encyclopedia of Human Geography; Springer: Berlin/Heidelberg, Germany, 2020; pp. 389–391. [Google Scholar]
  2. Raza, A.; Sime, F.B.; Cabot, P.J.; Maqbool, F.; Roberts, J.A.; Falconer, J.R. Solid nanoparticles for oral antimicrobial drug delivery: A review. Drug Discov. Today 2019, 24, 858–866. [Google Scholar] [CrossRef]
  3. Sharmin, S.; Rahaman, M.M.; Sarkar, C.; Atolani, O.; Islam, M.T.; Adeyemi, O.S. Nanoparticles as antimicrobial and antiviral agents: A literature-based perspective study. Heliyon 2021, 7, e06456. [Google Scholar] [CrossRef] [PubMed]
  4. Gandra, S.; Barter, D.M.; Laxminarayan, R. Economic burden of antibiotic resistance: How much do we really know? Clin. Microbiol. Infect. 2014, 20, 973–980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Fadda, A.A.; Mohammed, A.R.; Abdel-Galil, E. Synthesis and antimicrobial evaluation of some 4-quinolinylazo-N-pyrimidinyl benzenesulfonamide derivatives. Biointerface Res. Appl. Chem. 2020, 10, 4846–4852. [Google Scholar] [CrossRef]
  6. Grozav, A.; Fedoriv, M.; Chornous, V.; Yakovychuk, N.; Kemskyi, S.; Vovk, M. Synthesis and Bioevaluation of 5-Chloro-4-(1,3-Oxazol-5-yl)-1H-Pyrrole-3-Carboxyamides as Antimicrobial Agents. Biointerface Res. Appl. Chem. 2021, 11, 10595–10606. [Google Scholar] [CrossRef]
  7. Kumar, A.; Kumar, A. Design and synthesis of anti-convulsant and anti-bacterial activity of new hydrazone derivatives. Biointerface Res. Appl. Chem. 2020, 10, 5229–5236. [Google Scholar] [CrossRef]
  8. Eleraky, N.E.; Allam, A.; Hassan, S.B.; Omar, M.M. Nanomedicine Fight against Antibacterial Resistance: An Overview of the Recent Pharmaceutical Innovations. Pharmaceutics 2020, 12, 142. [Google Scholar] [CrossRef] [Green Version]
  9. Devrim, B.; Bozkır, A. Chapter 7—Nanocarriers and Their Potential Application as Antimicrobial Drug Delivery. In Nanostructures for Antimicrobial Therapy; Ficai, A., Grumezescu, A.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 169–202. [Google Scholar]
  10. Anghel, A.G.; Grumezescu, A.M.; Chirea, M.; Grumezescu, V.; Socol, G.; Iordache, F.; Oprea, A.E.; Anghel, I.; Holban, A.M. MAPLE Fabricated Fe3O4@Cinnamomum verum Antimicrobial Surfaces for Improved Gastrostomy Tubes. Molecules 2014, 19, 8981. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Canaparo, R.; Foglietta, F.; Giuntini, F.; Della Pepa, C.; Dosio, F.; Serpe, L. Recent Developments in Antibacterial Therapy: Focus on Stimuli-Responsive Drug-Delivery Systems and Therapeutic Nanoparticles. Molecules 2019, 24, 1991. [Google Scholar] [CrossRef] [Green Version]
  12. Gherasim, O.; Grumezescu, A.M.; Grumezescu, V.; Iordache, F.; Vasile, B.S.; Holban, A.M. Bioactive Surfaces of Polylactide and Silver Nanoparticles for the Prevention of Microbial Contamination. Materials 2020, 13, 768. [Google Scholar] [CrossRef] [PubMed]
  13. Polívková, M.; Hubáček, T.; Staszek, M.; Švorčík, V.; Siegel, J. Antimicrobial Treatment of Polymeric Medical Devices by Silver Nanomaterials and Related Technology. Int. J. Mol. Sci. 2017, 18, 419. [Google Scholar] [CrossRef] [Green Version]
  14. Petrachi, T.; Resca, E.; Piccinno, M.S.; Biagi, F.; Strusi, V.; Dominici, M.; Veronesi, E. An Alternative Approach to Investigate Biofilm in Medical Devices: A Feasibility Study. Int. J. Environ. Res. Public Health 2017, 14, 1587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. French, G.L. The continuing crisis in antibiotic resistance. Int. J. Antimicrob. Agents 2010, 36, S3–S7. [Google Scholar] [CrossRef]
  16. de la Fuente-Núñez, C.; Reffuveille, F.; Fernández, L.; Hancock, R.E.W. Bacterial biofilm development as a multicellular adaptation: Antibiotic resistance and new therapeutic strategies. Curr. Opin. Microbiol. 2013, 16, 580–589. [Google Scholar] [CrossRef] [PubMed]
  17. Li, Y.; Li, X.; Hao, Y.; Liu, Y.; Dong, Z.; Li, K. Biological and Physiochemical Methods of Biofilm Adhesion Resistance Control of Medical-Context Surface. Int. J. Biol. Sci. 2021, 17, 1769–1781. [Google Scholar] [CrossRef] [PubMed]
  18. Vallet-Regí, M.; Lozano, D.; González, B.; Izquierdo-Barba, I. Biomaterials against Bone Infection. Adv. Healthc. Mater. 2020, 9, 2000310. [Google Scholar] [CrossRef] [PubMed]
  19. Grumezescu, V.; Negut, I.; Grumezescu, A.M.; Ficai, A.; Dorcioman, G.; Socol, G.; Iordache, F.; Truşcă, R.; Vasile, B.S.; Holban, A.M. MAPLE fabricated coatings based on magnetite nanoparticles embedded into biopolymeric spheres resistant to microbial colonization. Appl. Surf. Sci. 2018, 448, 230–236. [Google Scholar] [CrossRef]
  20. Mihai, A.D.; Chircov, C.; Grumezescu, A.M.; Holban, A.M. Magnetite Nanoparticles and Essential Oils Systems for Advanced Antibacterial Therapies. Int. J. Mol. Sci. 2020, 21, 7355. [Google Scholar] [CrossRef] [PubMed]
  21. Jamal, M.; Ahmad, W.; Andleeb, S.; Jalil, F.; Imran, M.; Nawaz, M.A.; Hussain, T.; Ali, M.; Rafiq, M.; Kamil, M.A. Bacterial biofilm and associated infections. J. Chin. Med. Assoc. JCMA 2018, 81, 7–11. [Google Scholar] [CrossRef] [PubMed]
  22. Wu, K.; Yang, Y.; Zhang, Y.; Deng, J.; Lin, C. Antimicrobial activity and cytocompatibility of silver nanoparticles coated catheters via a biomimetic surface functionalization strategy. Int. J. Nanomed. 2015, 10, 7241–7252. [Google Scholar] [CrossRef]
  23. Koley, S.; Mukherjee, M. Genetic Basis of Biofilm Formation and Spread of Nosocomial Infections. In Analytical Methodologies for Biofilm Research; Springer: New York, NY, USA, 2021; pp. 269–298. [Google Scholar]
  24. Jenkins, D.R. Nosocomial infections and infection control. Medicine 2017, 45, 629–633. [Google Scholar] [CrossRef]
  25. Pircalabioru, G.G.; Chifiriuc, M.-C. Nanoparticulate drug-delivery systems for fighting microbial biofilms: From bench to bedside. Future Microbiol. 2020, 15, 679–698. [Google Scholar] [CrossRef] [PubMed]
  26. Barros, C.H.N.; Casey, E. A Review of Nanomaterials and Technologies for Enhancing the Antibiofilm Activity of Natural Products and Phytochemicals. ACS Appl. Nano Mater. 2020, 3, 8537–8556. [Google Scholar] [CrossRef]
  27. Besinis, A.; Hadi, S.D.; Le, H.R.; Tredwin, C.; Handy, R.D. Antibacterial activity and biofilm inhibition by surface modified titanium alloy medical implants following application of silver, titanium dioxide and hydroxyapatite nanocoatings. Nanotoxicology 2017, 11, 327–338. [Google Scholar] [CrossRef] [Green Version]
  28. Makvandi, P.; Wang, C.-Y.; Zare, E.N.; Borzacchiello, A.; Niu, L.-N.; Tay, F.R. Metal-Based Nanomaterials in Biomedical Applications: Antimicrobial Activity and Cytotoxicity Aspects. Adv. Funct. Mater. 2020, 30, 1910021. [Google Scholar] [CrossRef]
  29. Gherasim, O.; Popescu, R.C.; Grumezescu, V.; Mogoșanu, G.D.; Mogoantă, L.; Iordache, F.; Holban, A.M.; Vasile, B.Ș.; Bîrcă, A.C.; Oprea, O.-C.; et al. MAPLE Coatings Embedded with Essential Oil-Conjugated Magnetite for Anti-Biofilm Applications. Materials 2021, 14, 1612. [Google Scholar] [CrossRef]
  30. Ahmadabadi, H.Y.; Yu, K.; Kizhakkedathu, J.N. Surface modification approaches for prevention of implant associated infections. Colloids Surf. B Biointerfaces 2020, 193, 111116. [Google Scholar] [CrossRef]
