Next Article in Journal
Novel Megaplasmid Driving NDM-1-Mediated Carbapenem Resistance in Klebsiella pneumoniae ST1588 in South America
Next Article in Special Issue
Silver Nanoparticle-Based Combinations with Antimicrobial Agents against Antimicrobial-Resistant Clinical Isolates
Previous Article in Journal
Antiprotozoal Activity of Thymoquinone (2-Isopropyl-5-methyl-1,4-benzoquinone) for the Treatment of Leishmania major-Induced Leishmaniasis: In Silico and In Vitro Studies
Previous Article in Special Issue
Antivirulence Agent as an Adjuvant of β-Lactam Antibiotics in Treating Staphylococcal Infections
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nickel Nanoparticles: Applications and Antimicrobial Role against Methicillin-Resistant Staphylococcus aureus Infections

1
Noncommunicable Diseases Research Center, Fasa University of Medical Sciences, Fasa 7461686688, Iran
2
Department of Medical Laboratory Techniques, Faculty of Health and Medical Techniques, Imam Ja’afar Al-Sadiq University, Al Muthanna 9647555434, Iraq
3
Department of pharmacy, Al-Rasheed University College, Baghdad 9643675544, Iraq
4
Friedrich-Loeffler-Institut (Federal Research Institute for Animal Health), Institute of Bacterial Infections and Zoonoses, Naumburger Str. 96a, 07743 Jena, Germany
5
Animal Health Research Institute, Agriculture Research Center (ARC), Giza 12618, Egypt
*
Authors to whom correspondence should be addressed.
Antibiotics 2022, 11(9), 1208; https://doi.org/10.3390/antibiotics11091208
Submission received: 16 August 2022 / Revised: 30 August 2022 / Accepted: 2 September 2022 / Published: 7 September 2022

Abstract

:
Methicillin-resistant Staphylococcus aureus (MRSA) has evolved vast antibiotic resistance. These strains contain numerous virulence factors facilitating the development of severe infections. Considering the costs, side effects, and time duration needed for the synthesis of novel drugs, seeking efficient alternative approaches for the eradication of drug-resistant bacterial agents seems to be an unmet requirement. Nickel nanoparticles (NiNPs) have been applied as prognostic and therapeutic cheap agents to various aspects of biomedical sciences. Their antibacterial effects are exerted via the disruption of the cell membrane, the deformation of proteins, and the inhibition of DNA replication. NiNPs proper traits include high-level chemical stability and binding affinity, ferromagnetic properties, ecofriendliness, and cost-effectiveness. They have outlined pleomorphic and cubic structures. The combined application of NiNPs with CuO, ZnO, and CdO has enhanced their anti-MRSA effects. The NiNPs at an approximate size of around 50 nm have exerted efficient anti-MRSA effects, particularly at higher concentrations. NiNPs have conferred higher antibacterial effects against MRSA than other nosocomial bacterial pathogens. The application of green synthesis and low-cost materials such as albumin and chitosan enhance the efficacy of NPs for therapeutic purposes.

1. Background

Staphylococcus (S.) aureus is a ubiquitous pathogenic bacterium. Methicillin-resistant S. aureus (MRSA) is among the leading causes of nosocomial pathogens which decipher the multidrug-resistance (MDR) phenotype [1,2,3]. MRSA employs various mechanisms to resist drugs, such as cell wall thickening, efflux of compounds, enzymatic destruction, and target variation [4,5,6].
MRSA infections encompass a wide range of manifestations and are a major life-threatening priority worldwide [7,8]. MRSA is resistant to β-lactam antibiotics. These strains carry a modified penicillin-binding protein (PBP) known as PBP-2a (encoded by either mecA or mecC gene) inherently exhibiting resistance to various β-lactams such as oxacillin, methicillin, and cefoxitin [9]. MRSA is classified based on genotypic diversity and according to various staphylococcal cassette chromosome mec (SCCmec) types including SCCmec (I-XIII) as substantial markers of MRSA epidemiology. The SCCmec gene cassette carries the mecA or mecC genes in addition to other genes associated with aminoglycosides, macrolides, and fluoroquinolones resistance [10]. MRSA is classified into three main categories including healthcare (HA-MRSA), community (CA-MRSA), and livestock-associated (LA-MRSA) strains each with special virulence and resistance patterns [7,10]. HA-MRSA is responsible for nosocomial infections globally and usually carries SCCmec types I, II, or III. CA-MRSA has been recorded in patients without or by negligible contact with healthcare settings. These strains usually carry SCCmec elements IV, V, and pvl genes. The latter is associated with the Panton–Valentine leukocidin. The LA-MRSA isolates are mainly associated with livestock origins and usually carry SCCmec IVa and SCCmec V elements. Life-threatening infections caused by LA-MRSA have been previously recorded, highlighting the possibility of zoonotic risk [11,12]. In humans, MRSA can cause severe pyogenic infections of the skin and soft tissue, endocarditis, septic arthritis, pneumonia, osteomyelitis, and otitis media [9,13].

2. Pathogenicity of MRSA

Although the colonization of S. aureus on host surfaces is not harmful, overcoming the host’s innate immunity leads to invasive deep infections. MRSA causes various cutaneous and deep infections such as folliculitis, impetigo, cutaneous abscesses, pyomyositis, necrotizing pneumonia, and fasciitis [14]. HA-MRSA causes implant or surgical site and catheter-associated infections. Bacteremia-related infections include disseminated infections such as descending urinary tract infections, endocarditis, and osteomyelitis. Thereby, the eradication of the bacterium is a concern considering recurrent infections. The bacterium virulence regulation is exerted by a set of global regulatory circuits (two-component systems, TCS, accessory gene regulatory, Agr, and quorum-sensing, QS) which affect gene expression following environmental signals. Additionally, S. aureus responds to internal stimuli in the form of QS autoinduced signaling. AgrBDCA regulates RNA effector RNAIII [15,16].
Biofilm formation is an indispensable mechanism of resistance to antimicrobials and for the environment of host responses. The polysaccharide intercellular adhesin (PIA) mediates the bacterial binding to host cells. Some other major surface adhesin proteins include fibrinogen binding protein (FnBP) A and B, surface binding protein A (Spa), cell wall-anchored proteins (CWP), clumping factors (Clfs) A and B, and surface binding protein (SasG) [17,18,19,20].
Following the bacterial attachment to the cells and colonization, pathogenesis is initiated and developed via the production of toxins, exoenzymes such as exfoliative toxins (ETs), Panton–Valentine leukocidin (PVL), toxic shock syndrome toxin1 (TSST1), phenol-soluble modulins, leukotoxin and haemolysin, lipases, proteases and nucleases, and immunomodulators or immune evasion gene clusters (IEC1 and IEC2) [21].
Phenol-soluble modulins, leukotoxin, and haemolysin are known as pore-forming toxins which lyse the host cells. α-, β-, γ-, and δ-hemolysins of S. aureus cause the lysis of erythrocytes, epithelial and endothelial cells, monocytes, damage of the epithelium, and induction of apoptosis. Leukotoxins or PVL target and destroy white blood cells such as macrophages, monocytes, and neutrophils. These include LukDE, LukAB, LukS-PV, and LukF-PV. PVL is associated with soft tissue and skin infections in both MSSA and MRSA [22,23].
Phenol-soluble modulins (PSMs) including PSMα1–PSMα4 play a substantial role in bacterial pathogenesis via cell lysis, inflammation, immune regulation, and biofilm formation or detachment [24]. Exfoliative toxins (ETA-ETD) cause staphylococcal scaled skin syndrome (SSSS) which is associated with dehydration, loss of superficial skin layers, and secondary infections which are not significantly different from MSSA and MRSA [25]. Staphylococcal enterotoxins (SEs) and TSST-1 act as superantigens which are T-cells mitogens. SEs cause food poisoning and gastrointestinal problems such as emesis. TSST1 causes the release of extraordinary amounts of pro-inflammatory cytokines [26]. In addition to humans, MRSA has long been recognized globally to colonize numerous wild and domesticated livestock animals and develop infections [27,28,29,30,31]. The widespread distribution of MRSA among livestock is largely due to the indiscriminate prescription/consumption of antimicrobials for animal breeding or agricultural activities. For example, it has impacted more than 40% of pig farms, 20% of cattle farms, and 20% to 90% of turkey farms in Germany. Numerous studies have shown that there is a high risk of MRSA colonization and infection in humans who come into contact with livestock [32,33]. MRSA is repeatedly recorded in dairy farms as a cause of mastitis with failure in elimination due to its resistance against β-lactam antibiotics employed for related infections [34].
MRSA strains employ various virulence factors to invade the host and develop resistance [6]. The development of resistance to last-resort antimicrobial treatments such as glycopeptides (vancomycin and teicoplanin) is a crisis in the eradication of vancomycin-intermediate S. aureus (VISA) or vancomycin-resistant S. aureus (VRSA) [17,35]. For skin infections (impetigo) of MRSA, mupirocin and fusidic acid or alternative 1% hydrogen peroxide (H2O2) are recommended. For abscesses, in conditions of neutropenia, cell-mediated immunity deficiency, or severe infection, there is a need for co-trimoxazole or clindamycin treatment. For cellulitis and soft tissue infections, glycopeptides, tigecycline, and linezolid are useful. Susceptibility testing is strongly recommended prior to the utilization of antimicrobial drugs. Antimicrobial therapy necessity should be evaluated when infections are observed due to costs and risk of resistance evolution. The term “superbug” is defined as a strain developing vast antibiotic resistance [36].
Combination therapies have demonstrated substantial bactericidal effects against MRSA. The combination of carbapenems with linezolid, and each of imipenem/fosfomycin/gentamicin/oxacillin/rifampin and daptomycin have deciphered acceptable anti-MRSA effects [37,38,39]. A β-lactams and daptomycin combination facilitates its binding and mitigates resistance development [40,41]. Additionally, the combination of gentamicin/β-lactams/rifampin and vancomycin has outlined significant effects [42,43]. However, the combination of vancomycin and β-lactams has had nephrotoxicity according to clinical trials [44,45,46].
Natural combination therapies using Polysporin and Neosporin peptides have also exerted anti-MRSA effects [47]. Major chemotherapy approaches to combat MRSA have included telavancin, teicoplanin, ceftaroline, vancomycin, and oxazolidinones [48,49]. The combination of β-lactam and either arbekacin or vancomycin is recommended against MRSA. Notably, granulocyte counts are also recommended for antibiotics consumption [50,51]. Ceftaroline fosamil has exhibited substantial anti-MRSA effects [52,53,54]. Daptomycin and ceftaroline have exhibited higher bactericidal effects and linezolid has exerted strong effects on bacteremia.
As the first clinically applied anti-MRSA oxazolidinones, linezolid is a promising antibiotic with low resistance development to MRSA. However, epidemiological data have unraveled that linezolid resistance among MRSA isolates is in enhancement alongside novel antimicrobials quinupristin/dalfopristin (Q/D), daptomycin, and tigecycline [55,56]. Major drawbacks in the conventional approaches to MRSA eradication using antibiotics include non-specific effects which kill human beneficial bacteria, adverse/toxic effects, drug resistance development, and the high cost and time-consuming development of new drugs.

