Next Article in Journal
Description of AtCAX4 in Response to Abiotic Stress in Arabidopsis
Previous Article in Journal
Survival Pathways Are Differently Affected by Microgravity in Normal and Cancerous Breast Cells
Previous Article in Special Issue
Emerging Options for the Diagnosis of Bacterial Infections and the Characterization of Antimicrobial Resistance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pulling the Brakes on Fast and Furious Multiple Drug-Resistant (MDR) Bacteria

1
Center for Precision Medicine, Hochschule Furtwangen University, Jakob-Kienzle-Str. 17, 78054 Villingen-Schwenningen, Germany
2
Department of Biosciences, COMSATS University Islamabad, Park Road, Tarlai Kalan, Islamabad 45550, Pakistan
3
Mechano(bio)chemistry, Max Planck Institute for Colloids and Interfaces, Am Mühlenberg 1, Golm, 14476 Potsdam, Germany
4
EXIM Department, Fraunhofer Institute IZI, Leipzig, Schillingallee 68, 18057 Rostock, Germany
5
Faculty of Science, University of Tuebingen, Auf der Morgenstelle 8, 72076 Tuebingen, Germany
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2021, 22(2), 859; https://doi.org/10.3390/ijms22020859
Submission received: 22 December 2020 / Revised: 11 January 2021 / Accepted: 13 January 2021 / Published: 16 January 2021

Abstract

:
Life-threatening bacterial infections have been managed by antibiotics for years and have significantly improved the wellbeing and lifetime of humans. However, bacteria have always been one step ahead by inactivating the antimicrobial agent chemically or by producing certain enzymes. The alarming universal occurrence of multidrug-resistant (MDR) bacteria has compelled researchers to find alternative treatments for MDR infections. This is a menace where conventional chemotherapies are no longer promising, but several novel approaches could help. Our current review article discusses the novel approaches that can combat MDR bacteria: starting off with potential nanoparticles (NPs) that efficiently interact with microorganisms causing fatal changes in the morphology and structure of these cells; nanophotothermal therapy using inorganic NPs like AuNPs to destroy pathogenic bacterial cells; bacteriophage therapy against which bacteria develop less resistance; combination drugs that act on dissimilar targets in distinctive pathways; probiotics therapy by the secretion of antibacterial chemicals; blockage of quorum sensing signals stopping bacterial colonization, and vaccination against resistant bacterial strains along with virulence factors. All these techniques show us a promising future in the fight against MDR bacteria, which remains the greatest challenge in public health care.

1. Introduction

Infectious diseases around the globe, once cured with the help of the magical drugs “antibiotics”, are now becoming a menace due to ever-increasing microbial antibiotic resistance. This emerging resistance is due to the irrational use of antibiotics in humans, veterinary, and agriculture because of their easy and unregulated access, especially in developing countries. This antibiotic resistance is either acquired naturally or artificially (by transfer of resistance genes) [1,2,3]. The occurrence of antibiotic resistance first became evident when Staphylococci encountered the first commercially produced antibiotic, penicillin, which produced an enzyme (penicillinase) to degrade it. The continuous use of diverse drugs has imposed a selective pressure on bacteria, transforming them into “superbugs”, also known as multiple drug resistant (MDR) microorganisms [3]. MDR bacteria that are very difficult to treat include Pseudomonas aeruginosa, Acinetobacter baumannii, Escherichia coli, Klebsiella pneumoniae, vancomycin-resistant Enterococci (VRE), methicillin-resistant Staphylococcus aureus (MRSA), vancomycin-resistant S. aureus (VRSA), and extensively drug-resistant (XDR) Mycobacterium tuberculosis [4]. Bacteria can exhibit resistance either naturally, i.e., entirely lacking the target or having low-affinity targets; by having the potential to inactivate the antibiotics; low cell permeability and the presence of efficient efflux pumps, by the transfer of resistant genes found on plasmids, transposons, and bacteriophages [1,2]. Different sophisticated mechanisms have been evolved by bacteria for resistance to antibiotics and to protect themselves from being killed by these antimicrobial agents [5]. Among these different mechanisms, one of the most successful processes is to inactivate the antimicrobial agent by chemically altering or destroying it. It is accomplished by producing certain enzymes, e.g., aminoglycoside modifying enzymes that modify the amino groups of aminoglycoside molecules, β-lactamases that destroy the amide bond of the β-lactam ring of β-lactam antibiotics [6]. Another mechanism of resistance is by decreasing the influx of antimicrobial agents in Gram-negative bacteria due to the presence of outer membrane, i.e., vancomycin resistance [7,8,9]. Efflux pumps present in bacteria are capable of secreting the antibiotic out of the cell are also one of the reasons for resistance to antimicrobial agents [10]. Another common strategy of resistance is to prevent antibiotic action by interfering with the target site in several ways like protecting (tetracycline and fluoroquinolones resistance) or modifying the target site (rifamycin resistance). Much astonishing diversity is present within these categories of mechanisms, and several types of resistance may be possessed by a single strain [5,11]. The general modes of resistance in bacteria are diagrammatically shown in Figure 1. There are several mechanisms to prevent and combat these MDR bacteria. One of these includes nanoparticles (NPs) acting as a weapon against emerging antibiotic resistance. NPs (and nanoencapsulation platforms), due to their high surface area to volume ratio and functionalizable structural surface(s), can effectively interact with microorganisms causing fatal changes in the morphology and structure of these cells [12,13,14]. Nanophotothermal therapy is an approach in which inorganic NPs such as AuNPs can absorb magnetic radiation and convert it into heat, which can destroy pathogenic bacterial cells in the close locality. MDR bacteria are effectively killed by this technique [1,15,16,17,18]. Bacteriophage therapy is another mechanism to prevent the appearance of these lethal microorganisms. It is no less than a magical cure for many antimicrobial-resistant infections [3]. Bacteriophages are self-replicative, and bacteria develop less resistance against them when compared with the usage of antibiotics. To treat MDR bacteria, a suitable cocktail of phages is required [19].
Combination drugs that act on different targets in different pathways can be used in such a way that even if a bacterium is resistant to one of the drugs, the other drugs targeting different components of bacteria will disrupt them, thus minimizing their propagation [3]. Probiotics can also prevent antibiotic resistance by the secretion of antibacterial chemicals, diminishing the ability of bacteria to colonize the body, thereby reducing the use of antibiotics and hence the emergence of MDR [20,21]. Bacteria communicate with each other through quorum sensing; blockage of these signals would prevent bacterial colonization and therefore considerably reduce the need for antibiotics [22]. Vaccines lower the disease incidence and, concurrently, the need to use antimicrobial drugs. There is a potential prospect to develop vaccines against resistant bacterial strains as well as against virulence factors to step up the game against MDR bacteria [23]. Vaccines may be regarded as being superior to drugs due to being prophylactic and expressing multiple epitopes [24]. Figure 2 represents the different strategies and targets we can employ to fight multidrug resistance in bacteria.

2. Nanoparticles as a Weapon against Antibiotically Resistant Bacteria Extracorporally

Bacteria are responsible for causing a great variety of ailments each year around the globe, and the rapidly growing antibiotic resistance has caused an alarming situation in the field of medical microbiology. Infections due to MDR bacteria pose a great threat because they cause chronic disease states resulting in high rates of mortality, morbidity, and prolonged treatment costs [25]. MDR bacteria that are most problematic to deal with are XDR M. tuberculosis, A. baumannii, P. aeruginosa, MRSA, E. coli, and K. pneumoniae bearing NDM-1 (New Delhi metallo beta-lactamase-1), VRE, and VRSA [1,26,27]. NPs may serve as an imperative tool for fighting antibiotic resistance [28]. The general mode of antimicrobial action of different NPs is shown below, which is based on their exceptionally large surface area and functionalizable structure that enables them to effectively interact with microorganisms causing changes in the morphology and structure of bacterial cells (Figure 2) [1]. The basic mechanism includes anchoring the cell wall of a bacterium and penetrating it, altering the cell membrane permeability, or making it porous by the production of free radicals, ultimately causing cell death [1,16,27,29].
Metallic NPs cause bacterial cell death by interacting with the sulfur and phosphorus in the bacterial DNA bases, thereby destroying the DNA. NPs can also inhibit signal transduction and, consequently, bacterial growth by the dephosphorylation of peptide substrates on tyrosine residues [1,30,31]. Certain NPs can destroy the membrane potential and suppressing the ATPase activities to reduce the levels of ATP in the cell, while others by inhibiting ribosomal subunit from binding the tRNA molecules [1,32,33]. Li et al. reported that 2 nm AuNPs with cationic surface chemistry could interact with the cell membrane of Gram-positive and Gram-negative bacteria resulting in the formation of distinct aggregation patterns and promote bacterial cell lysis [34,35]. Correspondingly, Jiang et al. also established that bacterial cell membrane could be damaged by cationic AuNPs induced protuberance [35,36].

2.1. Metal Oxide, Nitric Oxide, and Chitosan NPs

Metal oxide NPs such as TiO2, CuO, and ZnO, by the production of reactive oxygen species (ROS), also act as antibacterial agents against MRSA and E. coli, but their efficiency is increased when they are coupled with AgNPs. They have photocatalytic activity due to a wide bandgap, which is attributed to the production of ROS [1,37,38]. CuONPs have shown to be effective against a variety of bacterial pathogens, including MRSA, E. coli, S. aureus, P. aeruginosa, N. meningitis, B. cereus, S. pyogenes, and A. baumannii [1,26,39,40]. The mode of action of AgNps is shown in Figure 3. When CuONPs are conjugated with AgNPs, their antibacterial activity is enhanced so that they can completely inhibit bacterial growth. ZnONPs have shown acute toxicity to antibiotic (methicillin)-resistant bacteria such as S. aureus and S. agalactiae. ZnONPs are internalized into the cells where they disorganize and damage the cell, cell membrane and also increase the oxidative stress that damages bacterial proteins, lipids, and DNA [1,41,42]. However, at low concentration, ZnONPs show slight toxicity indicating that the level of toxicity caused by these NPs depends upon their concentration. The colloidal suspension of ZnO is found to inhibit 90% of MRSA, E. faecalis, a high biofilm-producing strain S. epidermidis, and the growth of several other clinically relevant pathogens. ZnONPs were also shown to inhibit bacterial growth of methicillin-sensitive Staphylococcus aureus (MSSA), MRSA, and methicillin-resistant Staphylococcus epidermidis (MRSE) strains. Moreover, these NPs were also found to be effective against extended-spectrum β-lactamases-producing E. coli and K. pneumoniae apart from other bacteria like Vibrio cholera and Campylobacter jejuni [1,43,44,45,46].
Nitric-oxide-releasing NPs (NONPs) also act as antimicrobial agents against many antibiotic-resistant and sensitive bacteria, i.e., K. pneumoniae, E. faecalis, S. pyogenes, E. coli, and P. aeruginosa. NO is unstable in the presence of oxygen and reacts with oxygen or superoxide spontaneously to produce reactive nitrogen and oxygen intermediates that are toxic against cells and act as antimicrobial species. When the concentration of NO is greater than 1 µM, these intermediate species become significantly important because, at these concentrations, reactive nitrogen species (RNOS) like S-nitrosothiols (RSNO), peroxynitrite (OONO−), nitrogen dioxide (NO2) are produced [1,47,48]. Peroxynitrite and nitrogen dioxide have demonstrated NO-associated lipid damage, which shows the antimicrobial activities associated with NO [1,49,50].
NO interactions with proteins involve reactive thiols, heme groups, iron-sulfur clusters, phenolic or aromatic amino acid residues, tyrosyl radicals, or amines. Peroxynitrite and NO2 also nonspecifically oxidize proteins at many sites. NO can also inactivate the enzymes containing Fe-S clusters (e.g., aconitase, NADH dehydrogenase, succinate dehydrogenase), thereby suggesting that NO• (NO radicals) may cause the release of iron from metalloenzymes and result in iron depletion [1,51,52]. The enzymes DNA alkyl transferases have cysteine residues where the -SH group of cysteine residues reacts with NO, resulting in the formation of S-NO adducts. These adducts, in turn, inhibit the transfer of the alkyl group from guanine to the protein. Thus, NO inhibits DNA repair enzymes, which are particularly concerned with the repair of alkylation to DNA [53,54]. The sensitivity of prokaryotes to NPs treatment is higher because bacteria depend a great deal on iron-sulfur clusters as compared to mammalian cells. Consequently, it seems reasonable that efficient NO-releasing NPs have the potential to be effective against MDR bacteria [1,55]. The in vitro efficacy of NONPs has been assessed against several clinically significant Gram-positive (E. faecalis and S. pyogenes) and negative (E. coli, K. pneumoniae, and P. aeruginosa) isolates. It was found that the reduction in bacterial growth was NONPs dose-dependent for both Gram-negative as well as Gram-positive bacteria. The growth of Gram-negative bacterial isolates was inhibited within 24 h by NONPs; however, Gram-positive bacterial growth was inhibited within 8–16 h with lower NONPs concentrations [1,56,57].
Chitosan NPs are known to be suitable for non-invasive routes of drug administration (nasal, oral, ocular, and pulmonary routes) as they deliver the drug with reduced toxicity, increase the blood half-life of drugs as well as the efficiency of intravenous injections [58,59]. Biofilms produced by P. aeruginosa are one of the main challenges while treating (the skin) infections. In one study, gold nanorods (AuNR) were decorated with 1,2-Distearoyl-sn-glycero-3-phosphorylethanolamine(DSPE) phospholipids and exploited to destroy biofilms produced by P. aeruginosa In vitro. Roughly a ~6 log cycle reduction of the bacterial count was observed by applying DSPE-AuNR against P. aeruginosa, proving the fact that gold-based nanosystem is one of the effective alternatives to antibiotics for the eradication of biofilms [60]. In another report, the effect of hyperthermia of gold nanorods (GNR) against S. aureus and Propionibacterium acnes, the causative agents of acne vulgaris, were evaluated. Local heat was generated when functionalized GNR was excited by a laser beam. It resulted in a ≥99.99% reduction of viable bacterial count [61].

2.2. Nano-Photothermal Therapy of MDR Bacteria

Another technique employed for destroying biological cells is Photothermal therapy with NPs. In this technique, the electromagnetic radiation absorbed by the NPs is converted into heat, which is then transferred via thermal conduction to the bacteria or cells in close proximities [1,18,62]. AuNPs have been extensively studied for photothermal therapy of cancer [1,63,64,65]. It has been shown that pathogenic bacteria can also be selectively killed by using functionalized AuNPs. AuNPs, when conjugated with vancomycin, acquire a polygonal shape, due to which these NPs have the potential of absorbing near-infrared (NIR) light. Amoxicillin coated Au-NPs have increased in vivo stability [1,66,67]. NPs conjugated with vancomycin can effectively kill bacterial cells under illumination (>99%). It has been established that pathogenic E. coli can be photothermally lysed using Au nanorods [1,29,68], and P. aeruginosa can also be effectively killed by the same metallic nanorods conjugated with primary antibodies [1,69].
The bacterial cell viability also reduces significantly when a nanorod attaches to the surface of the bacterial cells and is exposed to near-infrared radiation. In addition, the MDR bacteria are photothermally destroyed by multifunctional popcorn-shaped magnetic iron core-shell gold nanoparticles. Results had shown that when MDR Salmonella DT104 bacterial cells were treated with M3038 antibody-conjugated hybrid platforms, they attached to bacterial cells, and localized heating at 670 nm light irradiation caused irreparable cellular damage and killed the bacteria within 10 min of exposure [70]. MRSA and E. coli had been reported to show a significant decrease in viable counts when exposed to 660 nm for 5 minutes, along with polysiloxane polymers containing embedded methylene blue and AuNPs [1,71,72].

2.3. Silver NPs Bactericidal Effect against Multidrug-Resistant Bacteria

Historically silver has been used for its antiseptic and bactericidal activity in dental alloys and open wounds as well as in Ayurveda and homeopathy. Gram-positive bacteria possess a thick peptidoglycan layer (30 nm) in comparison to Gram-negative bacteria (2–3 nm), and silver NPs are thought to anchor the cell wall leading to structural changes in the cell membrane, thus increasing the cell permeability. Therefore, uncontrolled transport through the cell membrane results in bacterial cell death [34]. In addition, AgNPs produce free radicals for membrane damage and may affect the proton motive force inhibiting the oxidative phosphorylation [55]. AgNPs can release Ag+ ions, which can disrupt cellular functions by interacting with thiol groups of many enzymes rendering them inactive [1,73,74]. Panacek et al. established that the colloidal AgNPs have substantial bactericidal activity against MRSA, Gram-positive and Gram-negative bacteria [55]. Ayala-Nunez et al. observed that the mode of action of AgNPs (100 nm size) is dose-dependent against MRSA and non-MRSA as their growth was inhibited at concentrations over 1.35 mg/mL when the inoculum was 105-CFU/mL [55,75]. Nanda and Saravanan reported the antimicrobial activity of AgNPs synthesized by aqueous Ag+ reduction with S. aureus against MRSA, MRSE, S. pyogenes, Salmonella typhi, and K. Pneumoniae and reported that AgNPs were most effective against MRSA followed by MRSE and S. pyogenes, but only moderate activity was observed against S. typhi and K. pneumoniae [55,76]. Humberto et al. found that AgNPs of concentration 30 to 100 mmol/L is effective against the erythromycin-resistant S. pyogenes, ampicillin-resistant E. coli, MDR P. aeruginosa, and drug-susceptible strains including Streptococcus spp., E. coli, and P. aeruginosa [55,77]. Pal et al. found that triangular AgNPs are more active than spherical NPs, which are again more active than rod-shaped NPs against E. coli. Morones et al. studied the effect of different concentrations of AgNPs (1–100 nm size) on E. coli and concluded that concentration over 75 µg/mL was sufficient for a significant decline in the bacterial progression. Shrivastava et al. found that the mode of action of AgNPs is dose-dependent and is more evidence against Gram-negative bacteria as compared to Gram-positive bacteria [55]. Recently, our group reported that AgNPs synthesized through green routes can check many susceptible and MDR bacteria very effectively [26]. The bactericidal effect of AgNPs against MDR bacteria is illustrated in Figure 4.
The antibacterial activity of NPs is directed by their physicochemical properties, which in turn depend upon the route(s) of synthesis [1,78]. It has been reported that NPs could be synthesized by Fungi like Fusarium oxysporum and Phanerochaete chrysosporium. Magnetotactic bacteria are also well known to biosynthesize magnetic iron oxide nanoparticles [79,80]. Biosynthetically produced AgNPs using fungus, yeast, bacteria, and plant extracts were found to have strong antibacterial efficacy against various MDR pathogens such as M. tuberculosis, P. aeruginosa, S. pneumoniae, MRSA, K. pneumoniae, MRSE, S. pyogenes, Bacillus spp., E. coli and S. typhi [1,77,81,82]. Green NPs have improved antibacterial activity because of their high surface area to volume ratio and surface reactivity as compared to the chemical NPs. In addition, the coating of biosynthetically produced NPs by phytochemicals increased their ability to inhibit bacterial growth in comparison to chemical NPs [1,83,84].