  31. Bahrami, A.; Delshadi, R.; Jafari, S.M. Active delivery of antimicrobial nanoparticles into microbial cells through surface functionalization strategies. Trends Food Sci. Technol. 2020, 99, 217–228. [Google Scholar] [CrossRef]
  32. Xie, S.; Tao, Y.; Pan, Y.; Qu, W.; Cheng, G.; Huang, L.; Chen, D.; Wang, X.; Liu, Z.; Yuan, Z. Biodegradable nanoparticles for intracellular delivery of antimicrobial agents. J. Control. Release 2014, 187, 101–117. [Google Scholar] [CrossRef]
  33. Liu, Z.; Ye, L.; Xi, J.; Wang, J.; Feng, Z.-g. Cyclodextrin polymers: Structure, synthesis, and use as drug carriers. Prog. Polym. Sci. 2021, 118, 101408. [Google Scholar] [CrossRef]
  34. Valenti, G.E.; Alfei, S.; Caviglia, D.; Domenicotti, C.; Marengo, B. Antimicrobial Peptides and Cationic Nanoparticles: A Broad-Spectrum Weapon to Fight Multi-Drug Resistance Not Only in Bacteria. Int. J. Mol. Sci. 2022, 23, 6108. [Google Scholar] [CrossRef] [PubMed]
  35. Tran, H.M.; Tran, H.; Booth, M.A.; Fox, K.E.; Nguyen, T.H.; Tran, N.; Tran, P.A. Nanomaterials for Treating Bacterial Biofilms on Implantable Medical Devices. Nanomaterials 2020, 10, 2253. [Google Scholar] [CrossRef] [PubMed]
  36. Delfi, M.; Ghomi, M.; Zarrabi, A.; Mohammadinejad, R.; Taraghdari, Z.B.; Ashrafizadeh, M.; Zare, E.N.; Agarwal, T.; Padil, V.V.T.; Mokhtari, B.; et al. Functionalization of Polymers and Nanomaterials for Biomedical Applications: Antimicrobial Platforms and Drug Carriers. Prosthesis 2020, 2, 117–139. [Google Scholar] [CrossRef]
  37. Lin, N.; Verma, D.; Saini, N.; Arbi, R.; Munir, M.; Jovic, M.; Turak, A. Antiviral nanoparticles for sanitizing surfaces: A roadmap to self-sterilizing against COVID-19. Nano Today 2021, 40, 101267. [Google Scholar] [CrossRef] [PubMed]
  38. Soares, S.; Sousa, J.; Pais, A.; Vitorino, C. Nanomedicine: Principles, Properties, and Regulatory Issues. Front. Chem. 2018, 6, 360. [Google Scholar] [CrossRef] [PubMed]
  39. Chapman, J.; Regan, F.; Sullivan, T. Nanoparticles in Anti-Microbial Materials: Use and Characterisation; Royal Society of Chemistry: Cambridge, UK, 2012; Volume 23. [Google Scholar]
  40. Nas, F.S.; Ali, M.; Aminu Muhammad, A. Application of Nanomaterials as Antimicrobial Agents: A Review. Arch. Nano Op. Acc. J. 2018, 1, 59–64. [Google Scholar]
  41. Kim, T.; Hyeon, T. Applications of inorganic nanoparticles as therapeutic agents. Nanotechnology 2013, 25, 012001. [Google Scholar] [CrossRef]
  42. Gilavand, F.; Saki, R.; Mirzaei, S.Z.; Lashgarian, H.E.; Karkhane, M.; Marzban, A. Green Synthesis of Zinc Nanoparticles Using Aqueous Extract of Magnoliae officinalis and Assessment of its Bioactivity Potentials. Biointerface Res. Appl. Chem. 2021, 11, 7765–7774. [Google Scholar] [CrossRef]
  43. Samrot, A.V.; Sahithya, C.S.; Sruthi, P.D.; Selvarani, A.J.; Raji, P.; Prakash, P.; Ponnaiah, P.; Petchi, I.; Pattammadath, S.; Purayil, S.K.; et al. Itraconazole Coated Super Paramagnetic Iron Oxide Nanoparticles for Antimicrobial Studies. Biointerface Res. Appl. Chem. 2020, 10, 6262–6269. [Google Scholar] [CrossRef]
  44. Tyagi, P.K.; Mishra, R.; Khan, F.; Gupta, D.; Gola, D. Antifungal Effects of Silver Nanoparticles Against Various Plant Pathogenic Fungi and its Safety Evaluation on Drosophila melanogaster. Biointerface Res. Appl. Chem. 2020, 10, 6587–6596. [Google Scholar] [CrossRef]
  45. Masri, A.; Anwar, A.; Khan, N.A.; Siddiqui, R. The Use of Nanomedicine for Targeted Therapy against Bacterial Infections. Antibiotics 2019, 8, 260. [Google Scholar] [CrossRef] [PubMed]
  46. Niculescu, A.-G.; Chircov, C.; Bîrcă, A.C.; Grumezescu, A.M. Nanomaterials Synthesis through Microfluidic Methods: An Updated Overview. Nanomaterials 2021, 11. [Google Scholar] [CrossRef] [PubMed]
  47. Arole, V.M.; Munde, S.V. Fabrication of nanomaterials by top-down and bottom-up approaches-an overview. J. Mater. Sci 2014, 1, 89–93. [Google Scholar]
  48. Shrimal, P.; Jadeja, G.; Patel, S. A review on novel methodologies for drug nanoparticle preparation: Microfluidic approach. Chem. Eng. Res. Des. 2020, 153, 728–756. [Google Scholar] [CrossRef]
  49. Niculescu, A.-G.; Chircov, C.; Grumezescu, A.M. Magnetite nanoparticles: Synthesis methods—A comparative review. Methods 2022, 199, 16–27. [Google Scholar] [CrossRef]
  50. Niculescu, A.-G.; Grumezescu, A.M. Polymer-Based Nanosystems—A Versatile Delivery Approach. Materials 2021, 14. [Google Scholar] [CrossRef]
  51. Zielińska, A.; Carreiró, F.; Oliveira, A.M.; Neves, A.; Pires, B.; Venkatesh, D.N.; Durazzo, A.; Lucarini, M.; Eder, P.; Silva, A.M. Polymeric nanoparticles: Production, characterization, toxicology and ecotoxicology. Molecules 2020, 25, 3731. [Google Scholar] [CrossRef]
  52. Niculescu, A.-G.; Bîrcă, A.C.; Grumezescu, A.M. New Applications of Lipid and Polymer-Based Nanoparticles for Nucleic Acids Delivery. Pharmaceutics 2021, 13, 2053. [Google Scholar] [CrossRef]
  53. Lundberg, S.; Karlsson, E.; Dahlberg, H.; Glansk, M.; Larsson, S.; Larsson, S.; Carlsson, K. Exosomes and Lipid Nanoparticles-the Future of Targeted Drug Delivery. Ph.D. Thesis, Uppsala University, Uppsala, Sweden, 2020. [Google Scholar]
  54. Tiplea, R.E.; Lemnaru, G.M.; Trusca, R.D.; Holban, A.; Kaya, M.G.A.; Dragu, L.D.; Ficai, D.; Ficai, A.; Bleotu, C. Antimicrobial Films based on Chitosan, Collagen, and ZnO for Skin Tissue Regeneration. Biointerface Res. Appl. Chem. 2021, 11, 11985–11995. [Google Scholar] [CrossRef]
  55. Scolari, I.R.; Páez, P.L.; Musri, M.M.; Petiti, J.P.; Torres, A.; Granero, G.E. Rifampicin loaded in alginate/chitosan nanoparticles as a promising pulmonary carrier against Staphylococcus aureus. Drug Deliv. Transl. Res. 2020, 10, 1403–1417. [Google Scholar] [CrossRef]
  56. Varela, M.F.; Stephen, J.; Lekshmi, M.; Ojha, M.; Wenzel, N.; Sanford, L.M.; Hernandez, A.J.; Parvathi, A.; Kumar, S.H. Bacterial Resistance to Antimicrobial Agents. Antibiotics 2021, 10, 593. [Google Scholar] [CrossRef] [PubMed]
  57. Mohid, S.A.; Bhunia, A. Combining Antimicrobial Peptides with Nanotechnology: An Emerging Field in Theranostics. Curr. Protein Pept. Sci. 2020, 21, 413–428. [Google Scholar] [CrossRef] [PubMed]
  58. Nag, M.; Lahiri, D.; Mukherjee, D.; Banerjee, R.; Garai, S.; Sarkar, T.; Ghosh, S.; Dey, A.; Ghosh, S.; Pattnaik, S.; et al. Functionalized Chitosan Nanomaterials: A Jammer for Quorum Sensing. Polymers 2021, 13, 2533. [Google Scholar] [CrossRef] [PubMed]
  59. Kong, J.; Zhang, S.; Shen, M.; Zhang, J.; Yoganathan, S. Evaluation of copper(I)-doped zinc oxide composite nanoparticles on both gram-negative and gram-positive bacteria. Colloids Surf. A Physicochem. Eng. Asp. 2022, 643, 128742. [Google Scholar] [CrossRef]
  60. Pietsch, F.; O’Neill, A.J.; Ivask, A.; Jenssen, H.; Inkinen, J.; Kahru, A.; Ahonen, M.; Schreiber, F. Selection of resistance by antimicrobial coatings in the healthcare setting. J. Hosp. Infect. 2020, 106, 115–125. [Google Scholar] [CrossRef] [PubMed]
  61. Nikaido, H. Multidrug resistance in bacteria. Annu. Rev. Biochem. 2009, 78, 119–146. [Google Scholar] [CrossRef] [Green Version]
  62. Wang, L.; Hu, C.; Shao, L. The antimicrobial activity of nanoparticles: Present situation and prospects for the future. Int. J. Nanomed. 2017, 12, 1227–1249. [Google Scholar] [CrossRef] [Green Version]