3. Nanoparticle Applications to Combat MRSA

The efficient, non-toxic, proper, accurate, and cost-effective eradication of MRSA infections has been recently achieved through the application of nanocarriers or nanoparticles (NPs) [57,58]. NPs have a size range of 1 to 100 nm. The physicochemical properties of NPs and lower costs are gaining attention for use as antimicrobial agents [59,60,61,62]. NPs have exhibited antimicrobial effects, particularly those synthesized using green methods [6]. NPs can also be used for the detection of MRSA [63]. The main mechanisms of NP antibacterial effects include an impairment in metabolism or bacterial integrity (CuNPs), replication and transcription disruption, tRNA, ATPases, membrane-bound enzymes and biofilm inhibition, protein denaturation (AgNPs), and reactive oxygen species (ROS) production [64,65,66]. Various NPs such as silver (Ag), gold (Au), and lower-cost NPs such as nickel (Ni), titanium-oxide, (TiO), zinc oxide (ZnO), silica (SiO2), and bismuth (Bi) NPs have deciphered efficient bactericidal effects against MRSA in vitro and in vivo [67,68,69,70,71]. Various methods of NP delivery to cells for antibacterial activity include polymeric NPs, liposomes, carbon NPs, and metal or metal-oxide NPs. ZnO could eliminate MRSA skin infection at 1875 mg/mL possibly via amino acid synthesis inhibition [70]. TiO2 NPs produce free radicals which kill MRSA [72]. Cu-doped ZnO nanorods have exhibited more potential effects than that of ZnO singly [73]. Cefotax-based magnetic NPs have exhibited promising anti-MRSA effects against isolates originating from livestock and dairy sources [58]. Interestingly, S,N-GQDs/NiO nanocomposites have exerted extraordinary anti-S. aureus effects in vitro with minimum inhibitory (MIC) and bactericidal (MBC) concentrations values of 0.4 and 0.8 mg/mL, respectively [74]. Nickel oxide nanoparticles (NiO NPs) exhibit promising traits such as biocompatibility, thermal and chemical stability, and interesting optical characteristics. The development of NiO nanocomposites has improved their bactericidal effects. The simple synthesis of a CdO–NiO–ZnO nanocomposite using the microwave method also exhibited antibacterial effects against Gram-positive and Gram-negative species [75,76]. In addition, NiOCuO-10%RGO also demonstrated substantial antibacterial activity [74]. NPs can specifically carry drugs and enter into cells via endocytosis. A few NPs for the eradication of MRSA have been used for clinical trials. PLGA-rifampicin NPs have decreased the MIC against MRSA from 0.0008 to 0.002 µg/mL [77]. Vancomycin-loaded hydroxyapatite increased the vancomycin release and improved bone regeneration [78]. The combination of levofloxacin and AgNPs unraveled a synergistic effect with 0.5 and 10 µM against control strains and MRSA, respectively [79]. The application of liposome-albumin-vancomycin has decreased toxicity and improved the anti-MRSA activity [80]. The liposomal formulation of amphotericin-b (AmBisome®) has also enhanced anti-MRSA efficiency and mitigated nephrotoxicity [81]. Various NP-based drug delivery systems used for skin treatment have included solid lipid nanoparticles, nanostructured lipid carriers, liposomes, niosomes, and nanoemulsions [82].
It was revealed that Chitosan/Gold Nanoparticle/Graphene Oxide could separate, identify, and eradicate MRSA superbugs within water contaminants. Chitosan has a positive charge and can trap these strains. In a study by Gupta, engineered polymer NPs exerted anti-biofilm effects against MRSA strains at non-toxic levels [83].
Mesoporous silica nanoparticles (MSNs) carrying enzymes have exhibited anti-biofilm (dispersal) effects against S. aureus, decreasing the bacterial cells efficiently after 24 and 48 h [84], respectively.

4. Advantages and Disadvantages of Nanoparticles

The reduced size of particles provides higher surface accessibility for ligand binding, easy and size adoptive production, rapid penetration and accumulation into cells, absorption and carriage of various compounds, higher drug deposition rates, and a lower rate of clearance from the body [85,86]. The facilitated and rapid penetration and long-lasting deposition of NP drugs are promising in regard to the control of MRSA infections. However, permeation enhancers may disrupt the lipids of the stratum corneum. Further challenges are considerable for various formulations of NPs such as drug expelling, encapsulation, physical stability, controlled release, transdermal delivery challenges (lipid carriers), viscosity differences, skin penetration, costs and large-scale production (niosomes), agglomeration, more frequent doses, systemic adsorption (nanocrystals), and risk of toxicity (polymeric NPs) [87,88]. It is worth mentioning that interactions of NPs with the human body enzymes and cytosol may affect host cells, as demonstrated for CuNPs compared to NiNPs, which necessitates the further evaluation of corresponding effects [89].
This review assessed published data between 2015 and 2022 discussing the involvement of NiNP or NiO NP applications in the treatment of MRSA infections. Found publications were screened based on their titles, abstracts, and full-text availability. English language texts were included herein. The flowchart of publications selection and content of this review is illustrated in Figure 1.

5. Nanoparticle Features and Synthesis

The diverse applications of NPs in medicine, environment, agriculture, catalysis, biosensing, and cancer theranostics have attracted interest for their synthesis using chemical, physical, and biological methods. NPs are synthesized through various approaches according to corresponding features of morphology, size, stability, and biocompatibility. These are commonly synthesized by microemulsion, hydrothermal, coprecipitation, sol-gel, ball milling, and biological methods [90]. In the physical method of synthesis, NP size and morphology are difficult to adjust, whereas, in chemical methods, it is possible by alterations in reaction conditions [91]. A hydrothermal method is the most suitable chemical method for NP synthesis owing to its easy and versatile nature and controllable synthesis of high crystalline and homogenous particles [92]. Noticeably, chemical methods leave adverse health and environmental effects and also consume large energy. Considering these, green synthesis (microbial and plant extracts) is promising in terms of considerably lower cost, pollution and toxicity, and health and environmental benefits [93]. In spite of seasonal and geographical limitations of herbs, low purity and amounts, green synthesis is preferred for the chemical synthesis of NPs [94].