2.4. Aluminum Oxide Nanoparticles

The antimicrobial effects of metallic NPs, particularly AgNPs, have been extensively reviewed, but there is not enough substantial literature regarding the antimicrobial activities of Aluminum oxide NPs. Aluminum oxide NPs, also called alumina, are normally known as corundum, which is the crystalline form of alumina (Al2O3) [33]. They adopt a corundum-like structure as oxygen forms a hexagonal packing, and Al+3 fills two-third of the lattice in the octahedral sites [85]. Alumina NPs possess a positive charge on their surface at almost neutral pH and are thermodynamically stable over a wide range of temperatures. They can disrupt the bacterial cell wall by producing ROS resulting in bacterial cell death [86]. Among the various diverse procedures for the synthesis of Alumina NPs, for instance, hydrothermal processing, sol–gel pyrolysis, sputtering; the laser ablation method is widely used as it is a quick and high-purity process [87]. Aluminum NPs have been documented to have a wide range of antimicrobial activities. Their growth-inhibitory effects on E. coli have been reported at 10–1000 µg/mL [88]. Balasubramanyam et al. have reported high sensitivity of Alumina against P. fluorescence as compared to the bulk materials [89]. In one study, aluminum oxide NPs were proven to be a good antibacterial agent against Gram-positive as well as Gram-negative bacteria. These NPs were synthesized by using aluminum sulfate and NaOH as precursors by the co-precipitation method. Their activity was analyzed against Gram-positive (S. aureus and Streptococcus mutans) and Gram-negative (E. coli and Proteus vulgaris) bacteria [90].
An oxide layer is produced over alumina nanoparticles, which then protects these from oxidation. These particles are affected by different pH concentrations, which lead to their different toxicity levels; for example, at neutral pH, these NPs have a positive surface charge due to which they would have an affinity towards the negatively charged surface of E. coli cells, thereby resulting in adhesion of alumina NPs over the bacterial surface. Moreover, these NPs also can serve as radical scavengers and cause distortion in bacterial cells [91].

2.5. Silicon NPs

Silicon dioxide (SiO2) is one of the important industrial additives with several applications. It is commonly used as a semiconductor in electronics when in the crystallized form [92]. The conversion to metal oxide increases the surface area, thereby providing better performance in applications. Jiang et al. have shown the binding of SiO2NPs to the bacterial cell walls as well as their higher toxicity as compared to their bulkier counterparts [93]. As limited data are available regarding the antibacterial activity of SiO2NPs, more research is needed to be done to understand their clear role. For treating multidrug-resistant M. tuberculosis, ethionamide (ETH) is one of the most important drugs used. In one study, it was reported that loading ETH into thermally carbonized-porous silicon (TCPSi) nanoparticles would result in enhancement of solubility and permeability of ETH at different pH-values as well as increased its metabolization process. It was found that ETH-conjugated SiNPs tend to reduce the dosing frequency of ETH for the treatment of multidrug-resistant M. tuberculosis [94]. Moreover, it was demonstrated in a study that porous silicon nanoparticles have the potential to be used as a means of a prolonged drug delivery system [95].

2.6. Gallium Nanoparticles (NPs)

There is a growing problem of drug-resistant strains of M. tuberculosis, due to which there is an urgent need for new treatments and novel drugs. In one study, targeted drug delivery using gallium (III) nano-formulations were used against drug-resistant M. tuberculosis, which showed promising anti-tuberculous activity. They also promoted maturation of the phagosome, which in turn result in the increased macrophage-mediated killing of the organism.
M. tuberculosis requires iron for growth and replication, gallium encapsulated in nanoparticles interferes with the cellular iron acquisition and utilization, which in turn inhibit the growth of this bacteria human monocyte-derived macrophages (MDMs). Delivery of Ga in the form of nanoparticles to macrophages open new pathways for the development of new therapeutic anti-tuberculous drugs [96]. Similar to M. tuberculosis, iron and gallium encapsulated in nanoparticles can also be used against HIV. GaNPs are readily internalized by the MDMs, and then sustained drug release causes significant growth inhibition of both HIV and M. tuberculosis. Iron-mediated enzymatic reactions are interrupted by GaNPs, which leads to growth inhibition of HIV–M. tuberculosis coinfection in macrophages, and it also modulates the release of cytokines that may contribute to HIV-TB pathogenesis [97]. Table 1 summarizes different nanoparticles that can be exploited as a weapon against multidrug-resistant bacteria.

3. Host Defense Peptides (HDPs)

The MDR infections are burgeoning at an alarming rate, and hardly any discoveries are taking place in the manufacturing of novel antibiotics to treat such stubborn bacterial infections. The seriousness of this matter creates a need for substitute strategies to treat bacterial infections. Host defense peptides (HDPs) are considered as some effective alternates that would help fight the resistant bacterial infections. These short cationic molecules are formed by the immune systems of many multicellular organisms. The evolution in nature has resulted in the formation of some remarkable HDPs, which possess diversity in structure as well as in biological activity. These natural peptides can be used as templates to generate a single synthetic molecule having the combined properties of antimicrobial and immunomodulatory compounds to fight the resistant bacteria when existing antibiotics fail to function. Defensins and cathelicidins are the two main families of the naturally existing HDPs [100].
Two novel peptides, brevinin1 HYba1 and brevinin1 HYba2, had been isolated from frog (Hydrophylax bahuvistara) skin secretions, and their hemolytic, cytotoxic, and antibacterial activities were investigated after designing acidic and amidated analogs. All the peptides, excluding acidic analogs, showed promising antimicrobial activity against tested Gram-positive and Gram-negative bacteria. These peptides also showed very low hemolysis on human erythrocytes. This study opened up an area to explore more natural sources of host defense peptides, which can be used as effective therapeutic agents [101].

4. Defensins

Defensins are cationic amphipathic peptides having approximately 30 amino acid residues. The three disulfide bonds in the structure stabilize the triple-stranded antiparallel β-sheet assembly [102]. Based on the arrangement of disulfide bonding, defensins are further divided into subfamilies: α, β, and θ. Among mammals, only human neutrophils and leukocyte granules possess α-defensins [103]. In most mammals, Paneth cells of the intestines are responsible for the production of these defensins [104]. They are synthesized as precursors initially and can become activated when the N-terminal segment is removed with the help of trypsin in humans (Figure 5). The concentration of α-defensins reaches 10 mg/mL when they are stimulated by microbes. This concentration is enough to tackle a resilient microbial infection [100].
Most of the epithelial cells express β-defensins; proinflammatory stimuli and infections are responsible for this expression of β-defensins. They can be found in mucosal sections of gastrointestinal, respiratory, and urogenital tracts as well as in the inflamed skin [100].
The θ-defensins are the rarest of all the three defensins and are cyclic molecules. Due to the cyclic structure of θ-defensins, their microbicidal action is resistant to the concentration of salt. θ-defensins are absent in mammals, including humans.
A variety of different defensins have also been identified in different fungi. In one study, 68 fungal defensin-like peptides (fDLPs) from five genera named Apophysomyces, Trichosporon, Scedosporium, Beauveria, and Lichtheimia had been reported. A new synthetic defensin called scedosporisin had been characterized, which shows good activity against Gram-positive bacteria. It killed several vancomycin-resistant Enterococci and MRSA while it showed less cytotoxicity and hemolysis. It was found out that scedosporisin-2 killed bacteria more rapidly as compared to the antibiotic vancomycin [105].

5. Cathelicidins

Cathelicidins are the second major group of HDPs and are categorized based on a production mechanism rather than a sequence match. The inactive precursors of cathelicidins consist of N-terminal cathelin-like domain, which is followed by a peptide region. These precursors are proteolytically cleaved to become mature and active HDPs [106]. Cathelicidins differ in sequence, length, as well as in structure. They have lengthy α-helical and β-hairpin folds along with some short linear molecules. These short 13 amino acid molecules are the initiators for designing synthetic peptides that have optimized biological activity.
Many types of cathelicidins, such as bactenecin, indolicidin, protegrins, and many others, are produced by the immune systems of bovine and porcine [106]. The human immune system is known to produce only one type of cathelicidin precursor protein hCAP18. This precursor is processed proteolytically to produce a mature cathelicidin LL-37 [107]. Disulfide bonds are absent in LL-37. However, it adopts the conformation of α-helical when it interacts with the lipid bilayers. Mice are known to have only one cathelicidin precursor, which is also proteolytically cleaved to produce a mature form, CRAMP. This mature peptide has a sequence identity of 67% with LL-37 [100].
S. aureus is responsible for many serious infections in humans that sometimes lead to sepsis or death also. In one study, six novel cathelicidins named CATHPb1–6 were identified from Python bivittatu. CATHPb1 was found with an excellent pharmacological and toxicological profile In vitro. Later on, it had been observed that CATHPb1 provides efficient protection to mice against MRSA/VRSA. CATHPb1 was found to be involved in rapidly modulating macrophages/monocytes as well as trafficking the neutrophils to the site of infection and also enhance their bactericidal functions. It also increases the levels of chemokines and reduces the release of proinflammatory cytokines. Therefore, it proved to be a novel therapeutic agent against MDR S. aureus [108]. Table 2 shows HDP and their target bacteria.

6. Antimicrobial Peptides (AMP)

Antimicrobial peptides (AMP) are evolutionarily conserved macromolecules produced by most living organisms ranging from prokaryotes to humans as a first line of defense. AMP are a part of innate immune response and have an ability to fight against pathogenic microbes. These small peptides ranging from 05 to 100 amino acid residues are generally cationic in nature and are folded into unique structures that facilitate their mode of action. AMP help to eradicate (pathogenic) bacteria either by killing them directly or by modulating the host immune response. They can be used against a number of microbes and have proven to be a promising agent when used as antibacterial either alone or in combination with other methods [102,106,109].
Antimicrobial photodynamic therapy is one of the novel approaches, which, in combination with antimicrobial peptides, has a great potential to act as a favorable tool against MDR bacteria. Photodynamic therapy is quite effective against Gram-positive bacteria, but combinational therapy with AMP makes it potent against Gram-negative too. Photodynamic AMP generate reactive oxygen species upon exposure to light of a certain wavelength and disrupt the cell wall and membrane resulting in cell death [110].
Biofilms are three-dimensional multicellular structures that form on natural and/or clinical surfaces. Biofilms are adaptively resistant to antibiotics, due to which they are difficult to treat as compared to their planktonic forms [111]. They are formed on various implanted devices, and these aggregates can only be removed by surgery [112]. Therefore, it is very challenging to treat biofilm-related infections; there is an urgent need for new therapeutic options to fight them out. In recent years, various approaches had been developed, such as bacteriophages, antibodies, quorum sensing antagonism, etc. One of the most promising approaches is the use of antibiofilm peptides (ABP), which are a class of the AMP; these proteinaceous entities can either be cationic or amphipathic [113] and are a part of the host defense peptides. The first-ever cationic peptide, nisin, was isolated in 1928 from Lactobacillus lactis. It was relatively stable at room temperature, but at pH 2–6, it showed high antimicrobial activity [114,115]. Natural AMP polylysine was isolated from Streptomyces albulus 346 and is now commercially produced for a variety of food applications as preservative agents [116]. ABP shows activity against a variety of resistant Gram-positive and Gram-negative bacterial strains. They are also found to be very effective against fungal microbes. Interactions of these peptides with bacterial components do not require any specific protein binding sites, and it is why they would theoretically not have any resistance emerged against them [109]. Recently, they have been put to work in many applications, including oral candidiasis, catheter-associated, and implant surface infections [117,118]. The human cathelicidin peptide LL-37 was shown to exhibit antibiofilm activity against S. aureus and E. coli [119]. The two tryptophan-rich cationic AMP KT2 and RT2 were found to show antibiofilm activity against enterohemorrhagic E. coli O157:H7, which is a multidrug-resistant strain. These two peptides did not only prevent the biofilm formation but also could eliminate mature biofilms [120]. It has also been reported that these peptides can be used in conjugation with other antimicrobial compounds to enhance their activity (synergistic effects) [121]. This synergy helps in lowering the concentration of antimicrobial compounds, which will, in turn, reduce the (toxic) side effects of these compounds and stop the spread of antimicrobial resistance [122]. Most commonly, these peptides permeabilize bacterial cell membranes, which lead to the death of cells either by causing large damage or small obstructions that will, in turn, disturb transmembrane potential leading to cell death [123]. Specifically, their mechanism of action has been explained by pore and non-pore models. For pore models, there are two theories; the toroidal pore model, in which antimicrobial peptides can affect the curvature of membrane and the barrel stave pore model, in which these will interact with the cell membrane and form a hydrophilic pore [124,125]. For non-pore models, there are many theories, such as the detergent model, the molecular shape model, and the carpet model [126]. Among these, the carpet model is the most common model. According to this model, peptide monomers form a layer on the surface of the membrane, which leads to the destabilization of a phospholipid bilayer, which then results in the breakdown of the membrane [127]. Much research has been carried out to develop antimicrobial peptides as effective antimicrobials, but hurdles are there because these have complex interactions with membranes as well as with each other. Very limited data are available on such peptides having antibiofilm properties. Therefore, more work is needed to be done to understand the proper mechanism of action [128]. However, we have shown various possible ways ABPs can get over a superbug and kill it when it resists antibiotic drugs in Figure 6.
ABP can also be obtained from the poisons of various animals like ants, wasps, bees, scorpions, and spiders. One such peptide, called Mastoparan peptide, isolated from Vespidae venom showed broad-spectrum antimicrobial activity against both Gram-positive and Gram-negative bacteria, Mycobacteria and fungi. In one study, two peptides, agelaia-MPI and polybia-MPII, isolated from wasps, showed bactericidal activity along with antibiofilm activity against biofilm-forming MDR Acinetobacter baumannii [129]. Five ocellatin peptides, ocellatin-PT2–PT6, had been isolated from frog Leptodactylus pustulatus skin secretion and used against an MDR opportunistic pathogen, P. aeruginosa, where they effectively killed the bacterial pathogen. Another ocellatin peptide, named ocellatin-PT3, inhibits the proliferation of established biofilms by direct killing of bacterial cells within biofilm [130].
The EPS of some bacterial species like non-typeable H. influenza, S. enterica serovar Typhimurium/Typhi and P. aeruginosa are responsible for the resistance against innate immune components, including AMP. This resistance is particularly due to the structure of the biofilm community. The polysaccharides and extracellular DNA (eDNA) of EPS bind the AMPs because of charge differences [131]. There are specific sensors in bacteria that are responsible for activating the resistance mechanisms against AMP upon exposure [132,133]. It had been demonstrated that in P. aeruginosa, the psrA gene encodes a transcriptional regulator which upregulates in response to the presence of subinhibitory concentrations of cationic AMP [134]. Table 3 describes possible mode(s) of action of (AMP) ABP with their advantages and limitations.