  63. Fernando, S.S.N.; Gunasekara, C.; Holton, J. Antimicrobial Nanoparticles: Applications and mechanisms of action. Sri Lankan J. Infect. Dis. 2018, 8, 2. [Google Scholar] [CrossRef]
  64. Varier, K.M.; Gudeppu, M.; Chinnasamy, A.; Thangarajan, S.; Balasubramanian, J.; Li, Y.; Gajendran, B. Nanoparticles: Antimicrobial Applications and Its Prospects. In Advanced Nanostructured Materials for Environmental Remediation; Springer: Cham, Switzerland, 2019; pp. 321–355. [Google Scholar] [CrossRef]
  65. Lee, N.-Y.; Ko, W.-C.; Hsueh, P.-R. Nanoparticles in the Treatment of Infections Caused by Multidrug-Resistant Organisms. Front. Pharmacol. 2019, 10, 1153. [Google Scholar] [CrossRef] [Green Version]
  66. 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]
  67. 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] [PubMed]
  68. Gupta, K.; Chundawat, T.S.; Malek, N. Antibacterial, Antifungal, Photocatalytic Activities and Seed Germination Effect of Mycosynthesized Silver Nanoparticles using Fusarium oxysporum. Biointerface Res. Appl. Chem. 2021, 11, 12082–12091. [Google Scholar] [CrossRef]
  69. Khan, K.; Javed, S. Functionalization of Inorganic Nanoparticles to Augment Antimicrobial Efficiency: A Critical Analysis. Curr. Pharm. Biotechnol. 2018, 19, 523–536. [Google Scholar] [CrossRef] [PubMed]
  70. Malaekeh-Nikouei, B.; Fazly Bazzaz, B.S.; Mirhadi, E.; Tajani, A.S.; Khameneh, B. The role of nanotechnology in combating biofilm-based antibiotic resistance. J. Drug Deliv. Sci. Technol. 2020, 60, 101880. [Google Scholar] [CrossRef]
  71. Karuppannan, S.K.; Ramalingam, R.; Mohamed Khalith, S.B.; Dowlath, M.J.H.; Darul Raiyaan, G.I.; Arunachalam, K.D. Characterization, antibacterial and photocatalytic evaluation of green synthesized copper oxide nanoparticles. Biocatal. Agric. Biotechnol. 2021, 31, 101904. [Google Scholar] [CrossRef]
  72. Caciandone, M.; Niculescu, A.-G.; Grumezescu, V.; Bîrcă, A.C.; Ghica, I.C.; Vasile, B.Ș.; Oprea, O.; Nica, I.C.; Stan, M.S.; Holban, A.M.; et al. Magnetite Nanoparticles Functionalized with Therapeutic Agents for Enhanced ENT Antimicrobial Properties. Antibiotics 2022, 11, 623. [Google Scholar] [CrossRef]
  73. Abduraimova, A.; Molkenova, A.; Duisembekova, A.; Mulikova, T.; Kanayeva, D.; Atabaev, T.S. Cetyltrimethylammonium Bromide (CTAB)-Loaded SiO2–Ag Mesoporous Nanocomposite as an Efficient Antibacterial Agent. Nanomaterials 2021, 11, 477. [Google Scholar] [CrossRef]
  74. Rashid, M.; Rabbi, M.A.; Ara, T.; Hossain, M.M.; Islam, M.S.; Elaissari, A.; Ahmad, H.; Rahman, M.M. Vancomycin conjugated iron oxide nanoparticles for magnetic targeting and efficient capture of Gram-positive and Gram-negative bacteria. RSC Adv. 2021, 11, 36319–36328. [Google Scholar] [CrossRef]
  75. Rayegan, A.; Allafchian, A.; Abdolhosseini Sarsari, I.; Kameli, P. Synthesis and characterization of basil seed mucilage coated Fe3O4 magnetic nanoparticles as a drug carrier for the controlled delivery of cephalexin. Int. J. Biol. Macromol. 2018, 113, 317–328. [Google Scholar] [CrossRef]
  76. Lage, W.C.; Sachs, D.; Nunes Ribeiro, T.A.; Tebaldi, M.L.; de Moura, Y.d.R.S.; Domingues, S.C.; Ferreira Soares, D.C. Mesoporous iron oxide nanoparticles loaded with ciprofloxacin as a potential biocompatible antibacterial system. Microporous Mesoporous Mater. 2021, 321, 111127. [Google Scholar] [CrossRef]
  77. Caciandone, M.; Niculescu, A.-G.; Roșu, A.R.; Grumezescu, V.; Negut, I.; Holban, A.M.; Oprea, O.; Vasile, B.Ș.; Bîrcă, A.C.; Grumezescu, A.M.; et al. PEG-Functionalized Magnetite Nanoparticles for Modulation of Microbial Biofilms on Voice Prosthesis. Antibiotics 2022, 11, 39. [Google Scholar] [CrossRef] [PubMed]
  78. de Lacerda Coriolano, D.; de Souza, J.B.; Bueno, E.V.; Medeiros, S.M.d.F.R.d.S.; Cavalcanti, I.D.L.; Cavalcanti, I.M.F. Antibacterial and antibiofilm potential of silver nanoparticles against antibiotic-sensitive and multidrug-resistant Pseudomonas aeruginosa strains. Braz. J. Microbiol. 2020, 52, 267–278. [Google Scholar] [CrossRef] [PubMed]
  79. Camacho-Jiménez, L.; Álvarez-Sánchez, A.R.; Mejía-Ruíz, C.H. Silver nanoparticles (AgNPs) as antimicrobials in marine shrimp farming: A review. Aquac. Rep. 2020, 18, 100512. [Google Scholar] [CrossRef]
  80. Hamad, A.; Khashan, K.S.; Hadi, A. Silver Nanoparticles and Silver Ions as Potential Antibacterial Agents. J. Inorg. Organomet. Polym. Mater. 2020, 30, 4811–4828. [Google Scholar] [CrossRef]
  81. Balakrishnan, K.; Casimeer, S.C.; Ghidan, A.Y.; Al Antary, T.M.; Singaravelu, A. Exploration of Antioxidant, Antibacterial Activities of Green Synthesized Hesperidin Loaded PLGA Nanoparticles. Biointerface Res. Appl. Chem. 2021, 11, 14520–14528. [Google Scholar] [CrossRef]
  82. Ameeduzzafar; Imam, S.S.; Abbas Bukhari, S.N.; Ahmad, J.; Ali, A. Formulation and optimization of levofloxacin loaded chitosan nanoparticle for ocular delivery: In-vitro characterization, ocular tolerance and antibacterial activity. Int. J. Biol. Macromol. 2018, 108, 650–659. [Google Scholar] [CrossRef] [PubMed]
  83. Alruwaili, N.K.; Zafar, A.; Imam, S.S.; Alharbi, K.S.; Alotaibi, N.H.; Alshehri, S.; Alhakamy, N.A.; Alzarea, A.I.; Afzal, M.; Elmowafy, M. Stimulus Responsive Ocular Gentamycin-Ferrying Chitosan Nanoparticles Hydrogel: Formulation Optimization, Ocular Safety and Antibacterial Assessment. Int. J. Nanomed. 2020, 15, 4717–4737. [Google Scholar] [CrossRef] [PubMed]
  84. Mohammadi, A.; Hosseini, S.M.; Hashemi, M. Emerging chitosan nanoparticles loading-system boosted the antibacterial activity of Cinnamomum zeylanicum essential oil. Ind. Crops Prod. 2020, 155, 112824. [Google Scholar] [CrossRef]
  85. Bagheri, R.; Ariaii, P.; Motamedzadegan, A. Characterization, antioxidant and antibacterial activities of chitosan nanoparticles loaded with nettle essential oil. J. Food Meas. Charact. 2021, 15, 1395–1402. [Google Scholar] [CrossRef]
  86. Andrade, L.F.d.; Apolinário, A.C.; Rangel-Yagui, C.O.; Stephano, M.A.; Tavares, L.C. Chitosan nanoparticles for the delivery of a new compound active against multidrug-resistant Staphylococcus aureus. J. Drug Deliv. Sci. Technol. 2020, 55, 101363. [Google Scholar] [CrossRef]
  87. Kaur, J.; Kour, A.; Panda, J.J.; Harjai, K.; Chhibber, S. Exploring Endolysin-Loaded Alginate-Chitosan Nanoparticles as Future Remedy for Staphylococcal Infections. AAPS PharmSciTech 2020, 21, 233. [Google Scholar] [CrossRef] [PubMed]
  88. Kadhum, W.N.; Zaidan, I.A. The synergistic effects of chitosan-alginate nanoparticles loaded with doxycycline antibiotic against multidrug resistant proteus mirabilis, Escherichia coli and enterococcus faecalis. Iraqi J. Sci. 2020, 61, 3187–3199. [Google Scholar]
  89. Kumar, D.; Kumar, S.; Kumar, S.; Rohatgi, S.; Kundu, P.P. Synthesis of rifaximin loaded chitosan-alginate core-shell nanoparticles (Rif@CS/Alg-NPs) for antibacterial applications. Int. J. Biol. Macromol. 2021, 183, 962–971. [Google Scholar] [CrossRef] [PubMed]
  90. Yoncheva, K.; Benbassat, N.; Zaharieva, M.M.; Dimitrova, L.; Kroumov, A.; Spassova, I.; Kovacheva, D.; Najdenski, H.M. Improvement of the Antimicrobial Activity of Oregano Oil by Encapsulation in Chitosan—Alginate Nanoparticles. Molecules 2021, 26, 7017. [Google Scholar] [CrossRef] [PubMed]