6. Importance of Nickel and Nickel-Oxide Nanoparticles

Nickel nanoparticles (NiNPs) including magnetic metal intermediate-cost particles have been studied vastly thanks to their myriad applications such as for magnetic sensors [95], memory devices [96], and for biomolecular separation [97]. The eventual proficiency of each material or NP reflects and depends on the structure, shape, size, and purity of NiNPs or derived materials. Nickel oxide (NiO) is an inevitably crucial part of today’s nanotechnology and intermediate-cost metal oxide is a cubical lattice structure [98]. The extraordinary chemical stability, high binding affinity, and ferromagnetic properties of NiNPs provide an indispensable field of study which includes their synthesis and application. They are applied mostly because of cost-effectiveness catalysts considering vast natural resource existence and driving reactions by alternate routes. These NPs contain numerous biomedical usages, including cell isolation, medicine delivery, magnetic resonance imaging, biomedical diagnostics, and more [99]. Considerable antibacterial attributes have been reported against Bacillus subtilis, Pseudomonas aeruginosa, Klebsiella pneumoniae, S. aureus, and many other microorganisms [100] (Figure 2). In addition, various forms of NP delivery have been represented in Figure 3.

7. Mechanism of Action of Nickel Nanoparticles

Antibacterial activities of NiNPs are related to nickel ion content which interpenetrate the bacterial cell and reach the surface of the bacterial cell membrane and intracellular milieu. This influx of nickel cations destroys organelles like ribosomes and affects bacterial metabolism. A study of the literature unravels how all of this happens, due to electrostatic attraction and the use of negatively charged intercellular microbial cells and positively charged nickel ions [103,104]. Nanoparticles exhibit great surface activity due to the large surface-to-volume proportion. Exposure of E. coli to NiNPs disrupts membrane morphology and transport. Nickel’s high affinity to sulfur- and phosphor-containing components such as DNA and proteins disrupts DNA replication and causes protein deformation [105].

8. Recent Data Regarding NiONP Effects against MRSA

Haghshenas, Leila, et al. [106] appraised the antibacterial efficacy of gold (AuNPs) and NiNPs separately and in combination with E. coli and S. aurous in milk. Transmission electron microscopy (TEM) and Uv-vis spectroscopy were used to measure the size and shape of these NPs. A broth microdilution procedure was used to assess the minimum inhibitory concentration (MIC) of the NPs. Then, the NPs’ effect on milk was investigated individually and in combination at 25 °C and 50 °C, respectively. These results displayed that AuNPs and NiNPs with mean sizes of 10 nm and 50 nm had a favorable bactericidal effect (p ≤ 0.05) against E. coli and S. aureus. MICs of AuNPs and NiNPs against S. aurous included 0.42 and 0.21 μg mL−1 and 0.84 and 0.42 μg mL−1 for E. coli, respectively. The data of this study revealed that the antibacterial effect of NPs in milk is temperature- and dose-dependent, and the greatest degree of effect was observed in the combined concentration at 50 °C. NiNPs were found to have a stronger antibacterial effect than AuNPs. Accordingly, S. aureus was more sensitive to NPs than E. coli. AuNPs and NiNPs were proposed as effective candidates for combating pathogens in the food system. This may be due to the charge of the cell wall surface which has a higher affinity for NiNPs.
Haider, Ali, et al. [107] developed a green plant (ginger and garlic) to replace the bactericidal and synthetic catalyst in the textile industry, reducing NiO NPs. The synthesis of NPs was corroborated using X-ray diffraction (XRD) and ultra-violet visible spectroscopy (UV-Vis), which has a potent absorption at 350 nm with a size between 16 to 52 nm for ginger and 11 to 59 nm for garlic. Scanning and transmission electron microscopy (SEM and TEM, respectively) confirmed pleomorphism with cubic and spherical NPs. In addition, the exact amounts of garlic and ginger extract (1:3.6 mL) combined for the synthesis of NiO NPs were effectively confirmed using Fourier transform infrared spectroscopy (FTIR). Garlic-reduced phytochemical NPs effectively degraded bactericidal activity against MRSA at higher concentrations (0.5, 1.0 mg/50 µL) as well as methylene blue (MB) dye. Finally, green synthesized NiO NPs were eminent factors for resolving drug resistance as well as eco-friendly catalytic factors that may be selected on an industrial measure.
In a study, NiO NPs were synthesized using photolysis and their anti-biofilm property was evaluated. XRD studies showed the presence of NiO NPs with high crystallinity. NiO particle size ranged from 13 to 31 nm. There were 42 samples of medical waste from various hospitals in Baghdad from 2 October to 12 October 2020. Isolation outcomes were recorded from 15 isolates of S. aureus. The results of both methods (well propagation method and PCR) showed that the percentages of MRSA in these two methods were 53.3% and 73.4%, respectively. The outcomes of MRSA biofilm formation showed that only four isolates (36.3%) were not able to produce biofilm using the plate microtiter method. On the other hand, other isolates (63.7%) were able to produce biofilms. Antimicrobial activity of various concentrations (10–100 µg/mL) ranged from 0–13 mm inhibition zones. The MIC concentration was 265 µg/mL (63.7%) from seven MRSA isolates and 530 µg/mL (36.4%) from four isolates. The results outlined that the hemolytic activity of 2.38%, 2.23%, 2.41%, and 2.69% corresponded to 0, 0.01, 0.1 and 1 mg/mL of NiO NPs, respectively [108].
In another study, NiO with sulfur, and nitrogen co-doped-graphene quantum dots-decorated NiO nanocomposites (S, N-GQDs/NiO) were synthesized using an easy hydrothermal technique. Then, their antibacterial activity against E. coli, S. aureus, P. aeruginosa, and MRSA was assessed. The prepared S,N-GQDs/NiO nanocomposites conferred the highest antibacterial effect against S. aureus among these species. The S, N-GQDs/NiO nanocomposite NPs exerted the highest effect against S. aureus (17 mm) in the disk diffusion method. The attendance of graphene quantum spots in S,N-GQDs/NiO nanocomposites comforts the ROS mechanism that leads to antibacterial activity [74].
A study inferred the structural and morphological properties of NiO NPs synthesized by a novel, convenient, eco-friendly, and cost-effective ash-supported approach [109]. The antibacterial property of NiNPs was attended using pathogenic strains including Gram-negative and Gram-positive bacteria by the cup and well diffusion method. Raman analysis confirmed the formation of pure NiO cubic phase at an annealing temperature of 700 °C. The SEM illuminated the particle morphology and size of 40 to 90 nm. The antibacterial activity of the NiO NPs was studied using three Gram-negative bacteria (Salmonella typhi, Pseudomonas aeruginosa, E. coli) and one Gram-positive bacterium (MRSA). Accordingly, MRSA was sensitive to NiONPs with a diameter of 26 mm inhibition zone [110].
NiO NPs with a green approach were synthesized using Eucalyptus globulus leaf extract and their bactericidal traits were measured. The synthesized NiO NPs were pleomorphic and ranged in size from 10 to 20 nm. The X-ray diffraction (XRD) analysis outlined an average size of NiO NPs as 19 nm. The bioactivity experiment showed the antibacterial and anti-biofilm properties of NiO NPs against extended spectrum beta lactamase (ESβL (+))-producing E. coli, P. aeruginosa, and S. aureus. The growth inhibition method showed a time-dependent decrease in the concentration of NiO NPs in the survival of treated cells. The inhibition of NiO NPs-induced biofilms was demonstrated using a sharp increase in the red fluorescence characteristic of propidium iodide (PI) when SEM illustrations of NiO NPs-treated cells were reduced and distorted by obvious depressions/indentations. In general, E. globulus leaf extract can be safely used to synthesize eco-friendly NiO NPs with the potential for the elimination of infections affecting human health (Table 1) [111].
In a study, Superparamagnetic Ni@2D-MoS2 nanosheets exerted anti-biofilm and antibacterial effects against MRSA and E. faecalis demonstrating the selective removal of these infections [112].

9. Future Prospects

Metal NPs such as Au, Ag, Zn, and Cu are believed to be toxic to eukaryotic cells at high levels. Hence proper and target-specific delivery is crucial. Since the stability of metal or metal oxide NPs is not ensured and even PEGylated NPs are removed from the bloodstream, their use for topical infections is more promising. Low-toxic or non-toxic materials such as albumin and chitosan are considerable for the carriage of drugs [113,114]. Considering the low costs of NiNPs, more investigations are needed to assess their efficiency in vivo and in clinical trials. In addition, combination therapies using NiNPs can be helpful for the eradication of extreme drug-resistant strains. As metal NPs have exerted acceptable anti-MRSA effects, the interactions of NPs with human body enzymes and cell structures need to be evaluated by further experiments to uncover their possible detrimental effects on the host [115].

10. Conclusions

Considering the drawbacks in MRSA infection eradication due to vast antibiotic resistance and the expression of numerous virulence factors to develop severe infections, particularly among vulnerable individuals, novel treatment choices are warranted. In addition, MRSA develops resistance to last-resort antibiotics during a time span, owing to a non-adherence to the prescription and consumption of antibiotics. Due to the high costs, side effects, and time duration needed for the synthesis of synthesized drugs, seeking efficient alternative approaches for the eradication of drug-resistant bacterial agents seems to be an unmet requirement. NPs have been applied as prognostic and therapeutic agents in various aspects of biomedical sciences. Lower-cost NPs such as NiNPs and their formulations have been utilized as antibacterial agents against MRSA infections. The application of green synthesis and low-cost materials such as albumin and chitosan enhance the efficacy of NPs for therapeutic purposes. NiNPs’ precise mechanism of action can be predicted and understood via in silico, in vitro, and in vivo studies.