7. Bacteriophage Therapy

Bacteriophages are diverse non-living biological entities that consist of DNA or RNA surrounded by a protein capsid. They are capable of reproducing independently and are ultimately dependent on bacterial hosts for survival. Phages normally bind themselves to specific receptors on the bacterial cell surface, release their genetic material into the host cell and then either incorporate this material into the bacterial genome and reproduce vertically from mother to daughter cells or invade the bacterial replication mechanism to produce the next-generation of phage offspring and lyse the cell. When a critical mass of phage offspring is reached, which can be from a few to over 1000 viral particles, depending on environmental factors, the lytic proteins are activated and hydrolyze the peptidoglycan (cell wall) of bacteria [137]. Scientists suggested that phages can be used as suitable antibiotic agents with having maximum efficiency.
Treatment of infectious diseases is becoming difficult and a threat to mankind due to a rise in antibiotic-resistant microbial strains [19]. To prevent and treat such resistant strains, phage therapy is becoming popular and gaining interest all over the world [138,139]. Pioneering (novel) antimicrobial approaches using phage products, or genetically manipulated phages, are being exploited to cope with bacterial infections and antibiotic resistance (Figure 7) [139]. Phages infect bacterial cells and produce endolysins that damage the bacterial cell wall by hydrolyzing the four main bonds of its peptidoglycan constituent in the lytic cycle [138,140,141]. An osmotic imbalance is the cause of lysis when the cell losses structural integrity upon peptidoglycan degradation. In the case of Gram-negative bacteria, the outer membrane is ruptured with the help of complexes (spanins), fusing both the inner and the outer membrane [139].
Bacteriophages have been successfully used against bacterial biofilms, in therapeutics (genetically modified form), in the food industry to minimize the bacterial load and to improve antibiotic potency (Figure 7c) [138,142]. Antibiotics delivered along with the phages permit delivery to specific cells and can cause an upsurge in local drug concentrations [139].
Bacteriophages are used both externally and internally to treat diseases that could otherwise not be cured by antibiotics [138]. For example, in a study, phage application significantly decreased the concentration of bacterial cells (in all patients’ sputum samples), improving overall health. Birds had also been reported to have a reduction in a load of Salmonella and Campylobacter in the poultry meat when a multivalent/cocktail of lytic bacteriophages was used to help the meat industry to produce safe and good quality edible products [138,143].
Bacteriophages have several advantages over antibiotics, i.e., they have an affinity for a specific bacterium, which helps in typing of that particular bacterium and causes its lysis, whereas using the (broad-spectrum) antibiotics would also harm the normal flora. Due to their replicative nature, there is no need to administer the bacteriophages repeatedly, and Most of the phages can be ingested as they can survive in the gastric environment [140,144] and are even lethal to MDR bacterial superbugs: E. faecalis, S. aureus, Klebsiella, A. baumannii, P. aeruginosa, and Escherichia coli [19]. Because of minimal side effects, they are also considered to be ecologically safe (i.e., harmless to humans, plants, and animals) [19,138,145].
However, the disadvantages of phage therapies must not be overlooked either. In the case of mixed infections, phages are not very effective due to their narrow host range, whereas broad-spectrum antibiotics work effectively in similar scenarios [138,146]. If the selected phage switches to the lysogenic life cycle, it will integrate its genome in the host cell leading to failure in phage therapy. Moreover, if phages are not sequenced before their administration in humans or animals, they may carry a toxic (deleterious) gene that can be harmful [138]. We should always use a cocktail of phages so that the risk of resistance development could be reversed or diminished [19,147].
The frequency of bacterial resistance to phages is significantly lower (10−7 to 10−8 per cell) compared with that of the resistance to antibiotics (frequency of mutation for one specific gene is 10−5 per cell). Bacterial cells can also become resistant to phages as their cell receptors are specific to different phages; constructing the “suitable” cocktail will help to achieve the maximum effectiveness of phage therapy [19].
Recent investigations using animal models have explored phage treatments against different bacteria, which have shown positive results. When challenged with gut-derived sepsis due to P. aeruginosa, oral administration of phages saved 66.7% of mice from mortality in comparison to 0% in the control group. In a hamster model of Clostridium difficile-induced ileocolitis, a single dose of phage synchronized with C. difficile administration was sufficient as prophylaxis against the infection; phage treatment post-infection saved 11 of 12 mice, whereas control animals administered with C. difficile and clindamycin died within 96 h [148].
Phage combinations also lowered C. difficile growth significantly in in vitro and limited proliferation in vivo using a hamster model [149]. Intraperitoneal administration of a single phage strain was effective to rescue 100% of mice in bacteremia models using vancomycin-resistant E. faecium [150], extended-spectrum β-lactamase producing E. coli [151], and imipenem-resistant P. aeruginosa [152].
Phage cocktails can be utilized in treating antibiotic-resistant P. aeruginosa infections of the skin, lungs, and gastrointestinal tract in animal models [153,154]. Reports also suggest that phages have the potential to restore sensitivity in antibiotic-resistant bacteria like the case of multidrug-resistant P. aeruginosa. Phages are combatting antibiotic-resistant bacteria by limiting their capacity to evolve resistance [155,156].
Unlike antibiotics, phages may evolve novel counter-defense mechanisms to counter bacterial resistance at a rate that can never be replicated by researchers developing antibiotics [157,158,159,160].
A. baumannii is a nosocomial pathogen that is rapidly evolving resistance against antibiotics. Two novel bacteriophages, named PBAB08 and PBAB25, were used against the MDR A. baumannii in a mouse model. Mice treated with a phage cocktail showed a 2.3-fold more survival rate along with a 1/100 reduction of the total number of A. baumannii in the lungs [161]. A person infected with MDR A. baumannii was treated with bacteriophages, and the patient was reported to be more alert than before, his craniotomy site and skin flap healed very well. He had lost all the symptoms of the infection and got healthy [162]. These findings suggested that the newly isolated phages could be used as effective therapeutic candidates against (MDR) A. baumannii.
Phage lysins are solely capable of lysing bacterial cells, and they have been identified as potential antimicrobial agents. These proteins are efficient, potent, and inactive against eukaryotic cells. Mice have been successfully saved from bacteremia through lysins caused by MDR A. baumannii [163], Streptococcus [164], and MRSA [165].
Using phage lysins and antibiotics in amalgamation may prove more effective at eradicating infections than by using antibiotics solely, as displayed in vitro and in vivo in a colon model using C. difficile [166]. All lysins do not show equal therapeutic potential, however as highlighted in an article [167], a highly potent lysin, PlySs2, was identified, which proved very effective against several pathogenic Streptococcus and Staphylococcus species, such as MRSA, and remained fully operational even after 10 freeze-&-thaw cycles. A recent study about the isolation and application of phage proteins has shown that lysins can cross the epithelial cell membrane to eradicate intracellular infections of S. pyogenes [168]. Phage lysins also interrupt vegetative cells, as displayed with B. anthracis lysin PlyG, which has the potential of attacking endospores of Bacillus, a major advantage over antibiotics. Table 4 enlists and summarizes different phages and phage cocktails used against resistant bacteria.

8. Immune Stimulation via Bacterial Extracts

Bacteria and bacterial extracts have been used for immunotherapy for several years. Recently, it has been discovered that these nonspecific immune activators trigger specific receptors of immune cells (and certain molecular signaling pathways); opened a new era of targeted immunotherapy. It can be achieved by using chemically synthesized molecules mimicking specific pathogen molecules. Bacterial extracts contain specifically and nonspecifically stimulating agents that activate innate and the adaptive immune system [171]. For the treatment of complicated as well as (other) infections that are caused by resistant strains, bacterial extracts can be used as adjuvants [172]. Bacterial extracts can activate macrophages and monocytes due to the presence of bacterial wall structures, like lipopolysaccharide or proteoglycans, which interact with Toll-like receptors (TLR) that are expressed over the surface of monocytes. Due to this interaction, monocytes are activated; they will differentiate into immature dendritic cells and then mature into dendritic cells, which are considered as suitable antigen-presenting cells, and the activation of this mechanism would cause a stimulation of the immune response. The antigen presentation over mature dendritic cells, in turn, stimulates T helper and B lymphocytes, following the maturation into plasma cells along with antibody production. These antibodies then cause bacterial opsonization, followed by destruction via macrophages [173].
Lipopolysaccharide (LPS) is a well-known virulence factor of bacteria that stimulates an innate immune response in hosts. In one study, LPS was extracted from the highly resistant isolates of Proteus mirabilis and incorporated into a liposomal delivery system. It was injected in rats via the intraperitoneal route, and its efficiency in stimulating immune responses was weighed by determining the Toll-like receptors and CD14 levels. The results showed that liposomes having incorporated LPS could release moderate levels of Toll-like receptors-4 (TLR4) that, in turn, enabled the immune system to clear pathogens [174].

9. Vaccination

Vaccination had been so fruitful in the world of multidrug resistance owing to its mechanism of action that diminishes the burden of the disease, thereby reducing the use of antibiotics and thus culminating the basis of MDR evolution. Since antibiotics are not being utilized (in this case), the resistance cannot develop, and selection cannot occur; neither of the pathogen at hand nor the “bystander” species present [175,176]. The vaccination process can considerably influence the MDR better if herd immunity is conferred [175,177], that is, protection of unvaccinated population by the vaccinated population such that they act as buffers, not being affected by the disease themselves and thereby reducing the chance of transmission to others. Moreover, for vaccines against bacteria like S. pneumoniae, S. aureus, and members of the family Enterobacteriaceae, which inhabit the nasopharynx, skin, gastrointestinal tract; there is the hypothetical likelihood that plummeting the density of bacterial populations by vaccination diminishes the prospects for genetic reassortment and recombination of resistant genes [175,178,179]. Vaccines that are generated specifically against virulence factors can be very powerful because these days, there are numerous genomic sequences available for almost all species. The state-of-the-art technologies and methods, such as reverse vaccinology, which can filter out the best models, are suited for enhanced immune responses. Some vaccines that are currently under production directing the aim towards virulence factors: (i) a tetra-subunit vaccine comprising of two capsular polysaccharides and two virulence-associated proteins (ClfA and MntC) against S. aureus, which is presently in phase 2b trials [175,180], (ii) three vaccines against C. difficile constructed on toxins A and B which are in phase 2 and 3 trials [181], (iii) a vaccine against P. aeruginosa founded on conserved outer membrane protein F/I fusion which is in phase 2/3 trials in an intensive care unit (ICU) patients [175,182], and (iv) a vaccine for Candida targeting T cell target protein, Als3 [175,183,184] is in the phase 2 trials. In recent times, it has been anticipated that directing vaccines against already resistant strains or even against resistance determinants themselves may be an actual way to thwart the selection pressure for antimicrobial resistance [23,175,185]. Antiresistance vaccines ought to be more operational against the drug-resistant strains in comparison to drug-susceptible ones by explicitly targeting resistant alleles of a conserved protein (for instance, a neuraminidase binding protein in the influenza virus) or by targeting proteins exclusively present in resistant clusters (such as ribosomal methylases deliberating macrolide resistance). Two vaccines are being developed under the shadow of this theory; that is, the resistance causing elements can be the foundation of vaccines providing a strong immune response. Protection against MRSA [175,182,186,187] can be sought through a vaccine that aims for the resistance-conferring extra penicillin-binding protein (PBP2a), while in Neisseria meningitidis, the target would be one of the essential penicillin-binding proteins [175,188]. The use of a variety of vaccines that may provide immunity against infectious agents is summarized in Table 5.

10. Combination Drug Therapy

Combination therapy is when a set of drugs are used to treat infections rather than a single drug (monotherapy). Infections caused by causative agents (M. tuberculosis, human immunodeficiency virus, Plasmodium parasite), which are predisposed to develop resistance, are being treated with this method [147]. The use of combination drug therapy acts in multiple dimensions. Recently, for the treatment of gonorrhea, which recommends ceftriaxone or cefixime plus azithromycin [191].

10.1. Combination Drug Combination Acting on Diverse Targets in Different Pathways

A classic example is a treatment modality used against M. tuberculosis infections currently prevalent in many developing nations like India. Four first-line drugs are used in this regimen: rifampicin (R), isoniazid (H), ethambutol (E), and pyrazinamide (Z); their targets are rifampicin (RNA polymerase inhibitor), isoniazid (enoyl reductase subunit of fatty acid synthase), ethambutol (an inhibitor of arabinosyl transferases involved in cell wall biosynthesis) and pyrazinamide (mechanism of action poorly understood) [147,192,193]. This method is highly effective since a bacterium may develop resistance by changing one of its targets; the combination drug strategy will still be effective against at least the other two pathways minimizing the chances of bacterial propagation.

10.2. Drug Combinations Acting on Diverse Targets in the Same Pathways

β-lactamase enzyme produced by Gram-positive bacteria opens up the β-lactam ring making it non-functional. Thus, this approach involves the use of a β-lactam antibiotic (amoxicillin) and β-lactamase enzyme inhibitor (clavulanic acid) [147,187]. Clavulanic acid degrades the enzyme, allowing the drug to destroy these microorganisms.

10.3. Drug Combination Acting on a Single Target, but in Multiple Dimensions

Streptogramins are made up of two active molecules that bind to the adjacent sites in the 50S ribosomal subunit near the peptidyl transferase center [194,195]. When both of these molecules are used simultaneously, they show 10–100-fold more potency as compared to using a single molecule alone [196]. Table 6 defines different combinations of drug therapy.

11. Novel Antibodies against MDR Bacteria

The conception of monoclonal antibody therapy against MDR superbugs stems from the fruitful usage of serum therapy against bacterial infections. The effectiveness of this technique was authenticated in the clinical trials and is in medical practice since the early 1900s entailing a particular antiserum serving as antimicrobial agents. Serum therapy was aborted with the advent of antibiotics, in some measure, owing to frequent toxicities and the failure to refine or generate antibodies for single determinants back then. Conversely, in the present day, technological and research progress in antibody engineering makes it conceivable to produce distinct, consistent, and completely human (humanized) mAbs with a particular antigen specificity.
A1102 is a humanized mouse gal-III mAb whose biologic activity in vitro and in vivo was demonstrated in experimental models of K. pneumoniae ST258 infections. It was revealed that upon passive immunization with A1102 before infection with ST258 whole bacteria or ST258-derived LPS increased the survival rate of endotoxin-sensitized mice and also protected rabbits from a lethal infection with ST258. It was shown by in vitro studies that the biological action of A1102 comprising complement and Fc independent LPS neutralization necessitated divalent binding and augmentation of human serum bactericidal eradication and complement-dependent macrophage uptake of ST258 [197].
It had been demonstrated that O25b-specific MAb ASN-4 retained its bactericidal activity against an MCR-1-positive colistin-resistant ST131-H30 strain by three mechanisms of action that are opsonophagocytosis, endotoxin neutralization, and complement-mediated killing. Subsequently, LPS O-antigen-targeting antibodies are thought to be an alternative way of combating MDR infections, including the emerging MCR-1-positive isolates [198]. The general mode of action of novel antibodies against bacteria is depicted in Figure 8.
Even though antibody-based approaches of prophylaxis may have substantial potential for the inhibition of bacterial infections, their conspicuous operating principle may correspondingly counterpart antibiotics. Combination drug therapy is broadly considered to diminish the occurrence of resistance, but antibacterial antibodies have likewise been established to deliver assistance when used alongside antibiotics comparative to management with antibiotics single-handedly. For instance, mice doctored with a comparable human dosage of tobramycin did not endure infection with a tobramycin-resistant P. aeruginosa experimental isolate; however, therapy with tobramycin and a sub-protective MEDI3902 (bispecific antibody targeting the P. aeruginosa type III secretion (T3S) protein PcrV and Psl exopolysaccharide) dosage ensured in mice survival and improved bacterial clearance. Comparable advantages have been described employing using mAbs adjunctively with antibiotics against bacterial infections [199]. Table 7 defines different antibodies against bacteria.

12. Carbon Monoxide-Releasing Molecules (CORMs)

Studies have been conducted on animal models regarding the usage of hem oxygenase activity intracellular product called carbon monoxide (CO), and its administration as a therapeutic agent showed beneficial effects in treating animal models suffering from inflammatory disorders and cardiovascular diseases as well as in the organ transplantation [200].
However, the same administration of CO in humans can be hazardous as the levels of systemic carboxyhemoglobin can be raised high enough to cause death. Prodrugs such as carbon monoxide-releasing molecules (CORMs) transfer CO into biological systems. They are considered a safer alternative since they do not affect the transport of oxygen by hemoglobin.
Among all the CORMs reported up to now, metal carbonyl complexes are the most suitable and popular class of compounds. Some other CORMs include oxalates, tertiary aldehydes, silacarboxylates, and boron carboxylates [201]. When the carbonyl groups are attached to the transition metals such as molybdenum, iron or ruthenium, the CORMs exhibit a unique ability to transfer CO in the cells and enhance the signal transduction mechanisms mediated by CO. Thus, CORMs, as donors of CO, presented an extensive range of biological activities. Moreover, some specific transition metal carbonyls have repeatedly presented cytoprotective properties as well as some curative activities in a wide range of cellular and in vivo models of diseased animals [202,203].

13. Probiotics

Probiotics are known to be live microorganisms, primarily belonging to the genus Lactobacillus and Bifidobacterium, which are well-known to have a very advantageous influence on the host organism they inhabit. Concerning antibiotics and MDR, probiotics are seen to participate in diminishing the threat of various infectious diseases, including gastroenteritis and respiratory tract infections. Simultaneous utilization of probiotics with antibiotics has been proven to lessen the occurrence, time period, and/or ruthlessness of antibiotic-linked gastroenteritis. This, in turn, has better effects on the adherence properties of the antibiotic, thereby enhancing the efficacy of these antibiotics. Lactobacillus strains are also known to be successful in the defense of the host against urinary tract infections. L. rhamnosus GR1 has an amazing capability to bind with epithelial cells, particularly in the vaginal tract and is not susceptible to spermicidal agents; therefore, the bacterium can avert binding and development of urinary microorganisms [204]. P. aeruginosa has intrinsic resistance to the majority of accessible antibiotics, including aminoglycosides, anti-pseudomonal penicillins, newer cephalosporins, and imipenem. The favorable effect of probiotics may be linked to their aptitude to inhibit the development of drug-resistant bacteria seemingly by the secretion of antibacterial chemicals including lactic acid, hydrogen peroxide, and more diminishing their ability to colonize the body, thereby reducing the use of antibiotics and therefore, the development of MDR as a whole [20,21].
In one current study, indigenous probiotic Lactobacilli and standard Lactobacillus strains were evaluated for their inhibitory activity against MDR K. pneumoniae. Probiotic lactobacilli strains; L. plantarum LMEM7, L. rhamnosus LMEM9, L. acidophilus LMEM8, and L. animalis LMEM6 isolated from curd samples and L. fermentum MTCC 9748 standard strains were used against K. pneumoniae. Results obtained showed that indigenous Lactobacilli could be used against MDR K. pneumoniae in place of antibiotic therapy, and more probiotic strains should be identified against different pathogens [205]. In a similar study, the antimicrobial effect of Propionibacterium freudenreichii derived from dairy had been identified against multidrug-resistant Salmonella heidelberg (SH) in turkey poults. Two strains were used; namely, P. freudenreichii freudenreichii B3523 (PF) and P. freudenreichii shermanii B4327 (PS). The analysis revealed that P. freudenreichii could be used as an alternative to antibiotics for preventing SH infections in poults [206]. Table 8 indicates probiotics use to combat bacteria.