  91. Ciro, Y.; Rojas, J.; Oñate-Garzon, J.; Salamanca, C.H. Synthesis, Characterisation and Biological Evaluation of Ampicillin–Chitosan–Polyanion Nanoparticles Produced by Ionic Gelation and Polyelectrolyte Complexation Assisted by High-Intensity Sonication. Polymers 2019, 11, 1758. [Google Scholar] [CrossRef] [Green Version]
  92. Qiu, Y.; Xu, D.; Sui, G.; Wang, D.; Wu, M.; Han, L.; Mu, H.; Duan, J. Gentamicin decorated phosphatidylcholine-chitosan nanoparticles against biofilms and intracellular bacteria. Int. J. Biol. Macromol. 2020, 156, 640–647. [Google Scholar] [CrossRef]
  93. Falciani, C.; Zevolini, F.; Brunetti, J.; Riolo, G.; Gracia, R.; Marradi, M.; Loinaz, I.; Ziemann, C.; Cossío, U.; Llop, J.; et al. Antimicrobial Peptide-Loaded Nanoparticles as Inhalation Therapy for Pseudomonas aeruginosa Infections. Int. J. Nanomed. 2020, 15, 1117–1128. [Google Scholar] [CrossRef] [Green Version]
  94. Da Costa, D.; Exbrayat-Héritier, C.; Rambaud, B.; Megy, S.; Terreux, R.; Verrier, B.; Primard, C. Surface charge modulation of rifampicin-loaded PLA nanoparticles to improve antibiotic delivery in Staphylococcus aureus biofilms. J. Nanobiotechnology 2021, 19, 12. [Google Scholar] [CrossRef]
  95. Vrouvaki, I.; Koutra, E.; Kornaros, M.; Avgoustakis, K.; Lamari, F.N.; Hatziantoniou, S. Polymeric Nanoparticles of Pistacia lentiscus var. chia Essential Oil for Cutaneous Applications. Pharmaceutics 2020, 12, 353. [Google Scholar] [CrossRef] [Green Version]
  96. De Mélo Silva, I.S.; do Amorim Costa Gaspar, L.M.; Rocha, A.M.O.; da Costa, L.P.; Tada, D.B.; Franceschi, E.; Padilha, F.F. Encapsulation of Red Propolis in Polymer Nanoparticles for the Destruction of Pathogenic Biofilms. AAPS PharmSciTech 2020, 21, 49. [Google Scholar] [CrossRef]
  97. Ucak, S.; Sudagidan, M.; Borsa, B.A.; Mansuroglu, B.; Ozalp, V.C. Inhibitory effects of aptamer targeted teicoplanin encapsulated PLGA nanoparticles for Staphylococcus aureus strains. World J. Microbiol. Biotechnol. 2020, 36, 69. [Google Scholar] [CrossRef] [PubMed]
  98. Ma, X.; Lang, J.; Chen, P.; Yang, R. Silver nanoparticles as an effective antimicrobial against otitis media pathogens. AIChE J. 2021, 67, e17468. [Google Scholar] [CrossRef] [PubMed]
  99. Al-Gethami, W.; Al-Qasmi, N. Antimicrobial Activity of Ca-Alginate/Chitosan Nanocomposite Loaded with Camptothecin. Polymers 2021, 13, 3559. [Google Scholar] [CrossRef] [PubMed]
  100. Rezaei, F.; Damoogh, S.; Reis, R.L.; Kundu, S.C.; Mottaghitalab, F.; Farokhi, M. Dual drug delivery system based on pH-sensitive silk fibroin/alginate nanoparticles entrapped in PNIPAM hydrogel for treating severe infected burn wound. Biofabrication 2020, 13, 015005. [Google Scholar] [CrossRef] [PubMed]
  101. Hassan, Y.A.; Khedr, A.I.M.; Alkabli, J.; Elshaarawy, R.F.M.; Nasr, A.M. Co-delivery of imidazolium Zn(II)salen and Origanum Syriacum essential oil by shrimp chitosan nanoparticles for antimicrobial applications. Carbohydr. Polym. 2021, 260, 117834. [Google Scholar] [CrossRef]
  102. Balabathula, P.; Whaley, S.G.; Janagam, D.R.; Mittal, N.K.; Mandal, B.; Thoma, L.A.; Rogers, P.D.; Wood, G.C. Lyophilized Iron Oxide Nanoparticles Encapsulated in Amphotericin B: A Novel Targeted Nano Drug Delivery System for the Treatment of Systemic Fungal Infections. Pharmaceutics 2020, 12, 247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Luis Enrique Jerez, P. Chapter 6 – Fungal Infections in Immunosuppressed Patients. In Immunodeficiency; Krassimir, M., Ed.; IntechOpen: Rijeka, Croatia, 2012. [Google Scholar]
  104. Patterson, T.F. The role of echinocandins, extended-spectrum azoles, and polyenes to treat opportunistic moulds and candida. Curr. Infect. Dis. Rep. 2006, 8, 442–448. [Google Scholar] [CrossRef]
  105. Quindós, G.; Gil-Alonso, S.; Marcos-Arias, C.; Sevillano, E.; Mateo, E.; Jauregizar, N.; Eraso, E. Therapeutic tools for oral candidiasis: Current and new antifungal drugs. Med. Oral Patol. Oral Cir. Bucal 2019, 24, e172–e180. [Google Scholar] [CrossRef] [PubMed]
  106. Araujo, V.H.S.; Duarte, J.L.; Carvalho, G.C.; Silvestre, A.L.P.; Fonseca-Santos, B.; Marena, G.D.; Ribeiro, T.d.C.; dos Santos Ramos, M.A.; Bauab, T.M.; Chorilli, M. Nanosystems against candidiasis: A review of studies performed over the last two decades. Crit. Rev. Microbiol. 2020, 46, 508–547. [Google Scholar] [CrossRef]
  107. Muñoz, J.E.; Rossi, D.C.P.; Jabes, D.L.; Barbosa, D.A.; Cunha, F.F.M.; Nunes, L.R.; Arruda, D.C.; Pelleschi Taborda, C. In Vitro and In Vivo Inhibitory Activity of Limonene against Different Isolates of Candida spp. J. Fungi 2020, 6, 183. [Google Scholar] [CrossRef]
  108. Biswas, K.D.; Choudhary, A.; Ghosh, S.K.; Biswas, S. Primary Laryngeal Aspergillosis in an Immunocompetent Host. Bengal J. Otolaryngol. Head Neck Surg. 2018, 26, 131–133. [Google Scholar] [CrossRef]
  109. Hussain, M.A.; Ahmed, D.; Anwar, A.; Perveen, S.; Ahmed, S.; Anis, I.; Shah, M.R.; Khan, N.A. Combination Therapy of Clinically Approved Antifungal Drugs Is Enhanced by Conjugation with Silver Nanoparticles. Int. Microbiol. 2019, 22, 239–246. [Google Scholar] [CrossRef] [PubMed]
  110. Alshahrani, S.M.; Khafagy, E.-S.; Riadi, Y.; Al Saqr, A.; Alfadhel, M.M.; Hegazy, W.A.H. Amphotericin B-PEG Conjugates of ZnO Nanoparticles: Enhancement Antifungal Activity with Minimal Toxicity. Pharmaceutics 2022, 14, 1646. [Google Scholar] [CrossRef] [PubMed]
  111. Shah, M.K.; Azad, A.K.; Nawaz, A.; Ullah, S.; Latif, M.S.; Rahman, H.; Alsharif, K.F.; Alzahrani, K.J.; El-Kott, A.F.; Albrakati, A.; et al. Formulation Development, Characterization and Antifungal Evaluation of Chitosan NPs for Topical Delivery of Voriconazole In Vitro and Ex Vivo. Polymers 2022, 14, 135. [Google Scholar] [CrossRef] [PubMed]
  112. Elshaer, E.E.; Elwakil, B.H.; Eskandrani, A.; Elshewemi, S.S.; Olama, Z.A. Novel Clotrimazole and Vitis vinifera loaded chitosan nanoparticles: Antifungal and wound healing efficiencies. Saudi J. Biol. Sci. 2022, 29, 1832–1841. [Google Scholar] [CrossRef]
  113. Orekhova, A.; Palocci, C.; Chronopoulou, L.; De Angelis, G.; Badiali, C.; Petruccelli, V.; D’Angeli, S.; Pasqua, G.; Simonetti, G. Poly-(lactic-co-glycolic) Acid Nanoparticles Entrapping Pterostilbene for Targeting Aspergillus Section Nigri. Molecules 2022, 27, 5424. [Google Scholar] [CrossRef]
  114. Yenice Gürsu, B. Potential antibiofilm activity of farnesol-loaded poly(DL-lactide-co-glycolide) (PLGA) nanoparticles against Candida albicans. J. Anal. Sci. Technol. 2020, 11, 43. [Google Scholar] [CrossRef]
  115. Alhowyan, A.A.; Altamimi, M.A.; Kalam, M.A.; Khan, A.A.; Badran, M.; Binkhathlan, Z.; Alkholief, M.; Alshamsan, A. Antifungal efficacy of Itraconazole loaded PLGA-nanoparticles stabilized by vitamin-E TPGS: In vitro and ex vivo studies. J. Microbiol. Methods 2019, 161, 87–95. [Google Scholar] [CrossRef]
  116. Mejía, S.P.; López, D.; Cano, L.E.; Naranjo, T.W.; Orozco, J. Antifungal Encapsulated into Ligand-Functionalized Nanoparticles with High Specificity for Macrophages. Pharmaceutics 2022, 14, 1932. [Google Scholar] [CrossRef]