Author Contributions

E.Z., A.G. and M.M. collected published papers and wrote the manuscript draft, A.A.M. conceptualized the study, H.T.A., E.G., M.E. and B.P. edited and approved the work. 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

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

The authors wish to thank the Noncommunicable Diseases Research Center, Fasa University of Medical Sciences.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

MRSAmethicillin-resistant Staphylococcus aureus
HA-MRSAhealthcare-associated MRSA
CA-MRSAcommunity-associated MRSA
PBPpenicillin-binding protein
WHOworld health organization
VISAvancomycin-intermediate S. aureus
VRSAvancomycin-resistant S. aureus
MSNsMesoporous silica nanoparticles
ESβLsExtended spectrum beta lactamases
FTIRFourier transform infrared spectroscopy
MBMethylene blue
MICMinimum inhibitory concentration
NiONPsNickel oxide nanoparticles
RBCsred blood cells
PIPropidium iodide
RBCsRed blood cells
ROSReactive oxygen species
SEMScanning electron microscopy
TEMTransmission electron microscopy
UV-VisUltra-violet visible spectroscopy
XRDX-ray diffraction

References

  1. Parente, D.M.; Cunha, C.; Mylonakis, E.; Timbrook, T.T. The Clinical Utility of Methicillin-Resistant Staphylococcus aureus (MRSA) Nasal Screening to Rule Out MRSA Pneumonia: A Diagnostic Meta-analysis With Antimicrobial Stewardship Implications. Clin. Infect. Dis. 2018, 67, 1–7. [Google Scholar] [CrossRef]
  2. Wong, J.W.; Ip, M.; Tang, A.; Wei, V.W.; Wong, S.Y.; Riley, S.; Read, J.M.; Kwok, K.O. Prevalence and risk factors of community-associated methicillin-resistant Staphylococcus aureus carriage in Asia-Pacific region from 2000 to 2016: A systematic review and meta-analysis. Clin. Epidemiol. 2018, 10, 1489–1501. [Google Scholar] [CrossRef]
  3. Zarenezhad, E.; Mosslemin, M.H.; Alborzi, A.; Anaraki-Ardakani, H.; Shams, N.; Khoshnood, M.M.; Zarenezhad, A. Efficient synthesis of 3, 4-dihydro-1 H-quinoxalin-2-ones and 1 H-quinolin-2-ones and evaluation of their anti-bacterial activity. J. Chem. Res. 2014, 38, 337–340. [Google Scholar] [CrossRef]
  4. Vestergaard, M.; Frees, D.; Ingmer, H. Antibiotic Resistance and the MRSA Problem. Microbiol. Spectr. 2019, 7, 18. [Google Scholar] [CrossRef]
  5. Mlynarczyk-Bonikowska, B.; Kowalewski, C.; Krolak-Ulinska, A.; Marusza, W. Molecular Mechanisms of Drug Resistance in Staphylococcus aureus. Int. J. Mol. Sci. 2022, 23, 8088. [Google Scholar] [CrossRef]
  6. Liu, W.-T.; Chen, E.-Z.; Yang, L.; Peng, C.; Wang, Q.; Xu, Z.; Chen, D.-Q. Emerging resistance mechanisms for 4 types of common anti-MRSA antibiotics in Staphylococcus aureus: A comprehensive review. Microb. Pathog. 2021, 156, 104915. [Google Scholar] [CrossRef]
  7. Nandhini, P.; Kumar, P.; Mickymaray, S.; Alothaim, A.S.; Somasundaram, J.; Rajan, M. Recent Developments in Methicillin-Resistant Staphylococcus aureus (MRSA) Treatment: A Review. Antibiotics 2022, 11, 606. [Google Scholar] [CrossRef]
  8. Baede, V.O.; David, M.Z.; Andrasevic, A.T.; Blanc, D.S.; Borg, M.; Brennan, G.; Catry, B.; Chabaud, A.; Empel, J.; Enger, H.; et al. MRSA surveillance programmes worldwide: Moving towards a harmonised international approach. Int. J. Antimicrob. Agents 2022, 59, 106538. [Google Scholar] [CrossRef] [PubMed]
  9. Lee, A.S.; De Lencastre, H.; Garau, J.; Kluytmans, J.; Malhotra-Kumar, S.; Peschel, A.; Harbarth, S. Methicillin-resistant Staphylococcus aureus. Nat. Rev. Dis. Primers 2018, 4, 18033. [Google Scholar] [CrossRef] [PubMed]
  10. Schnitt, A.; Tenhagen, B.-A. Risk Factors for the Occurrence of Methicillin-Resistant Staphylococcus aureus in Dairy Herds: An Update. Foodborne Pathog. Dis. 2020, 17, 585–596. [Google Scholar] [CrossRef] [Green Version]
  11. Krukowski, H.; Bakuła, Z.; Iskra, M.; Olender, A.; Bis-Wencel, H.; Jagielski, T. The first outbreak of methicillin-resistant Staphylococcus aureus in dairy cattle in Poland with evidence of on-farm and intrahousehold transmission. J. Dairy Sci. 2020, 103, 10577–10584. [Google Scholar] [CrossRef]
  12. Schnitt, A.; Lienen, T.; Wichmann-Schauer, H.; Cuny, C.; Tenhagen, B.-A. The occurrence and distribution of livestock-associated methicillin-resistant Staphylococcus aureus ST398 on German dairy farms. J. Dairy Sci. 2020, 103, 11806–11819. [Google Scholar] [CrossRef] [PubMed]
  13. Turner, N.A.; Sharma-Kuinkel, B.K.; Maskarinec, S.A.; Eichenberger, E.M.; Shah, P.P.; Carugati, M.; Holland, T.L.; Fowler, V.G., Jr. Methicillin-resistant Staphylococcus aureus: An overview of basic and clinical research. Nat. Rev. Microbiol. 2019, 17, 203–218. [Google Scholar] [CrossRef] [PubMed]
  14. Hanberger, H.; Walther, S.; Leone, M.; Barie, P.S.; Rello, J.; Lipman, J.; Marshall, J.C.; Anzueto, A.; Sakr, Y.; Pickkers, P.; et al. Increased mortality associated with meticillin-resistant Staphylococcus aureus (MRSA) infection in the Intensive Care Unit: Results from the EPIC II study. Int. J. Antimicrob. Agents 2011, 38, 331–335. [Google Scholar] [CrossRef]
  15. Queck, S.Y.; Jameson-Lee, M.; Villaruz, A.E.; Bach, T.-H.L.; Khan, B.A.; Sturdevant, D.E.; Ricklefs, S.M.; Li, M.; Otto, M. RNAIII-Independent Target Gene Control by the agr Quorum-Sensing System: Insight into the Evolution of Virulence Regulation in Staphylococcus aureus. Mol. Cell 2008, 32, 150–158. [Google Scholar] [CrossRef]
  16. Mahdally, N.H.; George, R.F.; Kashef, M.T.; Al-Ghobashy, M.; Murad, F.E.; Attia, A.S. Staquorsin: A Novel Staphylococcus aureus Agr-Mediated Quorum Sensing Inhibitor Impairing Virulence in vivo Without Notable Resistance Development. Front. Microbiol. 2021, 12, 700494. [Google Scholar] [CrossRef] [PubMed]
  17. Algammal, A.M.; Hetta, H.F.; Elkelish, A.; Alkhalifah, D.H.H.; Hozzein, W.N.; Batiha, G.E.-S.; El Nahhas, N.; Mabrok, M.A. Methicillin-Resistant Staphylococcus aureus (MRSA): One Health Perspective Approach to the Bacterium Epidemiology, Virulence Factors, Antibiotic-Resistance, and Zoonotic Impact. Infect. Drug Resist. 2020, 13, 3255–3265. [Google Scholar] [CrossRef]
  18. Cheung, G.Y.C.; Bae, J.S.; Liu, R.; Hunt, R.L.; Zheng, Y.; Otto, M. Bacterial virulence plays a crucial role in MRSA sepsis. PLoS Pathog. 2021, 17, e1009369. [Google Scholar] [CrossRef]
  19. Ghasemian, A.; Peerayeh, S.N.; Bakhshi, B.; Mirzaee, M. The Microbial Surface Components Recognizing Adhesive Matrix Molecules (MSCRAMMs) Genes among Clinical Isolates of Staphylococcus aureus from Hospitalized Children. Iran. J. Pathol. 2015, 10, 258–264. [Google Scholar]
  20. Ghasemian, A.; Peerayeh, S.N.; Bakhshi, B.; Mirzaee, M. Comparison of Biofilm Formation between Methicillin-Resistant and Methicillin-Susceptible Isolates of Staphylococcus aureus. Iran. Biomed. J. 2016, 20, 175–181. [Google Scholar] [CrossRef] [PubMed]