14. Quorum Sensing

Bacteria interact with each other through small chemical molecules (in a coordinated manner) known as quorum sensing. The buildup of quorum-sensing signals (QSS) in the growth medium reflects cell density, and as soon as a viable concentration is achieved, the QSS trigger transcription factors that in turn upregulate the signal synthase and several other genes. It is accepted that QSS govern virulence factor manifestation, which is our main interest coupled with bioluminescence, making biofilm, admission into stationary phase, sexual conjugation, sporulation, and transformation capability [22].
A huge number of hospital-acquired ailments are associated with infections instigated by biofilm molded on implanted devices. The consequence of these infections can be lengthier hospitalization, surgical operations, and even demise. A peptide molecule is known to hinder biofilm formation and ailments caused by S. aureus, which somehow blocks the quorum sensing (coordination) between bacteria. The heptapeptide, which was initially isolated from post exponential supernatants of S. xylosus, is now made in its amide form as a synthetic 7-aa molecule (YSPWTNF-NH2) termed “RNAIII-inhibiting peptide” (RIP). RIP inhibits cell adhesion and biofilm formation; the activity of the gene locus agr, thus preventing the production of a regulatory RNA molecule, RNAIII, that controls the production of toxins and affects the pathogenesis of S. aureus [135,136]. The mechanism through which RIP inhibits quorum-sensing mechanisms includes inhibition of the phosphorylation of a protein called “target of RNAIII activating protein”. The antibiotic concentration essential to destroy bacteria in the biofilm is 100–1000 times greater than that required to kill the same species outside a biofilm. The greater the use of antibiotics, the more the chances of resistance development against them. Thereby, by reducing the use of antibiotics achieved via quorum sensing inhibition, we can effectively prevent MDR in bacteria [135,207]. Figure 9 represents the mode of action how a quorum quencher can interfere with bacterial communication leading to biofilm inhibition.

15. Vaccines vs. Drugs: Who Is Going to Win?

Vaccines are majorly used as prophylactics as opposed to drugs. Vaccines confer immunity to individuals from the pathogens at hand, thereby preventing it from reaching massive numbers in the body, which in turn diminishes the buildup of genetic diversity or recombination frequency as well as culminates any chances for further communication to other hosts. For instance, a tuberculosis vaccine has been reported to reduce a bacterial population peak size up to five times the actual number attained in rodent subjects [24,208]. Furthermore, vaccines are designed as such to encompass more than one pathogenic antigen as well as multiple epitopes for the immune system to recognize and develop memory against [24,209]. Epitopes are recognized and processed by apparatuses of the immune system comparable to the method of biochemical molecules interacting with a drug. This shows that the immune response is actually like combinational drug therapy. With significant additional component effectors, however, it is revealed that roughly 100 distinctive tetanus-toxoid-specific antibodies can be witnessed in hale and hearty humans after getting a tetanus-toxoid booster vaccine, with distinct antibodies from one individual to another [24,209,210,211].

16. How Important Is the Accurate and Rapid Detection of the MDR Bacteria?

As already discussed in detail that antimicrobial resistance is a worldwide issue that leads to morbidity and mortality. This situation can be resolved by developing rapid diagnostics tools for quick profiling of pathogens, and their resistance [212]; one-way to fight this issue is by prescribing accurate antibiotics. This can be done by proper and rapid differentiation of bacterial and viral infections, and it is difficult to achieve because pathogen isolation, identification, and antibiotic resistance detection take time. This can be resolved by implementing point-of-care (POC) diagnostics. There are three classes of POC diagnostics; one that differentiates between bacterial and viral infections; second, in which detection and report of a specific pathogen are performed, and the third one is, it should not only detect and report a specific pathogen but also exhibit the presence of antibiotic resistance [213].
Many bacterial detection techniques, like infrared light-based devices, polymerase chain reaction (PCR), enzyme-linked immunosorbent assay (ELISA), and chemical assays, are available but are slow and require sophisticated equipment. Culturing is considered the gold standard for the detection of bacteria, but it requires several days to retrieve the final results. Hence, there is a need for those techniques that are not only fast but also sensitive in detecting bacteria. In a new method called whispering gallery mode (WGM), optical microcavity-based sensors were used to detect the bacteria Helicobacter hepaticus. It was shown that this bacterium could easily be detected using WGM optical microcavity-based sensors [214]. WGM sensors can react to environmental perturbations. Their sensitivity, together with the diversity of structures, leads to the development of these devices for a wide range of analytes. These are being used for detecting clinically relevant biomolecules as well as single-protein interactions [215]. Thus, this technique is proven to be faster and potentially sensitive.
Optical (bio)sensors facilitate us by monitoring polarization, intensity, phase, speed, and frequency of light. Whispering gallery mode (WGM) microresonators are at the forefront for over the last two decades [216]. They provide a label-free optical method to detect bacteria rapidly with high sensitivity [217]. They are increasingly used as transducers for detecting specific biomolecules. They can detect biomolecules in a label-free manner without any chemical modifications of the analyte. Additionally, the biophysical properties of biomolecules do not get altered due to probing light. WGM sensors detect biomolecules by recognizing the target analyte through molecular receptors. Thus, WGM frequency shifts that are specific to the analyte molecule are produced. Biomolecular detection is achieved by analyzing the binding of analyte molecules to the receptors and converting the binding event into optical and electrical signals, which are then detected and measured as spectra over time. Receptor molecules that are being used in WGM biosensing are oligonucleotides, antibodies, and proteins [218]. Due to the small mode size of the microcavity, the light which enters the microcavity can cycle multiples of times under TIR (total internal reflection) action. In the meantime, this light extends in the surrounding medium in the form of evanescent waves, and the intensity of the evanescent wave falloffs exponentially. Mode change occurs as a result of any interference in the evanescent field, and this results in substantial variations in the resonant characteristics like mode separation, frequency shifting, and linewidth broadening. This causes the enhancement of interaction between light and matter, which results in realizing unprecedented sensitivity detection [219].

17. Conclusions

Bacteria are naturally programmed to survive, and it is why they had always been one step ahead of humans. Nevertheless, it is high time to gain the advantage over these menacing creatures by the advent of modern science. To gain an edge over these MDR microorganisms, we must consume our time and energy towards rapid diagnostic systems to equip us with the knowledge of what we are dealing with. Moreover, it is proved through literature that approaches that have nonprotein targets in bacteria are the most promising ones as the bacteria have so far not been able to cope up with it efficiently. To end this continuous war, we must also keep on isolating new drugs with novel modes of action.

Author Contributions

A.A.K. performed the literature search with K.N.M., M.S. and A.T. The manuscript was written by together by A.A.K., K.N.M., A.S., M.S., M.R., S.N. and H.-P.D. The tables were prepared by S.N., S.R. and A.T. The images were made by K.N.M., M.R., M.S. and A.T. The complete writing, design and corrections of the manuscript was led and supervised by A.A.K. and H.-P.D. All authors have read and agreed to the published version of the manuscript.

Funding

We are grateful for support within the project “InfektResonator” (IGF 20,934 N) awarded by the Bundesministerium für Wirtschaft und Energie (BMWi), and for support by the Bundesministerium für Bildung und Forschung, BMBF, (FlowArray project 13FH121PX8), Germany.

Institutional Review Board Statement

Not Applicable.

Informed Consent Statement

Not Applicable.

Data Availability Statement

Not Applicable.

Conflicts of Interest

The authors declare that there is no conflict of interest.