  117. Hou, Y.; Yang, M.; Li, J.; Bi, X.; Li, G.; Xu, J.; Xie, S.; Dong, Y.; Li, D.; Du, Y. The enhancing antifungal effect of AD1 aptamer-functionalized amphotericin B-loaded PLGA-PEG nanoparticles with a low-frequency and low-intensity ultrasound exposure on C.albicans biofilm through targeted effect. NanoImpact 2021, 21, 100275. [Google Scholar] [CrossRef]
  118. Endo, E.H.; Makimori, R.Y.; Companhoni, M.V.P.; Ueda-Nakamura, T.; Nakamura, C.V.; Dias Filho, B.P. Ketoconazole-loaded poly-(lactic acid) nanoparticles: Characterization and improvement of antifungal efficacy in vitro against Candida and dermatophytes. J. Mycol. Médicale 2020, 30, 101003. [Google Scholar] [CrossRef] [PubMed]
  119. Al-Maghrabi, P.M.; Khafagy, E.-S.; Ghorab, M.M.; Gad, S. Influence of formulation variables on miconazole nitrate–loaded lipid based nanocarrier for topical delivery. Colloids Surf. B Biointerfaces 2020, 193, 111046. [Google Scholar] [CrossRef] [PubMed]
  120. Baldim, I.; Paziani, M.H.; Grizante Barião, P.H.; Kress, M.R.v.Z.; Oliveira, W.P. Nanostructured Lipid Carriers Loaded with Lippia sidoides Essential Oil as a Strategy to Combat the Multidrug-Resistant Candida auris. Pharmaceutics 2022, 14, 180. [Google Scholar] [CrossRef]
  121. Lee, Y.; Puumala, E.; Robbins, N.; Cowen, L.E. Antifungal Drug Resistance: Molecular Mechanisms in Candida albicans and Beyond. Chem. Rev. 2021, 121, 3390–3411. [Google Scholar] [CrossRef]
  122. Lim, C.S.Y.; Rosli, R.; Seow, H.F.; Chong, P.P. Candida and invasive candidiasis: Back to basics. Eur. J. Clin. Microbiol. Infect. Dis. 2012, 31, 21–31. [Google Scholar] [CrossRef] [PubMed]
  123. Jagaran, K.; Singh, M. Nanomedicine for COVID-19: Potential of Copper Nanoparticles. Biointerface Res. Appl. Chem. 2021, 11, 10716–10728. [Google Scholar] [CrossRef]
  124. Delshadi, R.; Bahrami, A.; McClements, D.J.; Moore, M.D.; Williams, L. Development of nanoparticle-delivery systems for antiviral agents: A review. J. Control. Release 2021, 331, 30–44. [Google Scholar] [CrossRef] [PubMed]
  125. Pokhrel, L.R.; Williams, F.; Cook, P.P.; O’Rourke, D.; Murray, G.; Akula, S.M. Preclinical efficacy and safety of novel SNAT against SARS-CoV-2 using a hamster model. Drug Deliv. Transl. Res. 2022, 12, 3007–3016. [Google Scholar] [CrossRef] [PubMed]
  126. León-Gutiérrez, G.; Elste, J.E.; Cabello-Gutiérrez, C.; Millán-Pacheco, C.; Martínez-Gómez, M.H.; Mejía-Alvarez, R.; Tiwari, V.; Mejía, A. A potent virucidal activity of functionalized TiO2 nanoparticles adsorbed with flavonoids against SARS-CoV-2. Appl. Microbiol. Biotechnol. 2022, 106, 5987–6002. [Google Scholar] [CrossRef]
  127. Joshy, K.S.; Susan, M.A.; Snigdha, S.; Nandakumar, K.; Laly, A.P.; Sabu, T. Encapsulation of zidovudine in PF-68 coated alginate conjugate nanoparticles for anti-HIV drug delivery. Int. J. Biol. Macromol. 2018, 107, 929–937. [Google Scholar] [CrossRef]
  128. Joshy, K.S.; George, A.; Snigdha, S.; Joseph, B.; Kalarikkal, N.; Pothen, L.A.; Thomas, S. Novel core-shell dextran hybrid nanosystem for anti-viral drug delivery. Mater. Sci. Eng. C 2018, 93, 864–872. [Google Scholar] [CrossRef] [PubMed]
  129. Belgamwar, A.V.; Khan, S.A.; Yeole, P.G. Intranasal dolutegravir sodium loaded nanoparticles of hydroxypropyl-beta-cyclodextrin for brain delivery in Neuro-AIDS. J. Drug Deliv. Sci. Technol. 2019, 52, 1008–1020. [Google Scholar] [CrossRef]
  130. Belgamwar, A.; Khan, S.; Yeole, P. Intranasal chitosan-g-HPβCD nanoparticles of efavirenz for the CNS targeting. Artif. Cells Nanomed. Biotechnol. 2018, 46, 374–386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Donalisio, M.; Argenziano, M.; Rittà, M.; Bastiancich, C.; Civra, A.; Lembo, D.; Cavalli, R. Acyclovir-loaded sulfobutyl ether-β-cyclodextrin decorated chitosan nanodroplets for the local treatment of HSV-2 infections. Int. J. Pharm. 2020, 587, 119676. [Google Scholar] [CrossRef]
  132. Parthiban, R.; Sathishkumar, S.; Ramakrishnan, P. Design and evaluation of acyclovir-loaded solid lipid nanoparticles for sustained release. Drug Invent. Today 2020, 14, 108–111. [Google Scholar]
  133. Javan, F.; Vatanara, A.; Azadmanesh, K.; Nabi-Meibodi, M.; Shakouri, M. Encapsulation of ritonavir in solid lipid nanoparticles: In-vitro anti-HIV-1 activity using lentiviral particles. J. Pharm. Pharmacol. 2017, 69, 1002–1009. [Google Scholar] [CrossRef]
  134. Figueira, T.N.; Domingues, M.M.; Illien, F.; Cadima-Couto, I.; Todorovski, T.; Andreu, D.; Sagan, S.; Castanho, M.A.R.B.; Walrant, A.; Veiga, A.S. Enfuvirtide-Protoporphyrin IX Dual-Loaded Liposomes: In Vitro Evidence of Synergy against HIV-1 Entry into Cells. ACS Infect. Dis. 2020, 6, 224–236. [Google Scholar] [CrossRef]
  135. Cao, S.; Slack, S.D.; Levy, C.N.; Hughes, S.M.; Jiang, Y.; Yogodzinski, C.; Roychoudhury, P.; Jerome, K.R.; Schiffer, J.T.; Hladik, F.; et al. Hybrid nanocarriers incorporating mechanistically distinct drugs for lymphatic CD4+ T cell activation and HIV-1 latency reversal. Sci. Adv. 2019, 5, eaav6322. [Google Scholar] [CrossRef] [Green Version]
  136. Snoussi, Y.; Sifaoui, I.; Khalil, A.M.; Bhakta, A.K.; Semyonov, O.; Postnikov, P.S.; Michely, L.; Pires, R.; Bastide, S.; Barroso, J.E.-P.; et al. Facile synthesis of silver decorated biochar as a novel and highly active biosourced anti-kinetoplastid agent. Mater. Today Commun. 2022, 32, 104126. [Google Scholar] [CrossRef]
  137. Badirzadeh, A.; Alipour, M.; Najm, M.; Vosoogh, A.; Vosoogh, M.; Samadian, H.; Hashemi, A.S.; Farsangi, Z.J.; Amini, S.M. Potential therapeutic effects of curcumin coated silver nanoparticle in the treatment of cutaneous leishmaniasis due to Leishmania major in-vitro and in a murine model. J. Drug Deliv. Sci. Technol. 2022, 74, 103576. [Google Scholar] [CrossRef]
  138. Antinori, S.; Schifanella, L.; Corbellino, M. Leishmaniasis: New insights from an old and neglected disease. Eur. J. Clin. Microbiol. Infect. Dis. 2012, 31, 109–118. [Google Scholar] [CrossRef] [PubMed]
  139. Durak, S.; Arasoglu, T.; Ates, S.C.; Derman, S. Enhanced antibacterial and antiparasitic activity of multifunctional polymeric nanoparticles. Nanotechnology 2020, 31, 175705. [Google Scholar] [CrossRef] [PubMed]
  140. Kanwal; Mungroo, M.R.; Anwar, A.; Ali, F.; Khan, S.; Abdullah, M.A.; Siddiqui, R.; Khan, K.M.; Khan, N.A. Synthetic nanoparticle-conjugated bisindoles and hydrazinyl arylthiazole as novel antiamoebic agents against brain-eating amoebae. Exp. Parasitol. 2020, 218, 107979. [Google Scholar] [CrossRef] [PubMed]
  141. Real, D.; Hoffmann, S.; Leonardi, D.; Salomon, C.; Goycoolea, F.M. Chitosan-based nanodelivery systems applied to the development of novel triclabendazole formulations. PLoS ONE 2018, 13, e0207625. [Google Scholar] [CrossRef]
  142. Wei, W.; Lu, M.; Xu, W.; Polyakov, N.E.; Dushkin, A.V.; Su, W.-k. Preparation of protamine-hyaluronic acid coated core-shell nanoparticles for enhanced solubility, permeability, and oral bioavailability of decoquinate. Int. J. Biol. Macromol. 2022, 218, 346–355. [Google Scholar] [CrossRef]