  21. Otto, M. MRSA virulence and spread. Cell. Microbiol. 2012, 14, 1513–1521. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Ahmad-Mansour, N.; Loubet, P.; Pouget, C.; Dunyach-Remy, C.; Sotto, A.; Lavigne, J.-P.; Molle, V. Staphylococcus aureus Toxins: An Update on Their Pathogenic Properties and Potential Treatments. Toxins 2021, 13, 677. [Google Scholar] [CrossRef]
  23. Kong, C.; Neoh, H.-M.; Nathan, S. Targeting Staphylococcus aureus Toxins: A Potential form of Anti-Virulence Therapy. Toxins 2016, 8, 72. [Google Scholar] [CrossRef]
  24. Salinas, N.; Colletier, J.-P.; Moshe, A.; Landau, M. Extreme amyloid polymorphism in Staphylococcus aureus virulent PSMα peptides. Nat. Commun. 2018, 9, 3512. [Google Scholar] [CrossRef]
  25. Bukowski, M.; Wladyka, B.; Dubin, G. Exfoliative toxins of Staphylococcus aureus. Toxins 2010, 2, 1148–1165. [Google Scholar] [CrossRef]
  26. Gaebler Vasconcelos, N.; de Lourdes Ribeiro de Souza da Cunha, M. Staphylococcal enterotoxins: Molecular aspects and detection methods. J. Public Health Epidemiol. 2010, 2, 29–42. [Google Scholar]
  27. Weese, J.S. Methicillin-Resistant Staphylococcus aureus in Animals. ILAR J. 2010, 51, 233–244. [Google Scholar] [CrossRef]
  28. Graveland, H.; Wagenaar, J.A.; Bergs, K.; Heesterbeek, H.; Heederik, D. Persistence of Livestock Associated MRSA CC398 in Humans Is Dependent on Intensity of Animal Contact. PLoS ONE 2011, 6, e16830. [Google Scholar] [CrossRef] [PubMed]
  29. Dorado-García, A.; Bos, M.E.; Graveland, H.; Van Cleef, B.A.; Verstappen, K.M.; Kluytmans, J.A.; Wagenaar, J.A.; Heederik, D.J. Risk factors for persistence of livestock-associated MRSA and environmental exposure in veal calf farmers and their family members: An observational longitudinal study. BMJ Open 2013, 3, e003272. [Google Scholar] [CrossRef] [PubMed]
  30. Van Loo, I.; Huijsdens, X.; Tiemersma, E.; De Neeling, A.; van de Sande-Bruinsma, N.; Beaujean, D.; Voss, A.; Kluytmans, J. Emergence of methicillin-resistant Staphylococcus aureus of animal origin in humans. Emerg. Infect. Dis. 2007, 13, 1834. [Google Scholar] [CrossRef]
  31. van Duijkeren, E.; Moleman, M.; van Oldruitenborgh-Oosterbaan, M.S.; Multem, J.; Troelstra, A.; Fluit, A.; van Wamel, W.; Houwers, D.; de Neeling, A.; Wagenaar, J. Methicillin-resistant Staphylococcus aureus in horses and horse personnel: An investigation of several outbreaks. Veter. Microbiol. 2010, 141, 96–102. [Google Scholar] [CrossRef]
  32. Idelevich, E.A.; Lanckohr, C.; Horn, D.; Wieler, L.H.; Becker, K.; Koeck, R. Multidrug-resistant bacteria in Germany. The impact of sources outside healthcare facilities. Bundesgesundheitsblatt Gesundh. Gesundh. 2016, 59, 113–123. [Google Scholar] [CrossRef] [Green Version]
  33. Köck, R.; Ballhausen, B.; Bischoff, M.; Cuny, C.; Eckmanns, T.; Fetsch, A.; Harmsen, D.; Goerge, T.; Oberheitmann, B.; Schwarz, S.; et al. The impact of zoonotic MRSA colonization and infection in Germany. Berl. Munch. Tierarztl. Wochenschr. 2015, 127, 384–398. [Google Scholar]
  34. Lienen, T.; Schnitt, A.; Hammerl, J.A.; Maurischat, S.; Tenhagen, B.-A. Genomic Distinctions of LA-MRSA ST398 on Dairy Farms From Different German Federal States With a Low Risk of Severe Human Infections. Front. Microbiol. 2021, 11, 575321. [Google Scholar] [CrossRef]
  35. Rybak, M.J.; Le, J.; Lodise, T.P.; Levine, D.P.; Bradley, J.S.; Liu, C.; Mueller, B.A.; Pai, M.P.; Wong-Beringer, A.; Rotschafer, J.C.; et al. Therapeutic Monitoring of Vancomycin for Serious Methicillin-resistant Staphylococcus aureus Infections: A Revised Consensus Guideline and Review by the American Society of Health-system Pharmacists, the Infectious Diseases Society of America, the Pediatric Infectious Diseases Society, and the Society of Infectious Diseases Pharmacists. Clin. Infect. Dis. 2020, 71, 1361–1364. [Google Scholar] [CrossRef] [PubMed]
  36. Washer, P.; Joffe, H. The “hospital superbug”: Social representations of MRSA. Soc. Sci. Med. 2006, 63, 2141–2152. [Google Scholar] [CrossRef] [PubMed]
  37. Chen, H.; Du, Y.; Xia, Q.; Li, Y.; Song, S.; Huang, X. Role of linezolid combination therapy for serious infections: Review of the current evidence. Eur. J. Clin. Microbiol. Infect. Dis. 2020, 39, 1043–1052. [Google Scholar] [CrossRef]
  38. Kelesidis, T.; Humphries, R.; Ward, K.; Lewinski, M.A.; Yang, O.O. Combination therapy with daptomycin, linezolid, and rifampin as treatment option for MRSA meningitis and bacteremia. Diagn. Microbiol. Infect. Dis. 2011, 71, 286–290. [Google Scholar] [CrossRef]
  39. Jacqueline, C.; Navas, D.; Batard, E.; Miegeville, A.-F.; Le Mabecque, V.; Kergueris, M.-F.; Bugnon, D.; Potel, G.; Caillon, J. In Vitro and In Vivo Synergistic Activities of Linezolid Combined with Subinhibitory Concentrations of Imipenem against Methicillin-Resistant Staphylococcus aureus. Antimicrob. Agents Chemother. 2005, 49, 45–51. [Google Scholar] [CrossRef]
  40. Mehta, S.; Singh, C.; Plata, K.B.; Chanda, P.K.; Paul, A.; Riosa, S.; Rosato, R.R.; Rosato, A.E. β-Lactams Increase the Antibacterial Activity of Daptomycin against Clinical Methicillin-Resistant Staphylococcus aureus Strains and Prevent Selection of Daptomycin-Resistant Derivatives. Antimicrob. Agents Chemother. 2012, 56, 6192–6200. [Google Scholar] [CrossRef]
  41. Alosaimy, S.; Sabagha, N.L.; Lagnf, A.M.; Zasowski, E.J.; Morrisette, T.; Jorgensen, S.C.J.; Trinh, T.D.; Mynatt, R.P.; Rybak, M.J. Monotherapy with Vancomycin or Daptomycin versus Combination Therapy with β-Lactams in the Treatment of Methicillin-Resistant Staphylococcus Aureus Bloodstream Infections: A Retrospective Cohort Analysis. Infect. Dis. Ther. 2020, 9, 325–339. [Google Scholar] [CrossRef]
  42. Dilworth, T.J.; Ibrahim, O.; Hall, P.; Sliwinski, J.; Walraven, C.; Mercier, R.-C. β-Lactams Enhance Vancomycin Activity against Methicillin-Resistant Staphylococcus aureus Bacteremia Compared to Vancomycin Alone. Antimicrob. Agents Chemother. 2014, 58, 102–109. [Google Scholar] [CrossRef]
  43. Morrisette, T.; Alosaimy, S.; Abdul-Mutakabbir, J.; Kebriaei, R.; Rybak, M. The Evolving Reduction of Vancomycin and Daptomycin Susceptibility in MRSA—Salvaging the Gold Standards with Combination Therapy. Antibiotics 2020, 9, 762. [Google Scholar] [CrossRef] [PubMed]
  44. Bellos, I.; Karageorgiou, V.; Pergialiotis, V.; Perrea, D. Acute kidney injury following the concurrent administration of antipseudomonal β-lactams and vancomycin: A network meta-analysis. Clin. Microbiol. Infect. 2020, 26, 696–705. [Google Scholar] [CrossRef]
  45. Tong, S.Y.; Lye, D.C.; Yahav, D.; Sud, A.; Robinson, J.O.; Nelson, J.; Archuleta, S.; Roberts, M.A.; Cass, A.; Davis, J. SEffect of vancomycin or daptomycin with vs without an antistaphylococcal β-lactam on mortality, bacteremia, relapse, or treatment failure in patients with MRSA bacteremia: A randomized clinical trial. JAMA 2020, 323, 527–537. [Google Scholar] [CrossRef]