References

  1. Singh, R.; Smitha, M.; Singh, S.P. The role of nanotechnology in combating multi-drug resistant bacteria. J. Nanosci. Nanotechnol. 2014, 14, 4745–4756. [Google Scholar] [CrossRef]
  2. Alekshun, M.N.; Levy, S.B. Molecular mechanisms of antibacterial multidrug resistance. Cell 2007, 128, 1037–1050. [Google Scholar] [CrossRef] [Green Version]
  3. Kaur, I. Novel strategies to combat antimicrobial resistance. J. Infect. Dis. Ther. 2016, 4, 292. [Google Scholar] [CrossRef]
  4. Alanis, A.J. Resistance to antibiotics: Are we in the post-antibiotic era? Arch. Med. Res. 2005, 36, 697–705. [Google Scholar] [CrossRef] [PubMed]
  5. Munita, J.M.; Arias, C.A. Mechanisms of antibiotic resistance. Virulence Mech. Bact. Pathog. 2016, 481–511. [Google Scholar] [CrossRef] [Green Version]
  6. Azucena, E.; Mobashery, S. Aminoglycoside-modifying enzymes: Mechanisms of catalytic processes and inhibition. Drug Resist. Updates 2001, 4, 106–117. [Google Scholar] [CrossRef] [PubMed]
  7. Gill, E.E.; Franco, O.L.; Hancock, R.E. Antibiotic adjuvants: Diverse strategies for controlling drug-resistant pathogens. Chem. Biol. Drug Des. 2015, 85, 56–78. [Google Scholar] [CrossRef]
  8. Bhattacharjya, S.; Ramamoorthy, A. Multifunctional host defense peptides: Functional and mechanistic insights from NMR structures of potent antimicrobial peptides. FEBS J. 2009, 276, 6465–6473. [Google Scholar] [CrossRef] [Green Version]
  9. Bhunia, A.; Domadia, P.N.; Torres, J.; Hallock, K.J.; Ramamoorthy, A.; Bhattacharjya, S. NMR structure of pardaxin, a pore-forming antimicrobial peptide, in lipopolysaccharide micelles mechanism of outer membrane permeabilization. J. Biol. Chem. 2010, 285, 3883–3895. [Google Scholar] [CrossRef] [Green Version]
  10. Blanco, P.; Hernando-Amado, S.; Reales-Calderon, J.A.; Corona, F.; Lira, F.; Alcalde-Rico, M.; Bernardini, A.; Sanchez, M.B.; Martinez, J.L. Bacterial multidrug efflux pumps: Much more than antibiotic resistance determinants. Microorganisms 2016, 4, 14. [Google Scholar] [CrossRef] [Green Version]
  11. Džidić, S.; Šušković, J.; Kos, B. Antibiotic resistance mechanisms in bacteria: Biochemical and genetic aspects. Food Technol. Biotechnol. 2008, 46, 11–21. [Google Scholar]
  12. Khan, S.; Akhtar, M.U.; Khan, S.; Javed, F.; Khan, A.A. Nanoniosome-encapsulated levoflaxicin as an antibacterial agent against Brucella. J. Basic Microbiol. 2020, 60, 281–290. [Google Scholar] [CrossRef] [PubMed]
  13. Talib, A.; Manzoor, K.; Ijaz, A.; Adnan, F.; Javed, F.; Khan, A. Encapsulated virgin coconut oil as a nanoscale in vitro solution against multiple drug resistant Staphylococcus aureus. Micro Nano Lett. 2020. [Google Scholar] [CrossRef]
  14. Talib, A.; Manzoor, K.N.; Ali, W.; Saeed, M.; Gondal, M.A.; Badshah, M.; Khan, A.A. Biogenic Copper Nanoparticles as a Nanoscale Solution to Address Multiple Drug Resistance in Bacteria. Pak. J. Zool. 2021, 53, 201–208. [Google Scholar] [CrossRef]
  15. Rai, M.; Yadav, A.; Gade, A. Silver nanoparticles as a new generation of antimicrobials. Biotechnol. Adv. 2009, 27, 76–83. [Google Scholar] [CrossRef]
  16. Sondi, I.; Salopek-Sondi, B. Silver nanoparticles as antimicrobial agent: A case study on E. coli as a model for Gram-negative bacteria. J. Colloid Interface Sci. 2004, 275, 177–182. [Google Scholar] [CrossRef]
  17. Kim, J.S.; Kuk, E.; Yu, K.N.; Kim, J.-H.; Park, S.J.; Lee, H.J.; Kim, S.H.; Park, Y.K.; Park, Y.H.; Hwang, C.-Y. Antimicrobial effects of silver nanoparticles. Nanomed. Nanotechnol. Biol. Med. 2007, 3, 95–101. [Google Scholar] [CrossRef]
  18. Zharov, V.P.; Mercer, K.E.; Galitovskaya, E.N.; Smeltzer, M.S. Photothermal nanotherapeutics and nanodiagnostics for selective killing of bacteria targeted with gold nanoparticles. Biophys. J. 2006, 90, 619–627. [Google Scholar] [CrossRef] [Green Version]
  19. Kutateladze, M.; Adamia, R. Bacteriophages as potential new therapeutics to replace or supplement antibiotics. Trends Biotechnol. 2010, 28, 591–595. [Google Scholar] [CrossRef]
  20. Jamalifar, H.; Rahimi, H.; Samadi, N.; Shahverdi, A.; Sharifian, Z.; Hosseini, F.; Eslahi, H.; Fazeli, M. Antimicrobial activity of different Lactobacillus species against multi-drug resistant clinical isolates of Pseudomonas aeruginosa. Iran. J. Microbiol. 2011, 3, 21. [Google Scholar]
  21. Gibbs, P. Novel uses for lactic acid fermentation in food preservation. J. Appl. Bacteriol. 1987, 63, 51s–58s. [Google Scholar] [CrossRef]
  22. Rutherford, S.T.; Bassler, B.L. Bacterial quorum sensing: Its role in virulence and possibilities for its control. Cold Spring Harb. Perspect. Med. 2012, 2, a012427. [Google Scholar] [CrossRef] [PubMed]
  23. Joice, R.; Lipsitch, M. Targeting imperfect vaccines against drug-resistance determinants: A strategy for countering the rise of drug resistance. PLoS ONE 2013, 8, e68940. [Google Scholar] [CrossRef] [PubMed]
  24. Kennedy, D.A.; Read, A.F. Why does drug resistance readily evolve but vaccine resistance does not? Proc. R. Soc. B Biol. Sci. 2017, 284, 20162562. [Google Scholar] [CrossRef] [PubMed]
  25. Gupta, A.; Landis, R.F.; Rotello, V.M. Nanoparticle-based antimicrobials: Surface functionality is critical. F1000Research 2016, 5. F1000 Faculty Rev-364. [Google Scholar] [CrossRef]
  26. Jehan, M.; Saeed, F.; Khan, Z.; Shah, M.; Sikandar, A.; Inayat, A.; Ali, S.; Khan, A.M.; Talib, A.; Aasim, M. Investigative evaluation of Cassia absus for antibacterial capacity and biomimetic synthesis of silver nanoparticles. IET Nanobiotechnol. 2018, 12, 1062–1066. [Google Scholar] [CrossRef] [PubMed]
  27. Kumar, M.; Curtis, A.; Hoskins, C. Application of nanoparticle technologies in the combat against anti-microbial resistance. Pharmaceutics 2018, 10, 11. [Google Scholar] [CrossRef] [Green Version]
  28. Aruguete, D.M.; Kim, B.; Hochella, M.F.; Ma, Y.; Cheng, Y.; Hoegh, A.; Liu, J.; Pruden, A. Antimicrobial nanotechnology: Its potential for the effective management of microbial drug resistance and implications for research needs in microbial nanotoxicology. Environ. Sci. Process. Impacts 2013, 15, 93–102. [Google Scholar] [CrossRef]
  29. Kim, C.-B.; Yi, D.K.; Kim, P.S.S.; Lee, W.; Kim, M.J. Rapid photothermal lysis of the pathogenic bacteria, Escherichia coli using synthesis of gold nanorods. J. Nanosci. Nanotechnol. 2009, 9, 2841–2845. [Google Scholar] [CrossRef] [Green Version]
  30. Shrivastava, S.; Bera, T.; Roy, A.; Singh, G.; Ramachandrarao, P.; Dash, D. Characterization of enhanced antibacterial effects of novel silver nanoparticles. Nanotechnology 2007, 18, 225103. [Google Scholar] [CrossRef]
  31. Feng, Q.L.; Wu, J.; Chen, G.Q.; Cui, F.; Kim, T.; Kim, J. A mechanistic study of the antibacterial effect of silver ions on Escherichia coli and Staphylococcus aureus. J. Biomed. Mater. Res. 2000, 52, 662–668. [Google Scholar] [CrossRef]
  32. Cui, Y.; Zhao, Y.; Tian, Y.; Zhang, W.; Lü, X.; Jiang, X. The molecular mechanism of action of bactericidal gold nanoparticles on Escherichia coli. Biomaterials 2012, 33, 2327–2333. [Google Scholar] [CrossRef] [PubMed]
  33. Rudramurthy, G.R.; Swamy, M.K.; Sinniah, U.R.; Ghasemzadeh, A. Nanoparticles: Alternatives against drug-resistant pathogenic microbes. Molecules 2016, 21, 836. [Google Scholar] [CrossRef] [PubMed]
  34. Hayden, S.C.; Zhao, G.; Saha, K.; Phillips, R.L.; Li, X.; Miranda, O.R.; Rotello, V.M.; El-Sayed, M.A.; Schmidt-Krey, I.; Bunz, U.H. Aggregation and interaction of cationic nanoparticles on bacterial surfaces. J. Am. Chem. Soc. 2012, 134, 6920–6923. [Google Scholar] [CrossRef] [PubMed]
  35. Li, X.; Robinson, S.M.; Gupta, A.; Saha, K.; Jiang, Z.; Moyano, D.F.; Sahar, A.; Riley, M.A.; Rotello, V.M. Functional gold nanoparticles as potent antimicrobial agents against multi-drug-resistant bacteria. ACS Nano 2014, 8, 10682–10686. [Google Scholar] [CrossRef]
  36. Zhao, Y.; Tian, Y.; Cui, Y.; Liu, W.; Ma, W.; Jiang, X. Small molecule-capped gold nanoparticles as potent antibacterial agents that target gram-negative bacteria. J. Am. Chem. Soc. 2010, 132, 12349–12356. [Google Scholar] [CrossRef]
  37. Obuya, E.A.; Joshi, P.C.; Gray, T.A.; Keane, T.C.; Jones, W.E., Jr. Application of Pt. TiO2 nanofibers in photosensitized degradation of Rhodamine B. Int. J. Chem. 2014, 6, 1–16. [Google Scholar]
  38. Raghunath, A.; Perumal, E. Metal oxide nanoparticles as antimicrobial agents: A promise for the future. Int. J. Antimicrob. Agents 2017, 49, 137–152. [Google Scholar] [CrossRef]
  39. Ren, G.; Hu, D.; Cheng, E.W.; Vargas-Reus, M.A.; Reip, P.; Allaker, R.P. Characterisation of copper oxide nanoparticles for antimicrobial applications. Int. J. Antimicrob. Agents 2009, 33, 587–590. [Google Scholar] [CrossRef]
  40. Maniprasad, P.; Santra, S. Novel copper (Cu) loaded core–shell silica nanoparticles with improved Cu bioavailability: Synthesis, characterization and study of antibacterial properties. J. Biomed. Nanotechnol. 2012, 8, 558–566. [Google Scholar] [CrossRef] [Green Version]
  41. Huang, Z.; Zheng, X.; Yan, D.; Yin, G.; Liao, X.; Kang, Y.; Yao, Y.; Huang, D.; Hao, B. Toxicological effect of ZnO nanoparticles based on bacteria. Langmuir 2008, 24, 4140–4144. [Google Scholar] [CrossRef]
  42. Król, A.; Pomastowski, P.; Rafińska, K.; Railean-Plugaru, V.; Buszewski, B. Zinc oxide nanoparticles: Synthesis, antiseptic activity and toxicity mechanism. Adv. Colloid Interface Sci. 2017, 249, 37–52. [Google Scholar] [CrossRef] [PubMed]
  43. Ansari, M.A.; Khan, H.M.; Khan, A.A.; Sultan, A.; Azam, A. Synthesis and characterization of the antibacterial potential of ZnO nanoparticles against extended-spectrum β-lactamases-producing Escherichia coli and Klebsiella pneumoniae isolated from a tertiary care hospital of North India. Appl. Microbiol. Biotechnol. 2012, 94, 467–477. [Google Scholar] [CrossRef] [PubMed]
  44. Salem, W.; Leitner, D.R.; Zingl, F.G.; Schratter, G.; Prassl, R.; Goessler, W.; Reidl, J.; Schild, S. Antibacterial activity of silver and zinc nanoparticles against Vibrio cholerae and enterotoxic Escherichia coli. Int. J. Med Microbiol. 2015, 305, 85–95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Xie, Y.; He, Y.; Irwin, P.L.; Jin, T.; Shi, X. Antibacterial activity and mechanism of action of zinc oxide nanoparticles against Campylobacter jejuni. Appl. Environ. Microbiol. 2011, 77, 2325–2331. [Google Scholar] [CrossRef] [Green Version]
  46. Reddy, L.S.; Nisha, M.M.; Joice, M.; Shilpa, P. Antimicrobial activity of zinc oxide (ZnO) nanoparticle against Klebsiella pneumoniae. Pharm. Biol. 2014, 52, 1388–1397. [Google Scholar] [CrossRef]
  47. Jones, M.L.; Ganopolsky, J.G.; Labbé, A.; Wahl, C.; Prakash, S. Antimicrobial properties of nitric oxide and its application in antimicrobial formulations and medical devices. Appl. Microbiol. Biotechnol. 2010, 88, 401–407. [Google Scholar] [CrossRef]
  48. Schairer, D.O.; Chouake, J.S.; Nosanchuk, J.D.; Friedman, A.J. The potential of nitric oxide releasing therapies as antimicrobial agents. Virulence 2012, 3, 271–279. [Google Scholar] [CrossRef] [Green Version]
  49. Rubbo, H.; Radi, R.; Trujillo, M.; Telleri, R.; Kalyanaraman, B.; Barnes, S.; Kirk, M.; Freeman, B.A. Nitric oxide regulation of superoxide and peroxynitrite-dependent lipid peroxidation. Formation of novel nitrogen-containing oxidized lipid derivatives. J. Biol. Chem. 1994, 269, 26066–26075. [Google Scholar] [CrossRef]
  50. Deupree, S.M.; Schoenfisch, M.H. Morphological analysis of the antimicrobial action of nitric oxide on Gram-negative pathogens using atomic force microscopy. Acta Biomater. 2009, 5, 1405–1415. [Google Scholar] [CrossRef] [Green Version]
  51. Drapier, J.; Pellat, C.; Henry, Y. Generation of EPR-detectable nitrosyl-iron complexes in tumor target cells cocultured with activated macrophages. J. Biol. Chem. 1991, 266, 10162–10167. [Google Scholar] [CrossRef]
  52. Ischiropoulos, H.; Al-Mehdi, A.B. Peroxynitrite-mediated oxidative protein modifications. FEBS Lett. 1995, 364, 279–282. [Google Scholar] [CrossRef] [Green Version]
  53. Laval, F.; Wink, D.; Laval, J. A discussion of mechanisms of NO genotoxicty: Implication of inhibition of DNA repair proteins. In Reviews of Physiology Biochemistry and Pharmacology; Springer: Berlin/Heidelberg, Germany, 1997; Volume 131, pp. 175–191. [Google Scholar]
  54. Laval, F.; Wink, D.A. Inhibition by nitric oxide of the repair protein, O 6-DNA-methyltransferase. Carcinogenesis 1994, 15, 443–447. [Google Scholar] [CrossRef] [PubMed]
  55. Rai, M.K.; Deshmukh, S.; Ingle, A.; Gade, A. Silver nanoparticles: The powerful nanoweapon against multidrug-resistant bacteria. J. Appl. Microbiol. 2012, 112, 841–852. [Google Scholar] [CrossRef]
  56. Friedman, A.; Blecher, K.; Sanchez, D.; Tuckman-Vernon, C.; Gialanella, P.; Friedman, J.M.; Martinez, L.R.; Nosanchuk, J.D. Susceptibility of Gram-positive and-negative bacteria to novel nitric oxide-releasing nanoparticle technology. Virulence 2011, 2, 217–221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Reighard, K.P.; Schoenfisch, M.H. Antibacterial action of nitric oxide-releasing chitosan oligosaccharides against Pseudomonas aeruginosa under aerobic and anaerobic conditions. Antimicrob. Agents Chemother. 2015, 59, 6506–6513. [Google Scholar] [CrossRef] [Green Version]
  58. Chen, M.-C.; Mi, F.-L.; Liao, Z.-X.; Hsiao, C.-W.; Sonaje, K.; Chung, M.-F.; Hsu, L.-W.; Sung, H.-W. Recent advances in chitosan-based nanoparticles for oral delivery of macromolecules. Adv. Drug Deliv. Rev. 2013, 65, 865–879. [Google Scholar] [CrossRef]
  59. Ghadi, A.; Mahjoub, S.; Tabandeh, F.; Talebnia, F. Synthesis and optimization of chitosan nanoparticles: Potential applications in nanomedicine and biomedical engineering. Casp. J. Intern. Med. 2014, 5, 156. [Google Scholar]
  60. Al-Bakri, A.G.; Mahmoud, N.N. Photothermal-induced antibacterial activity of gold nanorods loaded into polymeric hydrogel against Pseudomonas aeruginosa biofilm. Molecules 2019, 24, 2661. [Google Scholar] [CrossRef] [Green Version]
  61. Mahmoud, N.N.; Alkilany, A.M.; Khalil, E.A.; Al-Bakri, A.G. Nano-photothermal ablation effect of hydrophilic and hydrophobic functionalized gold nanorods on Staphylococcus aureus and Propionibacterium acnes. Sci. Rep. 2018, 8, 6881. [Google Scholar] [CrossRef] [Green Version]
  62. Pattani, V.P.; Tunnell, J.W. Nanoparticle-mediated photothermal therapy: A comparative study of heating for different particle types. Lasers Surg. Med. 2012, 44, 675–684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Pissuwan, D.; Cortie, C.H.; Valenzuela, S.M.; Cortie, M.B. Functionalised gold nanoparticles for controlling pathogenic bacteria. Trends Biotechnol. 2010, 28, 207–213. [Google Scholar] [CrossRef] [PubMed]
  64. Li, J.-L.; Gu, M. Gold-nanoparticle-enhanced cancer photothermal therapy. IEEE J. Sel. Top. Quantum Electron. 2009, 16, 989–996. [Google Scholar]
  65. Mocan, L.; Tabaran, F.A.; Mocan, T.; Pop, T.; Mosteanu, O.; Agoston-Coldea, L.; Matea, C.T.; Gonciar, D.; Zdrehus, C.; Iancu, C. Laser thermal ablation of multidrug-resistant bacteria using functionalized gold nanoparticles. Int. J. Nanomed. 2017, 12, 2255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Silvero, C.M.J.N.; Rocca, D.M.; de la Villarmois, E.A.; Fournier, K.; Lanterna, A.E.; Perez, M.F.; Becerra, M.C.; Scaiano, J.C. Selective photoinduced antibacterial activity of amoxicillin-coated gold nanoparticles: From one-step synthesis to in vivo cytocompatibility. ACS Omega 2018, 3, 1220–1230. [Google Scholar] [CrossRef]
  67. Huang, W.-C.; Tsai, P.-J.; Chen, Y.-C. Functional gold nanoparticles as photothermal agents for selective-killing of pathogenic bacteria. Nanomedicine 2007, 2. [Google Scholar] [CrossRef]
  68. Esmaeillou, M.; Zarrini, G.; Rezaee, M.A. Vancomycin capped with silver nanoparticles as an antibacterial agent against multi-drug resistance bacteria. Adv. Pharm. Bull. 2017, 7, 479. [Google Scholar] [CrossRef] [Green Version]
  69. Norman, R.S.; Stone, J.W.; Gole, A.; Murphy, C.J.; Sabo-Attwood, T.L. Targeted photothermal lysis of the pathogenic bacteria, Pseudomonas aeruginosa, with gold nanorods. Nano Lett. 2008, 8, 302–306. [Google Scholar] [CrossRef]
  70. Fan, Z.; Senapati, D.; Khan, S.A.; Singh, A.K.; Hamme, A.; Yust, B.; Sardar, D.; Ray, P.C. Popcorn-Shaped Magnetic Core-Plasmonic Shell Multifunctional Nanoparticles for the Targeted Magnetic Separation and Enrichment, Label-Free SERS Imaging, and Photothermal Destruction of Multidrug-Resistant Bacteria. Chem.-Eur. J. 2013, 19, 2839–2847. [Google Scholar] [CrossRef]