  143. Taghe, S.; Mirzaeei, S. Preparation and characterization of novel, mucoadhesive ofloxacin nanoparticles for ocular drug delivery. Braz. J. Pharm. Sci. 2019, 55, e17105. [Google Scholar] [CrossRef] [Green Version]
  144. Sanchez-Guzman, D.; Le Guen, P.; Villeret, B.; Sola, N.; Le Borgne, R.; Guyard, A.; Kemmel, A.; Crestani, B.; Sallenave, J.-M.; Garcia-Verdugo, I. Silver nanoparticle-adjuvanted vaccine protects against lethal influenza infection through inducing BALT and IgA-mediated mucosal immunity. Biomaterials 2019, 217, 119308. [Google Scholar] [CrossRef]
  145. Lina, T.T.; Johnson, S.J.; Wagner, R.D. Intravaginal poly-(D, L-lactic-co-glycolic acid)-(polyethylene glycol) drug-delivery nanoparticles induce pro-inflammatory responses with Candida albicans infection in a mouse model. PLoS ONE 2020, 15, e0240789. [Google Scholar] [CrossRef]
  146. Atinderpal, K.; Kapoor, N.; Gupta, S.; Tyag, A.; Sharma, R.K.; Ali, J.; Gabrani, R.; Dang, S. Development and characterization of green tea catechins and ciprofloxacin-loaded nanoemulsion for intravaginal delivery to treat urinary tract infection. Indian J. Pharm. Sci. 2018, 80, 442–452. [Google Scholar] [CrossRef]
  147. Brauner, B.; Semmler, J.; Rauch, D.; Nokaj, M.; Haiss, P.; Schwarz, P.; Wirth, M.; Gabor, F. Trimethoprim-Loaded PLGA Nanoparticles Grafted with WGA as Potential Intravesical Therapy of Urinary Tract Infections—Studies on Adhesion to SV-HUCs Under Varying Time, pH, and Drug-Loading Conditions. ACS Omega 2020, 5, 17377–17384. [Google Scholar] [CrossRef]
  148. Sarfraz, M.; Qamar, S.; Rehman, M.U.; Tahir, M.A.; Ijaz, M.; Ahsan, A.; Asim, M.H.; Nazir, I. Nano-Formulation Based Intravesical Drug Delivery Systems: An Overview of Versatile Approaches to Improve Urinary Bladder Diseases. Pharmaceutics 2022, 14, 1909. [Google Scholar] [CrossRef] [PubMed]
  149. Kothawade, S.; Bagul, U.; Kokare, C.; Giikwad, S.; Wakure, R.; Biyani, S.; Harne, C. Formulation development of antimicrobial zinc oxide nanoparticle loaded trans dermal patch by using 2 3 factorial design. Indo Am. J. Pharm. Res. 2019, 9, 483–493. [Google Scholar]
  150. Nasrollahzadeh, M.; Ganji, F.; Taghizadeh, S.M.; Vasheghani-Farahani, E.; Mohiti-Asli, M. Drug in adhesive transdermal patch containing antibiotic-loaded solid lipid nanoparticles. J. Biosci. Bioeng. 2022, in press. [CrossRef] [PubMed]
  151. Elfaky, M.A.; Sirwi, A.; Tolba, H.H.; Shaik, R.A.; Selmi, N.M.; Alattas, A.H.; Albreki, R.S.; Alshreef, N.M.; Gad, H.A. Development, Optimization, and Antifungal Assessment of Ocular Gel Loaded With Ketoconazole Cubic Liquid Crystalline Nanoparticles. J. Pharm. Sci. 2021, 110, 2210–2220. [Google Scholar] [CrossRef]
  152. Juknius, T.; Juknienė, I.; Tamulevičius, T.; Ružauskas, M.; Pamparienė, I.; Oberauskas, V.; Jurkevičiūtė, A.; Vasiliauskas, A.; Tamulevičius, S. Preclinical Study of a Multi-Layered Antimicrobial Patch Based on Thin Nanocomposite Amorphous Diamond Like Carbon Films with Embedded Silver Nanoparticles. Materials 2020, 13, 3180. [Google Scholar] [CrossRef] [PubMed]
  153. Rayyif, S.M.I.; Mohammed, H.B.; Curuțiu, C.; Bîrcă, A.C.; Grumezescu, A.M.; Vasile, B.Ș.; Dițu, L.M.; Lazăr, V.; Chifiriuc, M.C.; Mihăescu, G.; et al. ZnO Nanoparticles-Modified Dressings to Inhibit Wound Pathogens. Materials 2021, 14, 3084. [Google Scholar] [CrossRef]
  154. Ambrogi, V.; Pietrella, D.; Donnadio, A.; Latterini, L.; Di Michele, A.; Luffarelli, I.; Ricci, M. Biocompatible alginate silica supported silver nanoparticles composite films for wound dressing with antibiofilm activity. Mater. Sci. Eng. C 2020, 112, 110863. [Google Scholar] [CrossRef]
  155. Namviriyachote, N.; Muangman, P.; Chinaroonchai, K.; Chuntrasakul, C.; Ritthidej, G.C. Polyurethane-biomacromolecule combined foam dressing containing asiaticoside: Fabrication, characterization and clinical efficacy for traumatic dermal wound treatment. Int. J. Biol. Macromol. 2020, 143, 510–520. [Google Scholar] [CrossRef]
  156. Başaran, D.D.A.; Gündüz, U.; Tezcaner, A.; Keskin, D. Topical delivery of heparin from PLGA nanoparticles entrapped in nanofibers of sericin/gelatin scaffolds for wound healing. Int. J. Pharm. 2021, 597, 120207. [Google Scholar] [CrossRef]
  157. Ahmed, M.K.; Zayed, M.A.; El-Dek, S.I.; Hady, M.A.; El Sherbiny, D.H.; Uskoković, V. Nanofibrous ε-polycaprolactone scaffolds containing Ag-doped magnetite nanoparticles: Physicochemical characterization and biological testing for wound dressing applications in vitro and in vivo. Bioact. Mater. 2021, 6, 2070–2088. [Google Scholar] [CrossRef]
  158. Ahmed, M.K.; Menazea, A.A.; Abdelghany, A.M. Blend biopolymeric nanofibrous scaffolds of cellulose acetate/ε-polycaprolactone containing metallic nanoparticles prepared by laser ablation for wound disinfection applications. Int. J. Biol. Macromol. 2020, 155, 636–644. [Google Scholar] [CrossRef] [PubMed]
  159. Saravanan, S.; Nethala, S.; Pattnaik, S.; Tripathi, A.; Moorthi, A.; Selvamurugan, N. Preparation, characterization and antimicrobial activity of a bio-composite scaffold containing chitosan/nano-hydroxyapatite/nano-silver for bone tissue engineering. Int. J. Biol. Macromol. 2011, 49, 188–193. [Google Scholar] [CrossRef] [PubMed]
  160. Spirescu, V.A.; Niculescu, A.-G.; Slave, Ș.; Bîrcă, A.C.; Dorcioman, G.; Grumezescu, V.; Holban, A.M.; Oprea, O.-C.; Vasile, B.Ș.; Grumezescu, A.M.; et al. Anti-Biofilm Coatings Based on Chitosan and Lysozyme Functionalized Magnetite Nanoparticles. Antibiotics 2021, 10, 1269. [Google Scholar] [CrossRef] [PubMed]
  161. Spirescu, V.A.; Șuhan, R.; Niculescu, A.-G.; Grumezescu, V.; Negut, I.; Holban, A.M.; Oprea, O.-C.; Bîrcă, A.C.; Vasile, B.Ș.; Grumezescu, A.M.; et al. Biofilm-Resistant Nanocoatings Based on ZnO Nanoparticles and Linalool. Nanomaterials 2021, 11, 2564. [Google Scholar] [CrossRef] [PubMed]
  162. Prodana, M.; Stoian, A.B.; Burnei, C.; Ionita, D. Innovative Coatings of Metallic Alloys Used as Bioactive Surfaces in Implantology: A Review. Coatings 2021, 11, 649. [Google Scholar] [CrossRef]
  163. Florea, D.A.; Grumezescu, V.; Bîrcă, A.C.; Vasile, B.Ș.; Mușat, M.; Chircov, C.; Stan, M.S.; Grumezescu, A.M.; Andronescu, E.; Chifiriuc, M.C. Design, Characterization, and Antibacterial Performance of MAPLE-Deposited Coatings of Magnesium Phosphate-Containing Silver Nanoparticles in Biocompatible Concentrations. Int. J. Mol. Sci. 2022, 23, 7910. [Google Scholar] [CrossRef]
  164. Olar, R.; Badea, M.; Maxim, C.; Grumezescu, A.M.; Bleotu, C.; Măruţescu, L.; Chifiriuc, M.C. Anti-biofilm Fe3O4@C18-[1,3,4]thiadiazolo[3,2-a]pyrimidin-4-ium-2-thiolate Derivative Core-shell Nanocoatings. Materials 2020, 13, 4640. [Google Scholar] [CrossRef]
Figure 1. The main types of stimuli-responsive nanoparticles (NPs) for drug delivery.
Figure 1. The main types of stimuli-responsive nanoparticles (NPs) for drug delivery.
Ijms 23 13862 g001
Figure 2. Antimicrobial nanoformulation possibilities.
Figure 2. Antimicrobial nanoformulation possibilities.
Ijms 23 13862 g002
Figure 3. Schematic representation of commonly used synthesis methods to produce magnetite nanoparticles (MNPs): (a) co-precipitation; (b) thermal decomposition; (c) sol–gel; (d) microemulsion. Reprinted with permission from [49], © Elsevier, 2022.