  46. Wang, C.; Ye, C.; Liao, L.; Wang, Z.; Hu, Y.; Deng, C.; Liu, L. Adjuvant β-Lactam Therapy Combined with Vancomycin or Daptomycin for Methicillin-Resistant Staphylococcus aureus Bacteremia: A Systematic Review and Meta-analysis. Antimicrob. Agents Chemother. 2020, 64, e01377-20. [Google Scholar] [CrossRef] [PubMed]
  47. Mohamed, M.; Abdelkhalek, A.; Seleem, M. Evaluation of short synthetic antimicrobial peptides for treatment of drug-resistant and intracellular Staphylococcus aureus. Sci. Rep. 2016, 6, 29707. [Google Scholar] [CrossRef] [PubMed]
  48. Yao, C.-J.; Li, Y.-L.; Pu, M.-J.; Luo, L.-H.; Xiong, Q.; Xie, F.-J.; Li, T.-L.; Feng-Jiao, X. Aminoglycosides with Anti-MRSA Activity: A Concise Review. Curr. Top. Med. Chem. 2021, 21, 2483–2499. [Google Scholar] [CrossRef]
  49. Verma, R.; Verma, S.K.; Rakesh, K.P.; Girish, Y.R.; Ashrafizadeh, M.; Kumar, K.S.S.; Rangappa, K.S. Pyrazole-based analogs as potential antibacterial agents against methicillin-resistance staphylococcus aureus (MRSA) and its SAR elucidation. Eur. J. Med. Chem. 2021, 212, 113134. [Google Scholar] [CrossRef] [PubMed]
  50. Leng, B.; Yan, G.; Li, T.; Hou, N. Vancomycin-induced reversible pancytopenia and rash in a 16-month-old boy with osteomyelitis: A case report. Int. J. Clin. Pharmacol. Ther. 2020, 58, 242–246. [Google Scholar] [CrossRef]
  51. Niu, H.; Yang, T.; Wang, J.; Wang, R.; Cai, Y. Immunomodulatory Effect of Colistin and its Protective Role in Rats with Methicillin-Resistant Staphylococcus aureus-induced Pneumonia. Front. Pharmacol. 2021, 11, 602054. [Google Scholar] [CrossRef] [PubMed]
  52. Parish, D.; Scheinfeld, N. Ceftaroline fosamil, a cephalosporin derivative for the potential treatment of MRSA infection. Curr. Opin. Investig. Drugs 2008, 9, 201–209. [Google Scholar] [PubMed]
  53. Shirley, D.-A.T.; Heil, E.L.; Johnson, J.K. Ceftaroline Fosamil: A Brief Clinical Review. Infect. Dis. Ther. 2013, 2, 95–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Torres, A.; Soriano, A.; Rivolo, S.; Remak, E.; Peral, C.; Kantecki, M.; Ansari, W.; Charbonneau, C.; Hammond, J.; Grau, S.; et al. Ceftaroline Fosamil for the Empiric Treatment of Hospitalized Adults with cSSTI: An Economic Analysis from the Perspective of the Spanish National Health System. Clin. Econ. Outcomes Res. 2022, 14, 149–161. [Google Scholar] [CrossRef] [PubMed]
  55. Stefani, S.; Bongiorno, D.; Mongelli, G.; Campanile, F. Linezolid Resistance in Staphylococci. Pharmaceuticals 2010, 3, 1988–2006. [Google Scholar] [CrossRef]
  56. Shariati, A.; Dadashi, M.; Chegini, Z.; van Belkum, A.; Mirzaii, M.; Khoramrooz, S.S.; Darban-Sarokhalil, D. The global prevalence of Daptomycin, Tigecycline, Quinupristin/Dalfopristin, and Linezolid-resistant Staphylococcus aureus and coagulase–negative staphylococci strains: A systematic review and meta-analysis. Antimicrob. Resist. Infect. Control. 2020, 9, 1–20. [Google Scholar] [CrossRef]
  57. Li, C.; Li, Z.; Gan, Y.; Jiang, F.; Zhao, H.; Tan, J.; Yang, Y.Y.; Yuan, P.; Ding, X. Selective Capture, Separation, and Photothermal Inactivation of Methicillin-Resistant Staphylococcus aureus (MRSA) Using Functional Magnetic Nanoparticles. ACS Appl. Mater. Interfaces 2022, 14, 20566–20575. [Google Scholar] [CrossRef] [PubMed]
  58. Mohamed, M.B.E.D.; El-Ela, F.I.A.; Mahmoud, R.K.; Farghali, A.A.; Gamil, S.; Aziz, S.A.A.A. Cefotax-magnetic nanoparticles as an alternative approach to control Methicillin-Resistant Staphylococcus aureus (MRSA) from different sources. Sci. Rep. 2022, 12, 624. [Google Scholar] [CrossRef]
  59. Zhang, L.; Bhatti, M.M.; Michaelides, E.E.; Marin, M.; Ellahi, R. Hybrid nanofluid flow towards an elastic surface with tantalum and nickel nanoparticles, under the influence of an induced magnetic field. Eur. Phys. J. Spéc. Top. 2022, 231, 521–533. [Google Scholar] [CrossRef]
  60. Hou, Y.; Kondoh, H.; Ohta, T.; Gao, S. Size-controlled synthesis of nickel nanoparticles. Appl. Surf. Sci. 2005, 241, 218–222. [Google Scholar] [CrossRef]
  61. Abdollahi, A.; Mirzaei, E.; Amoozegar, F.; Moemenbellah-Fard, M.D.; Zarenezhad, E.; Osanloo, M. High Antibacterial Effect of Impregnated Nanofiber Mats with a Green Nanogel Against Major Human Pathogens. BioNanoScience 2021, 11, 549–558. [Google Scholar] [CrossRef]
  62. Qasemi, H.; Fereidouni, Z.; Karimi, J.; Abdollahi, A.; Zarenezhad, E.; Rasti, F.; Osanloo, M. Promising antibacterial effect of impregnated nanofiber mats with a green nanogel against clinical and standard strains of Pseudomonas aeruginosa and Staphylococcus aureus. J. Drug Deliv. Sci. Technol. 2021, 66, 102844. [Google Scholar] [CrossRef]
  63. Hulme, J. Application of Nanomaterials in the Prevention, Detection, and Treatment of Methicillin-Resistant Staphylococcus aureus (MRSA). Pharmaceutics 2022, 14, 805. [Google Scholar] [CrossRef] [PubMed]
  64. Nisar, P.; Ali, N.; Rahman, L.; Ali, M.; Shinwari, Z.K. Antimicrobial activities of biologically synthesized metal nanoparticles: An insight into the mechanism of action. JBIC J. Biol. Inorg. Chem. 2019, 24, 929–941. [Google Scholar] [CrossRef]
  65. Mendes, C.R.; Dilarri, G.; Forsan, C.F.; Sapata, V.d.M.R.; Lopes, P.R.M.; de Moraes, P.B.; Montagnolli, R.N.; Ferreira, H.; Bidoia, E.D. Antibacterial action and target mechanisms of zinc oxide nanoparticles against bacterial pathogens. Sci. Rep. 2022, 12, 2658. [Google Scholar] [CrossRef]
  66. Radzig, M.; Nadtochenko, V.; Koksharova, O.; Kiwi, J.; Lipasova, V.; Khmel, I. Antibacterial effects of silver nanoparticles on gram-negative bacteria: Influence on the growth and biofilms formation, mechanisms of action. Colloids Surf. B Biointerfaces 2013, 102, 300–306. [Google Scholar] [CrossRef]
  67. Nunez, N.V.A.; Villegas, H.H.L.; Turrent, L.D.C.I.; Padilla, C.R. Silver Nanoparticles Toxicity and Bactericidal Effect Against Methicillin-Resistant Staphylococcus aureus: Nanoscale Does Matter. Nanobiotechnology 2009, 5, 2–9. [Google Scholar] [CrossRef]
  68. Hemeg, H.A. Nanomaterials for alternative antibacterial therapy. Int. J. Nanomed. 2017, 12, 8211–8225. [Google Scholar] [CrossRef]
  69. Gwon, K.; Kim, Y.; Cho, H.; Lee, S.; Yang, S.-H.; Kim, S.-J.; Lee, D. Robust Copper Metal–Organic Framework-Embedded Polysiloxanes for Biomedical Applications: Its Antibacterial Effects on MRSA and In Vitro Cytotoxicity. Nanomaterials 2021, 11, 719. [Google Scholar] [CrossRef] [PubMed]
  70. Kadiyala, U.; Turali-Emre, E.S.; Bahng, J.H.; Kotov, N.A.; VanEpps, J.S. Unexpected insights into antibacterial activity of zinc oxide nanoparticles against methicillin resistant Staphylococcus aureus (MRSA). Nanoscale 2018, 10, 4927–4939. [Google Scholar] [CrossRef]