  71. Perni, S.; Piccirillo, C.; Pratten, J.; Prokopovich, P.; Chrzanowski, W.; Parkin, I.P.; Wilson, M. The antimicrobial properties of light-activated polymers containing methylene blue and gold nanoparticles. Biomaterials 2009, 30, 89–93. [Google Scholar] [CrossRef]
  72. Lima, E.; Guerra, R.; Lara, V.; Guzmán, A. Gold nanoparticles as efficient antimicrobial agents for Escherichia coli and Salmonella typhi. Chem. Cent. J. 2013, 7, 11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Ivask, A.; ElBadawy, A.; Kaweeteerawat, C.; Boren, D.; Fischer, H.; Ji, Z.; Chang, C.H.; Liu, R.; Tolaymat, T.; Telesca, D. Toxicity mechanisms in Escherichia coli vary for silver nanoparticles and differ from ionic silver. ACS Nano 2014, 8, 374–386. [Google Scholar] [CrossRef] [PubMed]
  74. Matsumura, Y.; Yoshikata, K.; Kunisaki, S.-I.; Tsuchido, T. Mode of bactericidal action of silver zeolite and its comparison with that of silver nitrate. Appl. Environ. Microbiol. 2003, 69, 4278–4281. [Google Scholar] [CrossRef] [Green Version]
  75. Ayala-Núñez, N.V.; 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]
  76. Nanda, A.; Saravanan, M. Biosynthesis of silver nanoparticles from Staphylococcus aureus and its antimicrobial activity against MRSA and MRSE. Nanomed. Nanotechnol. Biol. Med. 2009, 5, 452–456. [Google Scholar] [CrossRef]
  77. Lara, H.H.; Ayala-Núñez, N.V.; Turrent, L.D.; Padilla, C.R. Bactericidal effect of silver nanoparticles against multidrug-resistant bacteria. World J. Microbiol. Biotechnol. 2010, 26, 615–621. [Google Scholar] [CrossRef]
  78. Suresh, A.K.; Pelletier, D.A.; Doktycz, M.J. Relating nanomaterial properties and microbial toxicity. Nanoscale 2013, 5, 463–474. [Google Scholar] [CrossRef]
  79. Talib, A.; Khan, A.A.; Ahmed, H.; Jilani, G. The nano-magnetic dancing of bacteria hand-in-hand with oxygen. Braz. Arch. Biol. Technol. 2017, 60. [Google Scholar] [CrossRef]
  80. Talib, A.; Khan, Z.; Bokhari, H.; Hidayathula, S.; Jilani, G.; Khan, A.A. Respiring cellular nano-magnets. Mater. Sci. Eng. C 2017, 80, 526–531. [Google Scholar] [CrossRef]
  81. Vigneshwaran, N.; Kathe, A.A.; Varadarajan, P.; Nachane, R.P.; Balasubramanya, R. Biomimetics of silver nanoparticles by white rot fungus, Phaenerochaete chrysosporium. Colloids Surf. B Biointerfaces 2006, 53, 55–59. [Google Scholar] [CrossRef]
  82. Durán, N.; Marcato, P.D.; Alves, O.L.; De Souza, G.I.; Esposito, E. Mechanistic aspects of biosynthesis of silver nanoparticles by several Fusarium oxysporum strains. J. Nanobiotechnol. 2005, 3, 8. [Google Scholar]
  83. Antony, J.J.; Sivalingam, P.; Siva, D.; Kamalakkannan, S.; Anbarasu, K.; Sukirtha, R.; Krishnan, M.; Achiraman, S. Comparative evaluation of antibacterial activity of silver nanoparticles synthesized using Rhizophora apiculata and glucose. Colloids Surf. B Biointerfaces 2011, 88, 134–140. [Google Scholar] [CrossRef] [PubMed]
  84. Gunalan, S.; Sivaraj, R.; Rajendran, V. Green synthesized ZnO nanoparticles against bacterial and fungal pathogens. Prog. Nat. Sci. Mater. Int. 2012, 22, 693–700. [Google Scholar] [CrossRef] [Green Version]
  85. Martınez-Flores, E.; Negrete, J.; Villasenor, G.T. Structure and properties of Zn–Al–Cu alloy reinforced with alumina particles. Mater. Des. 2003, 24, 281–286. [Google Scholar] [CrossRef]
  86. Ravishankar Rai, V. Nanoparticles and Their Potential Application as Antimicrobials. 2011. Available online: http://citeseerx.ist.psu.edu/viewdoc/summary?doi=10.1.1.459.8922 (accessed on 21 June 2020).
  87. Ghorbani, H.R. A review of methods for synthesis of Al nanoparticles. Orient. J. Chem. 2014, 30, 1941–1949. [Google Scholar] [CrossRef] [Green Version]
  88. Sadiq, I.M.; Chowdhury, B.; Chandrasekaran, N.; Mukherjee, A. Antimicrobial sensitivity of Escherichia coli to alumina nanoparticles. Nanomed. Nanotechnol. Biol. Med. 2009, 5, 282–286. [Google Scholar] [CrossRef]
  89. Balasubramanyam, A.; Sailaja, N.; Mahboob, M.; Rahman, M.; Hussain, S.M.; Grover, P. In vitro mutagenicity assessment of aluminium oxide nanomaterials using the Salmonella/microsome assay. Toxicol. In Vitro 2010, 24, 1871–1876. [Google Scholar] [CrossRef]
  90. Manyasree, D.; Kiranmayi, P.; Kumar, R. Synthesis, characterization and antibacterial activity of aluminium oxide nanoparticles. Int. J. Pharm. Pharm. Sci. 2018, 10, 32–35. [Google Scholar]
  91. Chen, C.-W.; Hsu, C.-Y.; Lai, S.-M.; Syu, W.-J.; Wang, T.-Y.; Lai, P.-S. Metal nanobullets for multidrug resistant bacteria and biofilms. Adv. Drug Deliv. Rev. 2014, 78, 88–104. [Google Scholar] [CrossRef]
  92. Adams, L.K.; Lyon, D.Y.; Alvarez, P.J. Comparative eco-toxicity of nanoscale TiO2, SiO2, and ZnO water suspensions. Water Res. 2006, 40, 3527–3532. [Google Scholar] [CrossRef]
  93. Jiang, W.; Mashayekhi, H.; Xing, B. Bacterial toxicity comparison between nano-and micro-scaled oxide particles. Environ. Pollut. 2009, 157, 1619–1625. [Google Scholar] [CrossRef] [PubMed]
  94. Vale, N.; Correia, A.; Silva, S.; Figueiredo, P.; Mäkilä, E.; Salonen, J.; Hirvonen, J.; Pedrosa, J.; Santos, H.A.; Fraga, A. Preparation and biological evaluation of ethionamide-mesoporous silicon nanoparticles against Mycobacterium tuberculosis. Bioorg. Med. Chem. Lett. 2017, 27, 403–405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Polkovnikova, Y.A.; Lenshin, A.; Seredin, P.; Minakov, D. Porous silicon nanoparticles containing neurotropic drugs. Inorg. Mater. 2017, 53, 477–483. [Google Scholar] [CrossRef]
  96. Choi, S.-R.; Britigan, B.E.; Moran, D.M.; Narayanasamy, P. Gallium nanoparticles facilitate phagosome maturation and inhibit growth of virulent Mycobacterium tuberculosis in macrophages. PLoS ONE 2017, 12, e0177987. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Choi, S.-R.; Britigan, B.E.; Narayanasamy, P. Treatment of virulent Mycobacterium tuberculosis and HIV coinfected macrophages with gallium nanoparticles inhibits pathogen growth and modulates macrophage cytokine production. Msphere 2019, 4, e00443-19. [Google Scholar] [CrossRef] [Green Version]
  98. Li, W.; Cao, Z.; Liu, R.; Liu, L.; Li, H.; Li, X.; Chen, Y.; Lu, C.; Liu, Y. AuNPs as an important inorganic nanoparticle applied in drug carrier systems. Artif. Cells Nanomed. Biotechnol. 2019, 47, 4222–4233. [Google Scholar] [CrossRef] [Green Version]
  99. Zhao, X.; Drlica, K. Reactive oxygen species and the bacterial response to lethal stress. Curr. Opin. Microbiol. 2014, 21, 1–6. [Google Scholar] [CrossRef] [Green Version]
  100. Nijnik, A.; Hancock, R. Host defence peptides: Antimicrobial and immunomodulatory activity and potential applications for tackling antibiotic-resistant infections. Emerg. Health Threat. J. 2009, 2, 7078. [Google Scholar] [CrossRef] [Green Version]
  101. Vineeth Kumar, T.P.V.K.; Asha, R.; Shyla, G.; George, S. Identification and characterization of novel host defense peptides from the skin secretion of the fungoid frog, Hydrophylax bahuvistara (Anura: Ranidae). Chem. Biol. Drug Des. 2018, 92, 1409–1418. [Google Scholar] [CrossRef]
  102. Ganz, T. Defensins: Antimicrobial peptides of innate immunity. Nat. Rev. Immunol. 2003, 3, 710–720. [Google Scholar] [CrossRef]
  103. Agerberth, B.; Charo, J.; Werr, J.; Olsson, B.; Idali, F.; Lindbom, L.; Kiessling, R.; Jörnvall, H.; Wigzell, H.; Gudmundsson, G.H. The human antimicrobial and chemotactic peptides LL-37 and α-defensins are expressed by specific lymphocyte and monocyte populations. Blood J. Am. Soc. Hematol. 2000, 96, 3086–3093. [Google Scholar]
  104. Wehkamp, J.; Schmid, M.; Stange, E.F. Defensins and other antimicrobial peptides in inflammatory bowel disease. Curr. Opin. Gastroenterol. 2007, 23, 370–378. [Google Scholar] [CrossRef] [PubMed]
  105. Wu, J.; Liu, S.; Wang, H. Invasive fungi-derived defensins kill drug-resistant bacterial pathogens. Peptides 2018, 99, 82–91. [Google Scholar] [CrossRef] [PubMed]
  106. Zanetti, M. Cathelicidins, multifunctional peptides of the innate immunity. J. Leukoc. Biol. 2004, 75, 39–48. [Google Scholar] [CrossRef] [PubMed]
  107. Bowdish, D.M.; Davidson, D.J.; Lau, Y.E.; Lee, K.; Scott, M.G.; Hancock, R.E. Impact of LL-37 on anti-infective immunity. J. Leukoc. Biol. 2005, 77, 451–459. [Google Scholar] [CrossRef]
  108. Cai, S.; Qiao, X.; Feng, L.; Shi, N.; Wang, H.; Yang, H.; Guo, Z.; Wang, M.; Chen, Y.; Wang, Y. Python cathelicidin CATHPb1 protects against multidrug-resistant staphylococcal infections by antimicrobial-immunomodulatory duality. J. Med. Chem. 2018, 61, 2075–2086. [Google Scholar] [CrossRef]
  109. Wimley, W.C.; Hristova, K. Antimicrobial peptides: Successes, challenges and unanswered questions. J. Membr. Biol. 2011, 239, 27–34. [Google Scholar] [CrossRef] [Green Version]
  110. Liu, F.; Soh Yan Ni, A.; Lim, Y.; Mohanram, H.; Bhattacharjya, S.; Xing, B. Lipopolysaccharide neutralizing peptide–porphyrin conjugates for effective photoinactivation and intracellular imaging of Gram-negative bacteria strains. Bioconjug. Chem. 2012, 23, 1639–1647. [Google Scholar] [CrossRef]
  111. Jefferson, K.K. What drives bacteria to produce a biofilm? FEMS Microbiol. Lett. 2004, 236, 163–173. [Google Scholar] [CrossRef]
  112. Costerton, J.W.; Stewart, P.S.; Greenberg, E.P. Bacterial biofilms: A common cause of persistent infections. Science 1999, 284, 1318–1322. [Google Scholar] [CrossRef] [Green Version]
  113. Hancock, R.E.; Sahl, H.-G. Antimicrobial and host-defense peptides as new anti-infective therapeutic strategies. Nat. Biotechnol. 2006, 24, 1551–1557. [Google Scholar] [CrossRef] [PubMed]
  114. Rollema, H.S.; Kuipers, O.P.; Both, P.; De Vos, W.M.; Siezen, R.J. Improvement of solubility and stability of the antimicrobial peptide nisin by protein engineering. Appl. Environ. Microbiol. 1995, 61, 2873–2878. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Pepperney, A.; Chikindas, M.L. Antibacterial peptides: Opportunities for the prevention and treatment of dental caries. Probiot. Antimicrob. Proteins 2011, 3, 68. [Google Scholar] [CrossRef] [PubMed]
  116. Yoshida, T.; Nagasawa, T. ε-Poly-L-lysine: Microbial production, biodegradation and application potential. Appl. Microbiol. Biotechnol. 2003, 62, 21–26. [Google Scholar] [CrossRef] [PubMed]
  117. Melo, M.N.; Dugourd, D.; Castanho, M.A. Omiganan pentahydrochloride in the front line of clinical applications of antimicrobial peptides. Recent Pat. Anti-Infect. Drug Discov. 2006, 1, 201–207. [Google Scholar] [CrossRef] [PubMed]
  118. Kazemzadeh-Narbat, M.; Kindrachuk, J.; Duan, K.; Jenssen, H.; Hancock, R.E.; Wang, R. Antimicrobial peptides on calcium phosphate-coated titanium for the prevention of implant-associated infections. Biomaterials 2010, 31, 9519–9526. [Google Scholar] [CrossRef] [PubMed]
  119. Aka, S.T. Killing efficacy and anti-biofilm activity of synthetic human cationic antimicrobial peptide cathelicidin hCAP-18/LL37 against urinary tract pathogens. J. Microbiol. Infect. Dis. 2015, 5, 15–20. [Google Scholar] [CrossRef] [Green Version]
  120. Anunthawan, T.; De La Fuente-Núñez, C.; Hancock, R.E.; Klaynongsruang, S. Cationic amphipathic peptides KT2 and RT2 are taken up into bacterial cells and kill planktonic and biofilm bacteria. Biochim. Biophys. Acta (BBA)-Biomembr. 2015, 1848, 1352–1358. [Google Scholar] [CrossRef] [Green Version]
  121. Pletzer, D.; Coleman, S.R.; Hancock, R.E. Anti-biofilm peptides as a new weapon in antimicrobial warfare. Curr. Opin. Microbiol. 2016, 33, 35–40. [Google Scholar] [CrossRef] [Green Version]
  122. Ribeiro, S.M.; De La Fuente-Núñez, C.; Baquir, B.; Faria-Junior, C.; Franco, O.L.; Hancock, R.E. Antibiofilm peptides increase the susceptibility of carbapenemase-producing Klebsiella pneumoniae clinical isolates to β-lactam antibiotics. Antimicrob. Agents Chemother. 2015, 59, 3906–3912. [Google Scholar] [CrossRef] [Green Version]
  123. Jorge, P.; Lourenco, A.; Pereira, M.O. New trends in peptide-based anti-biofilm strategies: A review of recent achievements and bioinformatic approaches. Biofouling 2012, 28, 1033–1061. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Rapaport, D.; Shai, Y. Interaction of fluorescently labeled pardaxin and its analogues with lipid bilayers. J. Biol. Chem. 1991, 266, 23769–23775. [Google Scholar] [CrossRef]
  125. Ludtke, S.J.; He, K.; Heller, W.T.; Harroun, T.A.; Yang, L.; Huang, H.W. Membrane pores induced by magainin. Biochemistry 1996, 35, 13723–13728. [Google Scholar] [CrossRef] [PubMed]
  126. Wang, Z.; Shen, Y.; Haapasalo, M. Antibiofilm peptides against oral biofilms. J. Oral Microbiol. 2017, 9, 1327308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Gazit, E.; Miller, I.R.; Biggin, P.C.; Sansom, M.S.; Shai, Y. Structure and orientation of the mammalian antibacterial peptide cecropin P1 within phospholipid membranes. J. Mol. Biol. 1996, 258, 860–870. [Google Scholar] [CrossRef]
  128. Chung, P.Y.; Khanum, R. Antimicrobial peptides as potential anti-biofilm agents against multidrug-resistant bacteria. J. Microbiol. Immunol. Infect. 2017, 50, 405–410. [Google Scholar] [CrossRef]
  129. das Neves, R.C.; Mortari, M.R.; Schwartz, E.F.; Kipnis, A.; Junqueira-Kipnis, A.P. Antimicrobial and antibiofilm effects of peptides from venom of social Wasp and scorpion on multidrug-resistant Acinetobacter baumannii. Toxins 2019, 11, 216. [Google Scholar] [CrossRef] [Green Version]
  130. Bessa, L.J.; Eaton, P.; Dematei, A.; Plácido, A.; Vale, N.; Gomes, P.; Delerue-Matos, C.; Sa Leite, J.R.; Gameiro, P. Synergistic and antibiofilm properties of ocellatin peptides against multidrug-resistant Pseudomonas aeruginosa. Future Microbiol. 2018, 13, 151–163. [Google Scholar] [CrossRef] [Green Version]
  131. Gunn, J.S.; Bakaletz, L.O.; Wozniak, D.J. What’s on the outside matters: The role of the extracellular polymeric substance of gram-negative biofilms in evading host immunity and as a target for therapeutic intervention. J. Biol. Chem. 2016, 291, 12538–12546. [Google Scholar] [CrossRef] [Green Version]
  132. Otto, M. Bacterial sensing of antimicrobial peptides. In Bacterial Sensing and Signaling; Karger Publishers: Basel, Switzerland, 2009; Volume 16, pp. 136–149. [Google Scholar]
  133. Batoni, G.; Maisetta, G.; Lisa Brancatisano, F.; Esin, S.; Campa, M. Use of antimicrobial peptides against microbial biofilms: Advantages and limits. Curr. Med. Chem. 2011, 18, 256–279. [Google Scholar] [CrossRef]
  134. Gooderham, W.J.; Bains, M.; McPhee, J.B.; Wiegand, I.; Hancock, R.E. Induction by cationic antimicrobial peptides and involvement in intrinsic polymyxin and antimicrobial peptide resistance, biofilm formation, and swarming motility of PsrA in Pseudomonas aeruginosa. J. Bacteriol. 2008, 190, 5624–5634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Chung, P.Y.; Toh, Y.S. Anti-biofilm agents: Recent breakthrough against multi-drug resistant Staphylococcus aureus. Pathog. Dis. 2014, 70, 231–239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Kong, K.-F.; Vuong, C.; Otto, M. Staphylococcus quorum sensing in biofilm formation and infection. Int. J. Med. Microbiol. 2006, 296, 133–139. [Google Scholar] [CrossRef]
  137. Weinbauer, M.G. Ecology of prokaryotic viruses. FEMS Microbiol. Rev. 2004, 28, 127–181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Karthik, K.; Muneeswaran, N.S.; Manjunathachar, H.V.; Gopi, M.; Elamurugan, A.; Kalaiyarasu, S. Bacteriophages: Effective alternative to antibiotics. Adv. Anim. Vet. Sci. 2014, 2, 1–7. [Google Scholar] [CrossRef] [Green Version]
  139. Salmond, G.P.; Fineran, P.C. A century of the phage: Past, present and future. Nat. Rev. Microbiol. 2015, 13, 777–786. [Google Scholar] [CrossRef]
  140. Tiwari, R.; Dhama, K.; Kumar, A.; Rahal, A.; Kapoor, S. Bacteriophage therapy for safeguarding animal and human health: A review. Pak. J. Biol. Sci. PJBS 2014, 17, 301–315. [Google Scholar] [CrossRef] [Green Version]
  141. Dhama, K.; Chakraborty, S.; Wani, M.Y.; Verma, A.K.; Deb, R.; Tiwari, R.; Kapoor, S. Novel and emerging therapies safeguarding health of humans and their companion animals: A review. Pak. J. Biol. Sci. PJBS 2013, 16, 101. [Google Scholar] [CrossRef]
  142. Westwater, C.; Kasman, L.M.; Schofield, D.A.; Werner, P.A.; Dolan, J.W.; Schmidt, M.G.; Norris, J.S. Use of genetically engineered phage to deliver antimicrobial agents to bacteria: An alternative therapy for treatment of bacterial infections. Antimicrob. Agents Chemother. 2003, 47, 1301–1307. [Google Scholar] [CrossRef] [Green Version]