Figure 3. Schematic representation of commonly used synthesis methods to produce magnetite nanoparticles (MNPs): (a) co-precipitation; (b) thermal decomposition; (c) sol–gel; (d) microemulsion. Reprinted with permission from [49], © Elsevier, 2022.
Ijms 23 13862 g003
Figure 4. Schematic representation of commonly used synthesis methods to produce polymeric NPs: (a) solvent evaporation method; (b) emulsification/reverse salting-out method; (c) emulsification/solvent diffusion method; (d) nanoprecipitation method. Reprinted from an open-access source [50], adapted from [51].
Figure 4. Schematic representation of commonly used synthesis methods to produce polymeric NPs: (a) solvent evaporation method; (b) emulsification/reverse salting-out method; (c) emulsification/solvent diffusion method; (d) nanoprecipitation method. Reprinted from an open-access source [50], adapted from [51].
Ijms 23 13862 g004
Figure 5. Schematic representation of commonly used synthesis methods to produce lipid-based NPs: (a) hot high-pressure homogenization method; (b) cold high-pressure homogenization method; (c) solvent evaporation method; (d) microemulsion method. Reprinted from an open-access source [52], adapted from [53].
Figure 5. Schematic representation of commonly used synthesis methods to produce lipid-based NPs: (a) hot high-pressure homogenization method; (b) cold high-pressure homogenization method; (c) solvent evaporation method; (d) microemulsion method. Reprinted from an open-access source [52], adapted from [53].
Ijms 23 13862 g005
Figure 6. Visual representation of several nanosystems for antibacterial agents’ delivery. (a) Mesoporous iron oxide NPs loaded with ciprofloxacin. Reprinted with permission from [76], © Elsevier, 2021. (b) Polysorbate 20 micelles loaded in chitosan NPs for N′-((5-nitrofuran-2-yl)methylen)-2-benzhydrazide delivery. Reprinted with permission from [86], © Elsevier, 2020. (c) Chitosan-alginate core-shell NPs loaded with rifaximin. Reprinted with permission from [89], © Elsevier, 2021. (d) Phosphatidylcholine-chitosan liposome NPs for gentamicin delivery. Reprinted with permission from [92], © Elsevier, 2020.
Figure 6. Visual representation of several nanosystems for antibacterial agents’ delivery. (a) Mesoporous iron oxide NPs loaded with ciprofloxacin. Reprinted with permission from [76], © Elsevier, 2021. (b) Polysorbate 20 micelles loaded in chitosan NPs for N′-((5-nitrofuran-2-yl)methylen)-2-benzhydrazide delivery. Reprinted with permission from [86], © Elsevier, 2020. (c) Chitosan-alginate core-shell NPs loaded with rifaximin. Reprinted with permission from [89], © Elsevier, 2021. (d) Phosphatidylcholine-chitosan liposome NPs for gentamicin delivery. Reprinted with permission from [92], © Elsevier, 2020.
Ijms 23 13862 g006
Figure 7. Visual representation of several nanosystems for antifungal agents’ delivery. (a) Chitosan-based NPs loaded with voriconazole. Adapted from an open-access source [111]. (b) PLGA NPs loaded with farnesol. Reprinted from an open-access source [114]. (c) PLGA NPs functionalized with anti-F4/80 antibodies loaded for itraconazole delivery. Reprinted from an open-access source [116].
Figure 7. Visual representation of several nanosystems for antifungal agents’ delivery. (a) Chitosan-based NPs loaded with voriconazole. Adapted from an open-access source [111]. (b) PLGA NPs loaded with farnesol. Reprinted from an open-access source [114]. (c) PLGA NPs functionalized with anti-F4/80 antibodies loaded for itraconazole delivery. Reprinted from an open-access source [116].
Ijms 23 13862 g007
Figure 8. Visual representation of several nanosystems for antiviral agents’ delivery. (a) Amide-functionalized alginate NPs loaded with zidovudine. Reprinted with permission from [127], © Elsevier, 2018. (b) Dextran-stearic acid core shell NPs loaded with zidovudine. Reprinted with permission from [128], © Elsevier, 2018. (c) Sulfobutyl ether-β-cyclodextrin decorated chitosan nanodroplets for acyclovir delivery. Reprinted with permission from [131], © Elsevier, 2020.
Figure 8. Visual representation of several nanosystems for antiviral agents’ delivery. (a) Amide-functionalized alginate NPs loaded with zidovudine. Reprinted with permission from [127], © Elsevier, 2018. (b) Dextran-stearic acid core shell NPs loaded with zidovudine. Reprinted with permission from [128], © Elsevier, 2018. (c) Sulfobutyl ether-β-cyclodextrin decorated chitosan nanodroplets for acyclovir delivery. Reprinted with permission from [131], © Elsevier, 2020.
Ijms 23 13862 g008
Figure 9. (a) Visual representation for the fabrication of decoquinate delivery system and (b) systematic testing. Reprinted with permission from [142], © Elsevier, 2022.
Figure 9. (a) Visual representation for the fabrication of decoquinate delivery system and (b) systematic testing. Reprinted with permission from [142], © Elsevier, 2022.
Ijms 23 13862 g009
Figure 10. Overview of the possibilities of using NPs as delivery vehicles for antimicrobial applications.
Figure 10. Overview of the possibilities of using NPs as delivery vehicles for antimicrobial applications.
Ijms 23 13862 g010
Table 1. Classification of NP synthesis methods.
Table 1. Classification of NP synthesis methods.
Synthesis ApproachNature of Involved
Processes
Examples of Techniques
Top-down approachPhysical methodsBall milling, Laser ablation, Electron beam deposition, Sputtering, Aerosol spray
Bottom-up approachChemical methodsCo-precipitation, Thermal decomposition, Sol-gel, Microemulsion
Sonochemical, Hydrothermal, Microwave assisted, Chemical reduction, Electrochemical, Solvothermal
Biological methodsBacteria-based, Plant-based
Table 2. Examples of antibacterial nanoformulations.
Table 2. Examples of antibacterial nanoformulations.
NP TypeFabrication MethodPhysicochemical PropertiesAntimicrobial Agent(s)Targeted
Pathogen(s)
Ref.
Magnetite NPsCo-precipitationSize range: ~2.8–~4.7 nm
Shape: spherical
Streptomycin/
neomycin
S. aureus, P. aeruginosa[72]
Magnetite NPsCo-precipitationSize range: 10–20 nm
Average hydrodynamic diameter: 39.3 nm
Vancomycin B. subtilis, Streptococcus, E. coli[74]
Basil seed mucilage coated magnetite NPsCo-precipitationMean size: 6 nm
Specific surface area: 30.60 m2g−1
Cephalexin E. coli, S. typhimurium, S. aureus, B. cereus[75]
Mesoporous iron oxide NPsCo-precipitationAverage size: 78.34 ± 1.38 nm
Zeta potential: −18.45 ± 1.89 mV
Superficial area: 258.27 ± 7.51 m2g−1
Ciprofloxacin S. aureus[76]
Silanized mesoporous iron oxide NPsCo-precipitationAverage size: 86.32 ± 2.0 nm
Zeta potential: +5.76 ± 0.65 mV
Superficial area: 186.27 ± 6.68 m2g−1
Ciprofloxacin S. aureus[76]
CTAB-loaded mesoporous silica NPsHydrothermal methodSize range: ~100–110 nm
Shape: quasi-spherical
Silver NPsE. coli, S. aureus[73]
Chitosan NPsIonic gelationMean size: ranging from 161.90 ± 3.32 nm to 283.97 ± 4.21 nm
Zeta potential: ranging from +30.43 ± 1.08 to +21.87 ± 1.87 mV
Levofloxacin P. aeruginosa, S. aureus[82]
Chitosan NPsIonotropic gelationAverage size: 135.2 ± 3.24 nm
Shape: spherical
Gentamicin E. coli, S. aureus[83]
Chitosan NPsIonic gelationSize range: 20–80 nm
Hydrodynamic diameter: 141.4–181.6 nm
Zeta potential: ranging from +49.9 to +38.7 mV
Shape: spherical
C. zeylanicum essential oilE. coli, E. carotovora, P. fluorescens[84]
Chitosan NPsTwo-stage emulsion-ionic gelation methodMean size: ranging from 208.3 ± 44.5 to 369.4 ± 48.1 nm
Zeta potential: ranging from +30.1 ± 2.3 to +14.46 ± 0.9 mV
Nettle essential oilE. coli, S. aureus[85]
Polysorbate 20 micelles loaded in chitosan NPsIonic gelationAverage size: 321 nm
Zeta potential: +37 mV
Shape: spherical
N′-((5-nitrofuran-2-yl)methylen)-2-benzhydrazideMultidrug-resistant S. aureus[86]
Alginate-chitosan NPsIonic gelation Average hydrodynamic diameter: 380 ± 15 nm
Zeta potential: −28.5 ± 0.03
Rifampicin and ascorbic acidMSSA, MRSA[55]
Alginate-chitosan NPsCalcium ion-induced pre-gelation of alginate core and further complexation with chitosanAverage hydrodynamic diameter: 276.5 ± 42 nm
Zeta potential: −25 mV
LysMR-5S. aureus[87]
Chitosan-alginate NPsIonotropic gelationAverage size: 61.9 nm Doxycycline E. faecalis, P. mirabilis[88]
Chitosan-alginate core-shell NPsPrecipitation/coacervation methodSize range: 700–1150 nm
Zeta potential: −16.61 mV
Rifaximin E. coli, P. aeruginosa, Bacillus haynesii[89]
Chitosan-alginate NPsEmulsification and consequent electrostatic gelationAverage size: 320 nm
Zeta potential: −25 mV
Oregano oilMSSA, MRSA, E. faecalis, S. pyogenes, E. coli, P. aeruginosa, Y. enterocolitica[90]
Chitosan-polyanion NPsIonic gelation and polyelectrolyte complexation assisted by high-intensity sonicationAverage size: ranging from 130.7 to 249.2 nm
Zeta potential: ranging from +39.5 to +49.2 mV
AmpicillinS. aureus[91]
Phosphatidylcholine-chitosan liposome NPsIonic gelationAverage size: ~140 nm
Zeta potential: −19.5 mV
Gentamicin L. monocytogenes, S. aureus, P. aeruginosa, E. coli[92]
Dextran NPsIonic gelationAverage size: 18 nm
Zeta potential: −13 mV
SET-M33 peptideP. aeruginosa[93]
PLA NPs functionalized with poly-L-lysineSurfactant-free nanoprecipitationAverage hydrodynamic diameter: 162 ± 2 nm
Zeta potential: +40 ± 2 mV
Rifampicin S. aureus[94]
PLA/PVA NPsSolvent evaporation methodAverage size: 239.9 nm
Zeta potential: −29.1 mV
P. lentiscus L. var. chia essential oilE. coli, B. subtilis[95]
PLA/lecithinSolvent evaporation methodAverage size: 286.1 nm
Zeta potential: −34.5 mV
Pistacia lentiscus L. var. chia essential oilE. coli, B. subtilis[95]
PLGA NPsEmulsification solvent diffusion methodAverage size: 69.2 nm
Average hydrodynamic diameter: 224.23 ± 18.87 nm
Zeta potential: −32.1 ± 4.53 mV
Shape: spherical
Red propolis extractS. aureus, P. aeruginosa[96]
PLGA NPs functionalized with specific aptamersOil-in-water emulsification-evaporation methodAverage hydrodynamic size: 226.00 ± 5.57 nm
Zeta potential: 29.00 ± 1.18 mV
Teicoplanin S. aureus[97]
PVP-coated silver NPsChemical reductionAverage size: 9.23 ± 2.03 nm
Shape: spherical
Silver NPsS. pneumoniae, H. influenzae[98]
Abbreviations: CTAB—cetyltrimethylammonium bromide; MRSA—methicillin-resistant Staphylococcus aureus; MSSA—methicillin-sensitive Staphylococcus aureus; NP—nanoparticle; PLA—polylactic acid; PLGA—poly(lactic-co-glycolic acid); PVA—poly(vinyl alcohol); PVP—polyvinylpyrrolidone.