  71. Ahmad, A.; Sabir, A.; Iqbal, S.S.; Felemban, B.F.; Riaz, T.; Bahadar, A.; Hossain, N.; Khan, R.U.; Inam, F. Novel antibacterial polyurethane and cellulose acetate mixed matrix membrane modified with functionalized TiO2 nanoparticles for water treatment applications. Chemosphere 2022, 301, 134711. [Google Scholar] [CrossRef] [PubMed]
  72. Javed, R.; Ain, N.U.; Gul, A.; Ahmad, M.A.; Guo, W.; Ao, Q.; Tian, S. Diverse biotechnological applications of multifunctional titanium dioxide nanoparticles: An up-to-date review. IET Nanobiotechnol. 2022, 16, 171–189. [Google Scholar] [CrossRef]
  73. Abebe, B.; Zereffa, E.A.; Tadesse, A.; Murthy, H.C.A. A Review on Enhancing the Antibacterial Activity of ZnO: Mechanisms and Microscopic Investigation. Nanoscale Res. Lett. 2020, 15, 190. [Google Scholar] [CrossRef]
  74. Al-Shawi, S.G.; Andreevna Alekhina, N.; Aravindhan, S.; Thangavelu, L.; Elena, A.; Viktorovna Kartamysheva, N.; Rafkatovna Zakieva, R. Synthesis of NiO nanoparticles and sulfur, and nitrogen co doped-graphene quantum dots/nio nanocomposites for antibacterial application. J. Nanostruct. 2021, 11, 181–188. [Google Scholar]
  75. Kannan, K.; Radhika, D.; Nesaraj, A.; Sadasivuni, K.K.; Reddy, K.R.; Kasai, D.; Raghu, A.V. Photocatalytic, antibacterial and electrochemical properties of novel rare earth metal oxides-based nanohybrids. Mater. Sci. Energy Technol. 2020, 3, 853–861. [Google Scholar] [CrossRef]
  76. Munawar, T.; Iqbal, F.; Yasmeen, S.; Mahmood, K.; Hussain, A. Multi metal oxide NiO-CdO-ZnO nanocomposite–synthesis, structural, optical, electrical properties and enhanced sunlight driven photocatalytic activity. Ceram. Int. 2020, 46, 2421–2437. [Google Scholar] [CrossRef]
  77. Esmaeili, F.; Hosseini-Nasr, M.; Rad-Malekshahi, M.; Samadi, N.; Atyabi, F.; Dinarvand, R. Preparation and antibacterial activity evaluation of rifampicin-loaded poly lactide-co-glycolide nanoparticles. Nanomed. Nanotechnol. Biol. Med. 2007, 3, 161–167. [Google Scholar] [CrossRef]
  78. Jiang, J.-L.; Li, Y.-F.; Fang, T.-L.; Zhou, J.; Li, X.-L.; Wang, Y.-C.; Dong, J. Vancomycin-loaded nano-hydroxyapatite pellets to treat MRSA-induced chronic osteomyelitis with bone defect in rabbits. Inflamm. Res. 2011, 61, 207–215. [Google Scholar] [CrossRef] [PubMed]
  79. Saadh, M. Effect of silver nanoparticles on the antibacterial activity of Levofloxacin against methicillin-resistant Staphylococcus aureus. Eur. Rev. Med. Pharmacol. Sci. 2021, 25, 5507–5510. [Google Scholar] [PubMed]
  80. Wang, G.; Wang, J.-J.; Li, F.; To, S.-S.T. Development and Evaluation of a Novel Drug Delivery: Pluronics/SDS Mixed Micelle Loaded With Myricetin In Vitro and In Vivo. J. Pharm. Sci. 2016, 105, 1535–1543. [Google Scholar] [CrossRef]
  81. Adler-Moore, J.P.; Gangneux, J.-P.; Pappas, P.G. Comparison between liposomal formulations of amphotericin B. Sabouraudia 2016, 54, 223–231. [Google Scholar] [CrossRef]
  82. Ghasemiyeh, P.; Mohammadi-Samani, S. Potential of Nanoparticles as Permeation Enhancers and Targeted Delivery Options for Skin: Advantages and Disadvantages. Drug Des. Dev. Ther. 2020, 14, 3271–3289. [Google Scholar] [CrossRef]
  83. Jones, S.; Pramanik, A.; Kanchanapally, R.; Nellore, B.P.V.; Begum, S.; Sweet, C.; Ray, P.C. Multifunctional Three-Dimensional Chitosan/Gold Nanoparticle/Graphene Oxide Architecture for Separation, Label-Free SERS Identification of Pharmaceutical Contaminants, and Effective Killing of Superbugs. ACS Sustain. Chem. Eng. 2017, 5, 7175–7187. [Google Scholar] [CrossRef]
  84. Devlin, H.; Fulaz, S.; Hiebner, D.W.; O’Gara, J.P.; Casey, E. Enzyme-Functionalized Mesoporous Silica Nanoparticles to Target Staphylococcus aureus and Disperse Biofilms. Int. J. Nanomed. 2021, 16, 1929–1942. [Google Scholar] [CrossRef]
  85. Mubeen, B.; Ansar, A.N.; Rasool, R.; Ullah, I.; Imam, S.S.; Alshehri, S.; Ghoneim, M.M.; Alzarea, S.I.; Nadeem, M.S.; Kazmi, I. Nanotechnology as a Novel Approach in Combating Microbes Providing an Alternative to Antibiotics. Antibiotics 2021, 10, 1473. [Google Scholar] [CrossRef]
  86. Marzi, M.; Chijan, M.R.; Zarenezhad, E. Hydrogels as Promising Therapeutic Strategy for the Treatment of Skin Cancer. J. Mol. Struct. 2022, 1262, 133014. [Google Scholar] [CrossRef]
  87. Raza, S.; Matuła, K.; Karoń, S.; Paczesny, J. Resistance and Adaptation of Bacteria to Non-Antibiotic Antibacterial Agents: Physical Stressors, Nanoparticles, and Bacteriophages. Antibiotics 2021, 10, 435. [Google Scholar] [CrossRef]
  88. Mamun, M.M.; Sorinolu, A.J.; Munir, M.; Vejerano, E.P. Nanoantibiotics: Functions and Properties at the Nanoscale to Combat Antibiotic Resistance. Front. Chem. 2021, 9, 687660. [Google Scholar] [CrossRef]
  89. Malik, A.; Alshehri, M.A.; Alamery, S.F.; Khan, J.M. Impact of metal nanoparticles on the structure and function of metabolic enzymes. Int. J. Biol. Macromol. 2021, 188, 576–585. [Google Scholar] [CrossRef]
  90. Niculescu, A.-G.; Chircov, C.; Grumezescu, A.M. Magnetite nanoparticles: Synthesis methods—A comparative review. Methods 2021, 199, 16–27. [Google Scholar] [CrossRef] [PubMed]
  91. Cuenya, B.R. Synthesis and catalytic properties of metal nanoparticles: Size, shape, support, composition, and oxidation state effects. Thin Solid Films 2010, 518, 3127–3150. [Google Scholar] [CrossRef]
  92. Ali, A.; Shah, T.; Ullah, R.; Zhou, P.; Guo, M.; Ovais, M.; Tan, Z.; Rui, Y. Review on Recent Progress in Magnetic Nanoparticles: Synthesis, Characterization, and Diverse Applications. Front. Chem. 2021, 9, 629054. [Google Scholar] [CrossRef]
  93. Ying, S.; Guan, Z.; Ofoegbu, P.C.; Clubb, P.; Rico, C.; He, F.; Hong, J. Green synthesis of nanoparticles: Current developments and limitations. Environ. Technol. Innov. 2022, 26, 102336. [Google Scholar] [CrossRef]
  94. Jeevanandam, J.; Kiew, S.F.; Boakye-Ansah, S.; Lau, S.Y.; Barhoum, A.; Danquah, M.K.; Rodrigues, J. Green approaches for the synthesis of metal and metal oxide nanoparticles using microbial and plant extracts. Nanoscale 2022, 14, 2534–2571. [Google Scholar] [CrossRef]
  95. Wang, Z.K.; Kuok, M.H.; Ng, S.C.; Lockwood, D.J.; Cottam, M.G.; Nielsch, K.; Wehrspohn, R.B.; Gösele, U. Spin-Wave Quantization in Ferromagnetic Nickel Nanowires. Phys. Rev. Lett. 2002, 89, 027201. [Google Scholar] [CrossRef]
  96. Zheng, W.; Sun, C.Q. Electronic process of nitriding: Mechanism and applications. Prog. Solid State Chem. 2006, 34, 1–20. [Google Scholar] [CrossRef]
  97. Lee, K.-B.; Park, S.; Mirkin, C.A. Multicomponent Magnetic Nanorods for Biomolecular Separations. Angew. Chem. Int. Ed. 2004, 43, 3048–3050. [Google Scholar] [CrossRef]
  98. Hatamifard, A.; Nasrollahzadeh, M.; Sajadi, S.M. Biosynthesis, characterization and catalytic activity of an Ag/zeolite nanocomposite for base- and ligand-free oxidative hydroxylation of phenylboronic acid and reduction of a variety of dyes at room temperature. New J. Chem. 2016, 40, 2501–2513. [Google Scholar] [CrossRef]