  143. Carvalho, C.M.; Gannon, B.W.; Halfhide, D.E.; Santos, S.B.; Hayes, C.M.; Roe, J.M.; Azeredo, J. The in vivo efficacy of two administration routes of a phage cocktail to reduce numbers of Campylobacter coli and Campylobacter jejuni in chickens. BMC Microbiol. 2010, 10, 232. [Google Scholar] [CrossRef] [Green Version]
  144. Tiwari, R.; Dhama, K.; Wani, M.Y.; Verma, V.; Vaid, R.; Chauhan, R. Bacteriophage therapy: A novel tool for combating bacterial diseases of poultry—A review. J. Immunol. Immunopathol. 2011, 13, 55–66. [Google Scholar]
  145. Mattey, M.; Spencer, J. Bacteriophage therapy—cooked goose or Phoenix rising? Curr. Opin. Biotechnol. 2008, 19, 608–612. [Google Scholar] [CrossRef] [PubMed]
  146. Sulakvelidze, A.; Alavidze, Z.; Morris, J.G. Bacteriophage therapy. Antimicrob. Agents Chemother. 2001, 45, 649–659. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Salvatore, P.P.; Becerra, M.C.; Abel zur Wiesch, P.; Hinkley, T.; Kaur, D.; Sloutsky, A.; Cohen, T. Fitness costs of drug resistance mutations in multidrug-resistant Mycobacterium tuberculosis: A household-based case-control study. J. Infect. Dis. 2016, 213, 149–155. [Google Scholar] [CrossRef] [Green Version]
  148. Ramesh, V.; Fralick, J.A.; Rolfe, R.D. Prevention of Clostridium difficile-induced ileocecitis with bacteriophage. Anaerobe 1999, 5, 69–78. [Google Scholar] [CrossRef]
  149. Nale, J.Y.; Spencer, J.; Hargreaves, K.R.; Buckley, A.M.; Trzepiński, P.; Douce, G.R.; Clokie, M.R. Bacteriophage combinations significantly reduce Clostridium difficile growth in vitro and proliferation in vivo. Antimicrob. Agents Chemother. 2016, 60, 968–981. [Google Scholar] [CrossRef] [Green Version]
  150. Biswas, B.; Adhya, S.; Washart, P.; Paul, B.; Trostel, A.N.; Powell, B.; Carlton, R.; Merril, C.R. Bacteriophage therapy rescues mice bacteremic from a clinical isolate of vancomycin-resistant Enterococcus faecium. Infect. Immun. 2002, 70, 204–210. [Google Scholar] [CrossRef] [Green Version]
  151. Wang, J.; Hu, B.; Xu, M.; Yan, Q.; Liu, S.; Zhu, X.; Sun, Z.; Tao, D.; Ding, L.; Reed, E. Therapeutic effectiveness of bacteriophages in the rescue of mice with extended spectrum β-lactamase-producing Escherichia coli bacteremia. Int. J. Mol. Med. 2006, 17, 347–355. [Google Scholar] [CrossRef]
  152. Wang, J.; Hu, B.; Xu, M.; Yan, Q.; Liu, S.; Zhu, X.; Sun, Z.; Reed, E.; Ding, L.; Gong, J. Use of bacteriophage in the treatment of experimental animal bacteremia from imipenem-resistant Pseudomonas aeruginosa. Int. J. Mol. Med. 2006, 17, 309–317. [Google Scholar] [CrossRef] [Green Version]
  153. Watanabe, R.; Matsumoto, T.; Sano, G.; Ishii, Y.; Tateda, K.; Sumiyama, Y.; Uchiyama, J.; Sakurai, S.; Matsuzaki, S.; Imai, S. Efficacy of bacteriophage therapy against gut-derived sepsis caused by Pseudomonas aeruginosa in mice. Antimicrob. Agents Chemother. 2007, 51, 446–452. [Google Scholar] [CrossRef] [Green Version]
  154. Chan, B.K.; Sistrom, M.; Wertz, J.E.; Kortright, K.E.; Narayan, D.; Turner, P.E. Phage selection restores antibiotic sensitivity in MDR Pseudomonas aeruginosa. Sci. Rep. 2016, 6, 26717. [Google Scholar] [CrossRef]
  155. McCormick, A.W.; Whitney, C.G.; Farley, M.M.; Lynfield, R.; Harrison, L.H.; Bennett, N.M.; Schaffner, W.; Reingold, A.; Hadler, J.; Cieslak, P. Geographic diversity and temporal trends of antimicrobial resistance in Streptococcus pneumoniae in the United States. Nat. Med. 2003, 9, 424–430. [Google Scholar] [CrossRef] [PubMed]
  156. Control, C.F.D. Prevention, Emergence of Mycobacterium tuberculosis with extensive resistance to second-line drugs—Worldwide, 2000–2004. MMWR Morb. Mortal. Wkly. Rep. 2006, 55, 301. [Google Scholar]
  157. Alisky, J.; Iczkowski, K.; Rapoport, A.; Troitsky, N. Bacteriophages show promise as antimicrobial agents. J. Infect. 1998, 36, 5–15. [Google Scholar] [CrossRef]
  158. Thiel, K. Old dogma, new tricks—21st century phage therapy. Nat. Biotechnol. 2004, 22, 31–36. [Google Scholar] [CrossRef]
  159. Pirnay, J.-P.; De Vos, D.; Verbeken, G.; Merabishvili, M.; Chanishvili, N.; Vaneechoutte, M.; Zizi, M.; Laire, G.; Lavigne, R.; Huys, I. The phage therapy paradigm: Pret-a-porter or sur-mesure? Pharm. Res. 2011, 28, 934–937. [Google Scholar] [CrossRef]
  160. Brockhurst, M.A.; Morgan, A.D.; Fenton, A.; Buckling, A. Experimental coevolution with bacteria and phage: The Pseudomonas fluorescens—Φ2 model system. Infect. Genet. Evol. 2007, 7, 547–552. [Google Scholar] [CrossRef]
  161. Cha, K.; Oh, H.K.; Jang, J.Y.; Jo, Y.; Kim, W.K.; Ha, G.U.; Ko, K.S.; Myung, H. Characterization of two novel bacteriophages infecting multidrug-resistant (MDR) Acinetobacter baumannii and evaluation of their therapeutic efficacy in vivo. Front. Microbiol. 2018, 9, 696. [Google Scholar] [CrossRef] [Green Version]
  162. LaVergne, S.; Hamilton, T.; Biswas, B.; Kumaraswamy, M.; Schooley, R.; Wooten, D. Phage therapy for a multidrug-resistant Acinetobacter baumannii craniectomy site infection. In Open Forum Infectious Diseases; Oxford University Press US: New York, NY, USA, 2018; p. ofy064. [Google Scholar]
  163. Lood, R.; Winer, B.Y.; Pelzek, A.J.; Diez-Martinez, R.; Thandar, M.; Euler, C.W.; Schuch, R.; Fischetti, V.A. Novel phage lysin capable of killing the multidrug-resistant gram-negative bacterium Acinetobacter baumannii in a mouse bacteremia model. Antimicrob. Agents Chemother. 2015, 59, 1983–1991. [Google Scholar] [CrossRef] [Green Version]
  164. Witzenrath, M.; Schmeck, B.; Doehn, J.M.; Tschernig, T.; Zahlten, J.; Loeffler, J.M.; Zemlin, M.; Müller, H.; Gutbier, B.; Schütte, H. Systemic use of the endolysin Cpl-1 rescues mice with fatal pneumococcal pneumonia. Crit. Care Med. 2009, 37, 642–649. [Google Scholar] [CrossRef]
  165. Schmelcher, M.; Shen, Y.; Nelson, D.C.; Eugster, M.R.; Eichenseher, F.; Hanke, D.C.; Loessner, M.J.; Dong, S.; Pritchard, D.G.; Lee, J.C. Evolutionarily distinct bacteriophage endolysins featuring conserved peptidoglycan cleavage sites protect mice from MRSA infection. J. Antimicrob. Chemother. 2015, 70, 1453–1465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Wang, Q.; Euler, C.W.; Delaune, A.; Fischetti, V.A. Using a novel lysin to help control Clostridium difficile infections. Antimicrob. Agents Chemother. 2015, 59, 7447–7457. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Gilmer, D.B.; Schmitz, J.E.; Euler, C.W.; Fischetti, V.A. Novel bacteriophage lysin with broad lytic activity protects against mixed infection by Streptococcus pyogenes and methicillin-resistant Staphylococcus aureus. Antimicrob. Agents Chemother. 2013, 57, 2743–2750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Shen, Y.; Barros, M.; Vennemann, T.; Gallagher, D.T.; Yin, Y.; Linden, S.B.; Heselpoth, R.D.; Spencer, D.J.; Donovan, D.M.; Moult, J. A bacteriophage endolysin that eliminates intracellular streptococci. eLife 2016, 5, e13152. [Google Scholar] [CrossRef] [PubMed]
  169. Moghadam, M.T.; Khoshbayan, A.; Chegini, Z.; Farahani, I.; Shariati, A. Bacteriophages, a New Therapeutic Solution for Inhibiting Multidrug-Resistant Bacteria Causing Wound Infection: Lesson from Animal Models and Clinical Trials. Drug Des. Dev. Ther. 2020, 14, 1867. [Google Scholar] [CrossRef] [PubMed]
  170. Yang, H.; Wang, D.-B.; Dong, Q.; Zhang, Z.; Cui, Z.; Deng, J.; Yu, J.; Zhang, X.-E.; Wei, H. Existence of separate domains in lysin PlyG for recognizing Bacillus anthracis spores and vegetative cells. Antimicrob. Agents Chemother. 2012, 56, 5031–5039. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Krieg, A.M. CpG motifs: The active ingredient in bacterial extracts? Nat. Med. 2003, 9, 831–835. [Google Scholar] [CrossRef]
  172. Braido, F.; Tarantini, F.; Ghiglione, V.; Melioli, G.; Canonica, G. Bacterial lysate in the prevention of acute exacerbation of COPD and in respiratory recurrent infections. Int. J. Chronic Obstr. Pulm. Dis. 2007, 2, 335. [Google Scholar]
  173. Giovannini, M.; Salvini, F.; Riva, E. Bacterial extracts as immunomodulators for the prevention of recurrent respiratory infections in children. J. Med. Microb. Diagn 2014, 3, 136. [Google Scholar]
  174. Nile, R.S.; Darweesh, M.F.; Al-Rufaie, M.M. Liposomal-lipopolysaccharide vaccine extracted from Proteus mirabilis induces moderate TLR4 and CD14 production. Curr. Issues Pharm. Med. Sci. 2019, 32, 81–86. [Google Scholar] [CrossRef] [Green Version]
  175. Lipsitch, M.; Siber, G.R. How can vaccines contribute to solving the antimicrobial resistance problem? MBio 2016, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Chang, H.-H.; Cohen, T.; Grad, Y.H.; Hanage, W.P.; O’Brien, T.F.; Lipsitch, M. Origin and proliferation of multiple-drug resistance in bacterial pathogens. Microbiol. Mol. Biol. Rev. 2015, 79, 101–116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Fine, P.E. Herd immunity: History, theory, practice. Epidemiol. Rev. 1993, 15, 265–302. [Google Scholar] [CrossRef] [PubMed]
  178. Levin, B.R.; Cornejo, O.E. The population and evolutionary dynamics of homologous gene recombination in bacteria. PLoS Genet. 2009, 5, e1000601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Levin, B.R.; Stewart, F.M.; Rice, V.A. The kinetics of conjugative plasmid transmission: Fit of a simple mass action model. Plasmid 1979, 2, 247–260. [Google Scholar] [CrossRef]
  180. Chevereau, G.; Bollenbach, T. Systematic discovery of drug interaction mechanisms. Mol. Syst. Biol. 2015, 11, 807. [Google Scholar] [CrossRef] [PubMed]
  181. Chait, R.; Craney, A.; Kishony, R. Antibiotic interactions that select against resistance. Nature 2007, 446, 668–671. [Google Scholar] [CrossRef] [PubMed]
  182. Wood, K.B.; Wood, K.C.; Nishida, S.; Cluzel, P. Uncovering scaling laws to infer multidrug response of resistant microbes and cancer cells. Cell Rep. 2014, 6, 1073–1084. [Google Scholar] [CrossRef] [Green Version]
  183. Liu, Y.; Filler, S.G. Candida albicans Als3, a multifunctional adhesin and invasin. Eukaryot. Cell 2011, 10, 168–173. [Google Scholar] [CrossRef] [Green Version]
  184. Schmidt, C.S.; White, C.J.; Ibrahim, A.S.; Filler, S.G.; Fu, Y.; Yeaman, M.R.; Edwards, J.E., Jr.; Hennessey, J.P., Jr. NDV-3, a recombinant alum-adjuvanted vaccine for Candida and Staphylococcus aureus, is safe and immunogenic in healthy adults. Vaccine 2012, 30, 7594–7600. [Google Scholar] [CrossRef] [Green Version]
  185. Tekle, Y.I.; Nielsen, K.M.; Liu, J.; Pettigrew, M.M.; Meyers, L.A.; Galvani, A.P.; Townsend, J.P. Controlling antimicrobial resistance through targeted, vaccine-induced replacement of strains. PLoS ONE 2012, 7, e50688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Senna, J.P.; Roth, D.M.; Oliveira, J.S.; Machado, D.C.; Santos, D.S. Protective immune response against methicillin resistant Staphylococcus aureus in a murine model using a DNA vaccine approach. Vaccine 2003, 21, 2661–2666. [Google Scholar] [CrossRef]
  187. Lee, N.; Yuen, K.-Y.; Kumana, C.R. Clinical role of β-lactam/β-lactamase inhibitor combinations. Drugs 2003, 63, 1511–1524. [Google Scholar] [CrossRef] [PubMed]
  188. Zarantonelli, M.L.; Antignac, A.; Lancellotti, M.; Guiyoule, A.; Alonso, J.-M.; Taha, M.-K. Immunogenicity of meningococcal PBP2 during natural infection and protective activity of anti-PBP2 antibodies against meningococcal bacteraemia in mice. J. Antimicrob. Chemother. 2006, 57, 924–930. [Google Scholar] [CrossRef] [Green Version]
  189. Leuzzi, R.; Adamo, R.; Scarselli, M. Vaccines against Clostridium difficile. Hum. Vaccines Immunother. 2014, 10, 1466–1477. [Google Scholar] [CrossRef] [Green Version]
  190. Priebe, G.P.; Goldberg, J.B. Vaccines for Pseudomonas aeruginosa: A long and winding road. Expert Rev. Vaccines 2014, 13, 507–519. [Google Scholar] [CrossRef] [Green Version]
  191. Singh, V.; Bala, M.; Bhargava, A.; Kakran, M.; Bhatnagar, R. In vitro efficacy of 21 dual antimicrobial combinations comprising novel and currently recommended combinations for treatment of drug resistant gonorrhoea in future era. PLoS ONE 2018, 13, e0193678. [Google Scholar] [CrossRef]
  192. Ginsberg, A.M.; Spigelman, M. Challenges in tuberculosis drug research and development. Nat. Med. 2007, 13, 290–294. [Google Scholar] [CrossRef]
  193. Young, D.B.; Perkins, M.D.; Duncan, K.; Barry, C.E. Confronting the scientific obstacles to global control of tuberculosis. J. Clin. Investig. 2008, 118, 1255–1265. [Google Scholar] [CrossRef] [Green Version]
  194. Hansen, J.L.; Moore, P.B.; Steitz, T.A. Structures of five antibiotics bound at the peptidyl transferase center of the large ribosomal subunit. J. Mol. Biol. 2003, 330, 1061–1075. [Google Scholar] [CrossRef]
  195. Tu, D.; Blaha, G.; Moore, P.B.; Steitz, T.A. Structures of MLSBK antibiotics bound to mutated large ribosomal subunits provide a structural explanation for resistance. Cell 2005, 121, 257–270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Cocito, C. Antibiotics of the virginiamycin family, inhibitors which contain synergistic components. Microbiol. Rev. 1979, 43, 145. [Google Scholar] [CrossRef] [PubMed]
  197. Babb, R.; Pirofski, L.-A. Help is on the way: Monoclonal antibody therapy for multi-drug resistant bacteria. Virulence 2017, 8, 1055–1058. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Guachalla, L.M.; Ramoni, K.; Varga, C.; Mutti, M.; Ghazawi, A.; Pál, T.; Nagy, E.; Sonnevend, Á.; Nagy, G.; Szijártó, V. Retained activity of an O25b-specific monoclonal antibody against an Mcr-1-producing Escherichia coli sequence type 131 strain. Antimicrob. Agents Chemother. 2018, 62. [Google Scholar] [CrossRef] [Green Version]
  199. Thanabalasuriar, A.; Surewaard, B.G.; Willson, M.E.; Neupane, A.S.; Stover, C.K.; Warrener, P.; Wilson, G.; Keller, A.E.; Sellman, B.R.; DiGiandomenico, A. Bispecific antibody targets multiple Pseudomonas aeruginosa evasion mechanisms in the lung vasculature. J. Clin. Investig. 2017, 127, 2249–2261. [Google Scholar] [CrossRef]
  200. Motterlini, R.; Otterbein, L.E. The therapeutic potential of carbon monoxide. Nat. Rev. Drug Discov. 2010, 9, 728–743. [Google Scholar] [CrossRef]
  201. Romão, C.C.; Blättler, W.A.; Seixas, J.D.; Bernardes, G.J. Developing drug molecules for therapy with carbon monoxide. Chem. Soc. Rev. 2012, 41, 3571–3583. [Google Scholar] [CrossRef]
  202. García-Gallego, S.; Bernardes, G.J. Carbon-monoxide-releasing molecules for the delivery of therapeutic CO in vivo. Angew. Chem. Int. Ed. 2014, 53, 9712–9721. [Google Scholar] [CrossRef]
  203. Nobre, L.S.; Jeremias, H.; Romão, C.C.; Saraiva, L.M. Examining the antimicrobial activity and toxicity to animal cells of different types of CO-releasing molecules. Dalton Trans. 2016, 45, 1455–1466. [Google Scholar] [CrossRef]
  204. Ouwehand, A.C.; Forssten, S.; Hibberd, A.A.; Lyra, A.; Stahl, B. Probiotic approach to prevent antibiotic resistance. Ann. Med. 2016, 48, 246–255. [Google Scholar] [CrossRef]
  205. Halder, D.; Mandal, S. Insights into the antagonism of Lactobacillus fermentum curd isolate against Gram-positive and Gram-negative pathogenic bacteria. Biosci. Biotechnol. Res. Commun. 2018, 11, 461–468. [Google Scholar] [CrossRef]
  206. Nair, D.V.; Kollanoor Johny, A. Characterizing the antimicrobial function of a dairy-originated probiotic, Propionibacterium freudenreichii, against multidrug-resistant Salmonella enterica serovar Heidelberg in turkey poults. Front. Microbiol. 2018, 9, 1475. [Google Scholar] [CrossRef] [PubMed]
  207. Kiran, M.D.; Adikesavan, N.V.; Cirioni, O.; Giacometti, A.; Silvestri, C.; Scalise, G.; Ghiselli, R.; Saba, V.; Orlando, F.; Shoham, M. Discovery of a quorum-sensing inhibitor of drug-resistant staphylococcal infections by structure-based virtual screening. Mol. Pharmacol. 2008, 73, 1578–1586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Smith, D.; Wiegeshaus, E.; Navalkar, R.; Grover, A. Host-parasite relationships in experimental airborne tuberculosis I. Preliminary studies in BCG-vaccinated and nonvaccinated animals. J. Bacteriol. 1966, 91, 718–724. [Google Scholar] [CrossRef] [Green Version]
  209. Janeway, C.A.; Travers, P.; Walport, M.; Shlomchik, M. Immunobiology: The Immune System in Health and Disease. 2005; Garland Science: New York, NY, USA, 2017; Volume 6. [Google Scholar]
  210. Davis, M.M.; Bjorkman, P.J. T-cell antigen receptor genes and T-cell recognition. Nature 1988, 334, 395–402. [Google Scholar] [CrossRef]
  211. Alberts, B.; Johnson, A.; Lewis, J.; Raff, M.; Roberts, K.; Walter, P. B Cells Antibodies. In Molecular Biology of the Cell, 4th ed.; Garland Science: New York, NY, USA, 2002. [Google Scholar]