Table 3. Examples of antifungal nanoformulations.
Table 3. Examples of antifungal nanoformulations.
NP TypeFabrication MethodPhysicochemical PropertiesAntimicrobial Agent(s)Targeted
Pathogen(s)
Ref.
Silver NPsChemical reductionAverage size: 80 nm
Shape: spherical
Nystatin C. albicans, A. brasiliensis[109]
Silver NPsChemical reductionAverage size: 25 nm
Shape: spherical
Fluconazole C. albicans, A. brasiliensis[109]
ZnO-PEGylated NPsNanoemulsification Average size: 662.3 ± 24.7 nm
Zeta potential: −14.2 ± 0.94 mV
Amphotericin BC. albicans, C. neoformans[110]
Chitosan-based NPsIonic gelationAverage size: ranging from 167 ± 8.23 to 475 ± 15.30 nm
Zeta potential: ranging from 39 ± 2.56 to 45 ± 3.11 mV
Shape: spherical
Voriconazole C albicans[111]
Chitosan NPsIonic gelationAverage size: 35.4 nm
Zeta potential: +31 mV
Seedless V. vinifera and clotrimazoleC. albicans, A. niger[112]
Chitosan-alginate NPsEmulsification and consequent electrostatic gelationAverage size: 320 nm
Zeta potential: −25 mV
Oregano oilC albicans[90]
PLGA NPsn/rAverage size: 50 nm
Zeta potential: −25 mV
Coumarin 6 and pterostilbene A. brasiliensis[113]
PLGA NPsEmulsion evaporation methodAverage size: 140 nmFarnesol C albicans[114]
PLGA NPsNanoprecipitation and single emulsion solvent evaporation methodsAverage size: 176.96 ± 24.32 nm
Zeta potential: −24.7 ± 1.04 mV
Shape: spherical
Itraconazole C albicans[115]
PLGA NPs functionalized with anti-F4/80 antibodiesNanoemulsion Average size: 226.66 ± 13.05 nm
Zeta potential: −27.9 ± 0.26 mV
Itraconazole H. capsulatum[116]
Aptamer-functionalized PLGA-PEG NPsDouble emulsification methodAverage size: 273.9 ± 1.14 nm
Zeta potential: −20 mV
Amphotericin BC albicans[117]
PLA NPsNanoprecipitation Mean size: 188.5 nm
Zeta potential: 4.80 mV
Shape: spherical
Ketoconazole T. rubrum, T. mentagrophytes, M. gypseum, C. albicans, C. dubliniensis, C. krusei, C. parapsilosis, C. tropicalis[118]
SLNsHigh shear homogenization and ultrasonication Average size: ranging from 244.2 ± 27.2 to 493.6 ± 35.3 nm
Zeta potential: ranging from −21.6 ± 7.05 to −1.4 ± 6.84 mV
Miconazole nitrateC. albicans[119]
NLCHot emulsification methodAverage size: ranging from 213.1 ± 1.7 to 445.5 ± 8.7 nm
Zeta potential: ranging from −63.8 ± 8.7 to −93.1 ± 2.7 mV
L. sidoides essential oilC. auris[120]
Abbreviations: n/r—not reported; NLC—nanostructured lipid carriers; NP—nanoparticle; PEG—polyethylene glycol; PLA—polylactic acid; PLGA—poly(lactic-co-glycolic acid); SLN—solid lipid nanoparticle.
Table 4. Examples of antiviral nanoformulations.
Table 4. Examples of antiviral nanoformulations.
NP TypeFabrication MethodPhysicochemical
Properties
Antimicrobial Agent(s)Targeted
Pathogen(s)
Ref.
NH2-functionalized silver NPsUV irradiation and chemical reductionAverage hydrodynamic diameter: 5.0 nm
Zeta potential: 22 mV
Docetaxel SARS-CoV-2[125]
Amide-functionalized alginate NPsEmulsion solvent evaporation methodAverage size: ranging from 437 ± 2.3 to 473 ± 2.6 nm
Zeta potential: ranging from −42.16 ± 3.2 to −34.13 ± 1.61 mV
Shape: spherical
Zidovudine HIV[127]
Dextran-stearic acid core shell NPsDouble emulsion solvent evaporation methodAverage size: ranging from 356 ± 2.06 to 730 ± 1.22 nm
Zeta potential: ranging from −31.6 ± 2.12 to −20.9 ± 1.7 mV
Zidovudine HIV[128]
HPβCD NPs crosslinked with diphenyl carbonateCross-linkingAverage size: ranging from 72.47 ± 4.8 to 106.5 ± 5.6 nm
Zeta potential: ranging from −30.3 ± 4.1 to −7.77 ± 3.1 mV
Dolutegravir sodiumHIV[129]
Chitosan-g-HPβCD NPsIonic gelationAverage size: ranging from 198 ± 4.4 471.3 ± 5.3 nm
Zeta potential: ranging from 3.14 ± 2.6 to 11.5 ± 2.1 mV
Efavirenz HIV[130]
Sulfobutyl ether-β-cyclodextrin decorated chitosan nanodropletsElectrostatic interactionAverage size: 395.4 ± 12.6 nm
Zeta potential: 19.98 ± 3.02 mV
Acyclovir HSV type 2[131]
SLNs Emulsification and low-temperature solidificationAverage size: 180 ± 1.2 nm
Zeta potential: −25 mV
AcyclovirHSV[132]
SLNsSolvent emulsification evaporation and double emulsion methodsMean size: ranging from 178.7 ± 4.5 to 254.3 ± 16.6 nm
Zeta potential: ranging from 39.35 ± 1.2 to 50.80 ± 4.8 mV
Ritonavir HIV-1[133]
Abbreviations: HIV—human immunodeficiency virus; HPβCD—2-Hydroxypropyl-beta-cyclodextrin; HSV—herpes simplex virus; LUVs—large unilamellar vesicles; NP—nanoparticle; SARS-CoV-2—severe acute respiratory syndrome coronavirus 2; SLN—solid lipid nanoparticle.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mercan, D.-A.; Niculescu, A.-G.; Grumezescu, A.M. Nanoparticles for Antimicrobial Agents Delivery—An Up-to-Date Review. Int. J. Mol. Sci. 2022, 23, 13862. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232213862

AMA Style

Mercan D-A, Niculescu A-G, Grumezescu AM. Nanoparticles for Antimicrobial Agents Delivery—An Up-to-Date Review. International Journal of Molecular Sciences. 2022; 23(22):13862. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232213862

Chicago/Turabian Style

Mercan, Doina-Antonia, Adelina-Gabriela Niculescu, and Alexandru Mihai Grumezescu. 2022. "Nanoparticles for Antimicrobial Agents Delivery—An Up-to-Date Review" International Journal of Molecular Sciences 23, no. 22: 13862. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232213862

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