  99. Jaji, N.-D.; Lee, H.L.; Hussin, M.H.; Akil, H.; Zakaria, M.R.; Othman, M.B.H. Advanced nickel nanoparticles technology: From synthesis to applications. Nanotechnol. Rev. 2020, 9, 1456–1480. [Google Scholar] [CrossRef]
  100. Iqbal, J.; Abbasi, B.A.; Mahmood, T.; Hameed, S.; Munir, A.; Kanwal, S. Green synthesis and characterizations of Nickel oxide nanoparticles using leaf extract of Rhamnus virgata and their potential biological applications. Appl. Organomet. Chem. 2019, 33, e4950. [Google Scholar] [CrossRef]
  101. De, M.; Ghosh, P.S.; Rotello, V.M. Applications of Nanoparticles in Biology. Adv. Mater. 2008, 20, 4225–4241. [Google Scholar] [CrossRef]
  102. Materón, E.M.; Miyazaki, C.M.; Carr, O.; Joshi, N.; Picciani, P.H.; Dalmaschio, C.J.; Davis, F.; Shimizu, F.M. Magnetic nanoparticles in biomedical applications: A review. Appl. Surf. Sci. Adv. 2021, 6, 100163. [Google Scholar] [CrossRef]
  103. Galdiero, S.; Falanga, A.; Berisio, R.; Grieco, P.; Morelli, G.; Galdiero, M. Antimicrobial Peptides as an Opportunity Against Bacterial Diseases. Curr. Med. Chem. 2015, 22, 1665–1677. [Google Scholar] [CrossRef]
  104. Xing, K.; Zhu, X.; Peng, X.; Qin, S. Chitosan antimicrobial and eliciting properties for pest control in agriculture: A review. Agron. Sustain. Dev. 2015, 35, 569–588. [Google Scholar] [CrossRef] [Green Version]
  105. Chaudhary, R.G.; Tanna, J.A.; Gandhare, N.V.; Rai, A.R.; Juneja, H.D. Synthesis Of Nickel Nanoparticles: Microscopic Investigation, An Efficient Catalyst And Effective Antibacterial Activity. Adv. Mater. Lett. 2015, 6, 990–998. [Google Scholar] [CrossRef]
  106. Haghshenas, L.; Faraji, A. Evaluation of the effect of Gold and Nickel nanoparticles on Escherichia coliand Staphylococcus aurousbacteria in milk. J. Micro Nano Biomed. 2016, 1, 1–6. [Google Scholar]
  107. Haider, A.; Ijaz, M.; Ali, S.; Haider, J.; Imran, M.; Majeed, H.; Shahzadi, I.; Ali, M.M.; Khan, J.A.; Ikram, M. Green Synthesized Phytochemically (Zingiber officinale and Allium sativum) Reduced Nickel Oxide Nanoparticles Confirmed Bactericidal and Catalytic Potential. Nanoscale Res. Lett. 2020, 15, 50. [Google Scholar] [CrossRef]
  108. Rheima, A.M.; Al Marjani, M.F.; Aboksour, M.F.; Mohammed, S.H. Evaluation of Anti-Biofilm Formation Effect of Nickel Oxide Nanoparticles (NiO-NPs) Against Methicillin-Resistant Staphylococcus Aureus (MRSA). Int. J. Nanosci. Nanotechnol. 2021, 17, 221–230. [Google Scholar]
  109. Deotale, A.J.; Nandedkar, R.V.; Sinha, A.K.; Upadhyay, A.; Mondal, P.; Srivastava, A.K.; Deb, S.K. Effect of isochronal annealing on phase transformation studies of iron oxide nanoparticles. Bull. Mater. Sci. 2015, 38, 599–606. [Google Scholar] [CrossRef]
  110. Deotale, A.J.; Singh, U.; Golani, M.; Hajela, K.; Nandedkar, R.V. Raman spectroscopic study of nickel oxide nanoparticles and its antibacterial activity. In AIP Conference Proceedings; AIP Publishing LLC: Melville, NY, USA, 2021. [Google Scholar] [CrossRef]
  111. Saleem, S.; Ahmed, B.; Khan, M.S.; Al-Shaeri, M.; Musarrat, J. Inhibition of growth and biofilm formation of clinical bacterial isolates by NiO nanoparticles synthesized from Eucalyptus globulus plants. Microb. Pathog. 2017, 111, 375–387. [Google Scholar] [CrossRef]
  112. Ali, S.R.; De, M. Superparamagnetic Nickel Nanocluster-Embedded MoS2 Nanosheets for Gram-Selective Bacterial Adhesion and Antibacterial Activity. ACS Biomater. Sci. Eng. 2022, 8, 2932–2942. [Google Scholar] [CrossRef]
  113. Rashad, M.M.; El-Kemary, N.M.; Amer, S.; El-Kemary, M. Bovine serum albumin/chitosan-nanoparticle bio-complex; spectroscopic study and in vivo toxicological—Hypersensitivity evaluation. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2021, 253, 119582. [Google Scholar] [CrossRef]
  114. Scutera, S.; Argenziano, M.; Sparti, R.; Bessone, F.; Bianco, G.; Bastiancich, C.; Castagnoli, C.; Stella, M.; Musso, T.; Cavalli, R. Enhanced Antimicrobial and Antibiofilm Effect of New Colistin-Loaded Human Albumin Nanoparticles. Antibiotics 2021, 10, 57. [Google Scholar] [CrossRef] [PubMed]
  115. Yeo, W.W.Y.; Maran, S.; Kong, A.S.-Y.; Cheng, W.-H.; Lim, S.-H.E.; Loh, J.-Y.; Lai, K.-S. A Metal-Containing NP Approach to Treat Methicillin-Resistant Staphylococcus aureus (MRSA): Prospects and Challenges. Materials 2022, 15, 5802. [Google Scholar] [CrossRef]
Figure 1. Flowchart of study procedure in this review.
Figure 1. Flowchart of study procedure in this review.
Antibiotics 11 01208 g001
Figure 2. Applications of NiO NPs [92,101,102].
Figure 2. Applications of NiO NPs [92,101,102].
Antibiotics 11 01208 g002
Figure 3. Nanocarriers as skin penetration enhancers in dermal drug delivery. NE: nanoemulsion, SLN: solid lipid nanoparticles, NLC: nanostructured lipid carriers.
Figure 3. Nanocarriers as skin penetration enhancers in dermal drug delivery. NE: nanoemulsion, SLN: solid lipid nanoparticles, NLC: nanostructured lipid carriers.
Antibiotics 11 01208 g003
Table 1. NiNP concentrations and conditions against MRSA in vitro.
Table 1. NiNP concentrations and conditions against MRSA in vitro.
NanoparticleMICMechanism of ActionConditionsReference
NiNPs0.21 (µg/mL)NDIn vitro[106]
NiO NPs1 mg/50 µLNDIn vitro[107]
NiO NPs265 µg/mLNDIn vitro[108]
(S,N-GQDs/NiO) NPsND *NDIn vitro[74]
NiO NPsND **NDIn vitro[110]
NiO NPs0.8 mg/mL ***NDIn vitro[111]
MIC: minimum inhibitory concentration, ND: not determined, ND *: disk diffusion test (17 mm), ND **: disk diffusion test (17 mm), *** MBC: (minimum bactericidal concentration) included 1.6 mg/mL.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zarenezhad, E.; Abdulabbas, H.T.; Marzi, M.; Ghazy, E.; Ekrahi, M.; Pezeshki, B.; Ghasemian, A.; Moawad, A.A. Nickel Nanoparticles: Applications and Antimicrobial Role against Methicillin-Resistant Staphylococcus aureus Infections. Antibiotics 2022, 11, 1208. https://doi.org/10.3390/antibiotics11091208

AMA Style

Zarenezhad E, Abdulabbas HT, Marzi M, Ghazy E, Ekrahi M, Pezeshki B, Ghasemian A, Moawad AA. Nickel Nanoparticles: Applications and Antimicrobial Role against Methicillin-Resistant Staphylococcus aureus Infections. Antibiotics. 2022; 11(9):1208. https://doi.org/10.3390/antibiotics11091208

Chicago/Turabian Style

Zarenezhad, Elham, Hussein T. Abdulabbas, Mahrokh Marzi, Esraa Ghazy, Mohammad Ekrahi, Babak Pezeshki, Abdolmajid Ghasemian, and Amira A. Moawad. 2022. "Nickel Nanoparticles: Applications and Antimicrobial Role against Methicillin-Resistant Staphylococcus aureus Infections" Antibiotics 11, no. 9: 1208. https://doi.org/10.3390/antibiotics11091208

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