  212. Dhesi, Z.; Enne, V.I.; Justin, O.G.; Gant, V.; Livermore, D.M. Rapid and Point-of-Care Testing in Respiratory Tract Infections: An Antibiotic Guardian? ACS Pharmacol. Transl. Sci. 2020, 3, 401–417. [Google Scholar] [CrossRef]
  213. Mitsakakis, K.; Kaman, W.E.; Elshout, G.; Specht, M.; Hays, J.P. Challenges in identifying antibiotic resistance targets for point-of-care diagnostics in general practice. Future Microbiol. 2018, 13, 1157–1164. [Google Scholar] [CrossRef] [Green Version]
  214. Anderson, M.E.; O’Brien, E.C.; Grayek, E.N.; Hermansen, J.K.; Hunt, H.K. The detection of Helicobacter hepaticus using whispering-gallery mode microcavity optical sensors. Biosensors 2015, 5, 562–576. [Google Scholar] [CrossRef] [Green Version]
  215. Jiang, X.; Qavi, A.J.; Huang, S.H.; Yang, L. Whispering-Gallery Sensors. Matter 2020, 3, 371–392. [Google Scholar] [CrossRef]
  216. Jiang, X.; Qavi, A.J.; Huang, S.H.; Yang, L. Whispering gallery microsensors: A review. arXiv 2018, arXiv:1805.00062. [Google Scholar]
  217. Vahala, K.J. Optical microcavities. Nature 2003, 424, 839–846. [Google Scholar] [CrossRef] [PubMed]
  218. Foreman, M.R.; Swaim, J.D.; Vollmer, F. Whispering gallery mode sensors. Adv. Opt. Photonics 2015, 7, 168–240. [Google Scholar] [CrossRef] [PubMed]
  219. Cai, L.; Pan, J.; Zhao, Y.; Wang, J.; Xiao, S. Whispering Gallery Mode Optical Microresonators: Structures and Sensing Applications. Phys. Status Solidi (A) 2020, 217, 1900825. [Google Scholar] [CrossRef]
Figure 1. The possible ways a bacterium resists the action of an antibiotic drug.
Figure 1. The possible ways a bacterium resists the action of an antibiotic drug.
Ijms 22 00859 g001
Figure 2. A pictorial presentation of various strategies, targets and effector molecules that can be used to curb the multiple drug resistance in bacterial pathogens.
Figure 2. A pictorial presentation of various strategies, targets and effector molecules that can be used to curb the multiple drug resistance in bacterial pathogens.
Ijms 22 00859 g002
Figure 3. Different modes of action possible for the eradication (killing) of bacterial cells through silver (Ag) nanoparticles [1].
Figure 3. Different modes of action possible for the eradication (killing) of bacterial cells through silver (Ag) nanoparticles [1].
Ijms 22 00859 g003
Figure 4. Mode of action (bactericidal effects) of AgNPs against bacteria against MRSA.
Figure 4. Mode of action (bactericidal effects) of AgNPs against bacteria against MRSA.
Ijms 22 00859 g004
Figure 5. Promyelocytes in the bone marrow synthesize α-defensins. 94 amino acid preprodefensin (purple) is biosynthesized in the ribosomes; the 19 amino-acid N-terminal signal sequence is cleaved and it converted to a 75 amino-acid prodefensin (brown). Subsequent cleavage of residues generates a 29–30 amino acid mature defensin (green). During phagocytosis (pathogens), defensin-rich primary granules fuse with phagocytic vacuoles and high concentrations of defensins are generated.
Figure 5. Promyelocytes in the bone marrow synthesize α-defensins. 94 amino acid preprodefensin (purple) is biosynthesized in the ribosomes; the 19 amino-acid N-terminal signal sequence is cleaved and it converted to a 75 amino-acid prodefensin (brown). Subsequent cleavage of residues generates a 29–30 amino acid mature defensin (green). During phagocytosis (pathogens), defensin-rich primary granules fuse with phagocytic vacuoles and high concentrations of defensins are generated.
Ijms 22 00859 g005
Figure 6. Possible mechanisms antimicrobial peptides can kill bacterial superbugs.
Figure 6. Possible mechanisms antimicrobial peptides can kill bacterial superbugs.
Ijms 22 00859 g006
Figure 7. Phage-derived antimicrobial techniques. Novel antimicrobial strategies derived from phages and their products. (a) Phages target specific bacterial pathogens and thus cause the lysis of that particular bacterial cell wall; (b) phages produce enzymes that target particular bacterial pathogens; (c) phages can be used to transfer antibiotic-sensitive genes into drug-resistant bacteria.
Figure 7. Phage-derived antimicrobial techniques. Novel antimicrobial strategies derived from phages and their products. (a) Phages target specific bacterial pathogens and thus cause the lysis of that particular bacterial cell wall; (b) phages produce enzymes that target particular bacterial pathogens; (c) phages can be used to transfer antibiotic-sensitive genes into drug-resistant bacteria.
Ijms 22 00859 g007
Figure 8. Mode of action of novel antibodies how they bind to multidrug-resistant (MDR) bacteria and present to macrophages and destroy them.
Figure 8. Mode of action of novel antibodies how they bind to multidrug-resistant (MDR) bacteria and present to macrophages and destroy them.
Ijms 22 00859 g008
Figure 9. Quorum sensing can be blocked by producing and apply chemical analogs, which render the bacteria unable to communicate and hence pathogenicity-related expression is affected.
Figure 9. Quorum sensing can be blocked by producing and apply chemical analogs, which render the bacteria unable to communicate and hence pathogenicity-related expression is affected.
Ijms 22 00859 g009
Table 1. Use of different nanoparticles as a weapon against drug-resistant bacteria.
Table 1. Use of different nanoparticles as a weapon against drug-resistant bacteria.
AgentTarget Bacteria/DiseasesMode of Action/DescriptionNotes (Advantages (a), Limitations (l), Combination Strategy (c))Ref.
Nanoparticles
AuNP with cationic surface chemistryGram-positive and Gram-negative bacteriaInteraction with cell membrane → formation of aggregates → bacterial cell lysis; cause protuberance(a) Unique electronic, sensing, optical, and biochemical properties[34,35,36,98]
CuONPVariety pathogens, including MRSA, E. coli, S. aureus, P. aeruginosa, N. meningitis, B. cereus, S. pyogenes, A. baumanniiROS → induce oxidative stress(c) Antibacterial activity enhanced by conjugation with AgNPs (photocatalytic activity attributed to the production of ROS)[1,38,39,40,99]
ZnONPs, colloidal ZnO suspensionMRSA, S. agalactiae, MRSE, MSSA, ESBL-producing E. coli and K. pneumoniae, Vibrio cholera, Campylobacter jejuni, E. faecalis, S. epidermidis, and other clinically relevant pathogensDisorganization and damage of cell, cell membrane after internalization; damage of proteins, lipids, and DNA via oxidative stress(l) Level of toxicity concentration dependent
(c) Antibacterial activity enhanced by conjugation with AgNPs
[1,41,42,43,44,45,46]
Nitric-oxide-releasing NPs (NONPs)Antibiotic-resistant and sensitive bacteria, i.e., K. pneumoniae, E. faecalis, S. pyogenes, E. coli, and P. aeruginosaFormation of cell toxic reactive nitrogen and oxygen intermediates, NO-associated lipid damage, iron depletion, inhibition of DNA repair enzymes(a) NO is unstable so spontaneously generate reactive intermediates[47,48,49,50,51,52,53,54]
Gold nanorodsP. aeruginosaConjugated with primary antibodies(a) Eradicate biofilms
(c) 1,2-Distearoyl-sn-glycero-3-phosphorylethanolamine (DSPE)
[60,69]
S. aureus and Propionibacterium acnesLocal hyperthermia by laser beam excited functionalized gold nanorods(a) Enhanced reduction of viable bacterial count[61]
AgNPs, colloidal AgNPsGram-positive and Gram-negative bacteria; drug-susceptible strains including Streptococcus spp., E. coli, and P. aeruginosa;
MRSA, MRSE, (erythromycin-resistant) S. pyogenes, (ampicillin-resistant) E. coli, MDR P. aeruginosa
AgNPs anchor to cell wall leading to increased cell permeability by structural changes → uncontrolled transport through cell membrane;
Membrane damage caused by AgNP produced free radicals;
released Ag+ ions inactivate enzymes by interacting with thiol groups of enzymes
(a) Bactericidal against Gram-positive as well as Gram-negative bacteria
(1) Dose-dependent
[55,74,75,76]
Biosynthetically produces AgNPs using fungus, yeast, bacteria, and plant extractsM. tuberculosis, P. aeruginosa, S. pneumoniae, MRSA, K. pneumoniae, MRSE, S. pyogenes, Bacillus spp., E. coli and S. typhiInhibit cell wall synthesis, protein synthesis, which is mediated by the 30 s ribosomal subunit, and nucleic acid synthesis(a) Strong antibacterial efficacy against various MDR pathogens[81,82,83]
Aluminum oxide NPsE. coli, Pseudomonas fluorescence, Staphylococcus aureus, Streptococcus mutans, Proteus vulgaris) Disruption of bacterial cell wall by producing ROS,
Serve as radical scavengers leading to distortion in bacterial cells
(a) Thermodynamically stable over a wide range of temperatures[88,89,90]
Ethionamide (ETH)-conjugated SiNPs
(silicon)
Multidrug-resistant M. tuberculosisEnhance solubility and permeability of ETH at different pH-values(a) Reduction of dosing frequency of ETH for the treatment of multidrug-resistant M. tuberculosis
(c) thermally carbonized-porous silicon (TCPSi) loaded with ethionamide (ETH)
[94]
Gallium (III) nano-formulationsDrug-resistant M. tuberculosisTargeted drug delivery,
Promotion of maturation of phagosome → increased macrophage-mediated killing,
Interruption of iron-mediated enzymatic reactions
(a) Active against resistant bacteria like M. tuberculosis, HIV[96]
Nano-photothermal therapy
AuNPsGram-positive bacteria, Gram-negative bacteriaElectromagnetic radiation absorbed by the NPs converted into heat → transferred via thermal conduction to bacteria in close proximities(c) Conjugated with vancomycin or amoxicillin[62,63,67]
Au nanorodsPathogenic E. coliGenerate heat that lyses bacteria(c) Heat generated by using continuous-wave laser irradiation or near-infrared laser[29]
Au nanorodsP. aeruginosaNanorod attach to the bacterial cell surface allows the cell to expose to near-infrared radiation(c) Conjugated with primary antibodies[1,69]
Multifunctional popcorn-shaped magnetic iron core-shell gold nanoparticlesSalmonella DT104Selective and irreparable cellular-damage(c) Conjugated with Salmonella DT104 specific antibody[70]
Polysiloxane polymers containing embedded methylene blue and AuNPsMRSA and E. coliLight-induced production of singlet oxygen and other reactive oxygen species by the methylene blue and gold nanoparticles enhanced activity of methylene blue(a) Significant reduction of viable cell count
(1) Require exposure to light and polymer formation
[1,71,72]
Table 2. Host defense peptides (HDP) against antibiotic-resistant bacteria.
Table 2. Host defense peptides (HDP) against antibiotic-resistant bacteria.
CategoryAgentTarget Bacteria/DiseasesMode of Action/DescriptionNotes (Advantages (a), Limitations (l), Combination Strategy (c))StatusRef
HDPBrevinin1 HYba1 and brevinin1 HYba2Several Gram-positive and Gram-negative bacteriaHemolytic, cytotoxic and antibacterial activities(a) Very low hemolysis on human erythrocytesHep 3B cancer cell line[101]
HDP (Defensin)68 fungal defensin-like peptides (fDLPs)A variety of bacterial, fungal and viral pathogens(i) Cationic amphipathic peptides having approximately 30 amino acid residues
(ii) From five genera named Apophysomyces, Trichosporon, Scedosporium, Beauveria, and Lichtheimia had been reported
(a) Higher antibacterial potential with lower cellular toxicitiesIn vitro and in vivo[105]
Scedosporisin (synthetic defensin)Gram-positive bacteria, vancomycin-resistant Enterococci, MRSAScedosporisin-2 killed bacteria more rapidly as compared to the antibiotic vancomycin(a) Low cytotoxicity and hemolysis on humanIn vitro[105]
HDP (Cathelicidin)Bactenecin, indolicidin, protegrins,…S. pyogenes and MRSA, VISA, ListeriaProduced by the immune systems of bovine and porcine(a) Broad bacterial lytic properties, stability and higher efficacyIn vitro and in vivo[106]
Human cathelicidin LL-37Antibiofilm activity against S. aureus and E. coliHuman immune system is known to produce only one type of cathelicidin precursor protein, hCAP18 → processed proteolytically to produce mature LL-37(1) Exact mechanisms of interaction between LL-37 and immune cells have not been yet clarifiedIn vitro and in vivo[107]
CATHPb1–6 (six novel cathelicidins identified from Python bivittatu)S. aureus (MRSA/VRSA)Involved in modulating macrophages/monocytes;
trafficking neutrophils to the site of infection and also enhance their bactericidal functions; increases levels of chemokines and reduces release of proinflammatory cytokines
(a) Provides protection via different administration routesIn vitro and in vivo (mice)[108]
Table 3. A list of different antimicrobial peptides (AMP) (antibiofilm peptides (ABP)) molecules that can be used against multidrug-resistant (pathogenic) bacteria.
Table 3. A list of different antimicrobial peptides (AMP) (antibiofilm peptides (ABP)) molecules that can be used against multidrug-resistant (pathogenic) bacteria.
CategoryAgentTarget Bacteria/DiseasesMode of Action/DescriptionNotes (Advantages (a), Limitations (l), Combination Strategy (c))StatusRef.
ABPNisinVariety of Gram-positiveCationic peptide(1) Requires proper optimization of pHIn vitro[114]
ABP + HDPCombination of peptides and defensin proteinsVariety of Gram-positive and Gram-negative bacteria; effective against fungi tooProteinaceous entities can either be cationic or amphipathic(a) Not require any specific protein binding sitesIn vitro[113]
ABPAntimicrobial peptides KT2 and RT2Antibiofilm activity against MDR enterohemorrhagic E. coli O157:H7Tryptophan-rich cationic
peptides permeabilize bacterial cell membranes → lead to death of cells by causing large damage or small obstructions that disturb transmembrane potential
(a) Not only prevent biofilm formation but also can eliminate mature biofilms
(l) Interactions with membrane and each other
(c) Combination with other antimicrobial compounds to enhance activity → lower concentration of antimicrobial compounds
In vivo[100,120]
Agelaia-MPI and Polybia-MPIIMDR Acinetobacter baumannii, several Gram-positive and Gram-negative bacteria, Mycobacteria as well as fungiIsolated from wasps; bactericidal activity along with antibiofilm activity(l) Production costs
(l) Peptidases and proteases lead to low stability of peptides in human serum → (c) Increased stability in combination with other molecules (e.g., polyethylene glycol)
In vitro[129]
Ocellatin-PT2–PT6Opportunistic pathogen Pseudomonas aeruginosaOcellatin-PT3 inhibits proliferation of established biofilms by directly killing bacterial cells(a) Novel antimicrobial agent(l) Works better in combination BS antibiotics
in vitro[130]
QS + ABP“RNAIII-inhibiting peptide” (RIP)Biofilm formation and ailments caused by S. aureusInhibition of phosphorylation of “target of RNAIII activating protein” → quorum sensing inhibition, prevention of MDR in bacteria(a) Inhibits cell adhesion and biofilm formationIn vitro and in vivo (cellulitis)[135,136]
Table 4. Bacteriophages as a therapeutic option against bacteria.
Table 4. Bacteriophages as a therapeutic option against bacteria.
AgentTarget Bacteria/DiseasesMode of Action/DescriptionNotes (Advantages (a), Limitations (l), Combination Strategy (c))StatusRef
Bacteriophages
ϕMR11, KP DP1, SA DP1, PA DP4, EC DP3E. faecalis, S. aureus, Klebsiella, A. baumannii, P. aeruginosa, and Escherichia coliPhages bind to specific receptors on bacterial cell surface → infects bacterial cells → production of endolysins that damage bacterial cell wall by hydrolyzing four main bonds of peptidoglycan,
Rupture of outer membrane via complexes (spanins) (Gram-negative bacteria)
(a) Applied externally and internally;
high affinity for specific bacterium (normal flora not attacked)
only one administration (replicative nature);
can survive in the gastric environment;
minimal side effects → ecologically safe; frequency of bacterial resistance to phages significantly lower compared with resistance to antibiotics
(l) Not very effective in mixed infections (narrow host range)
In vivo (mice)[19,169]
Lytic phage strain (KPP10)P. aeruginosaDecreased numbers of viable P. aeruginosa cells in blood, liver, and spleen as well as levels of inflammatory cytokines in blood and liver(c) Oral administrationAnimal models[153]
CD140Clostridium difficile-induced ileocolitisPhage administration prophylaxis against infection(1) Specific against Clostridium difficilehamster[148]
ØCDHM1–ØCDHM6, ØCDHS1, ENB6 and C33, Ø9882,
ØA392, and
KPP10
Clostridium difficile, vancomycin-resistant E. faecium, extended-spectrum β-lactamase producing E. coli, imipenem-resistant and MDR P. aeruginosaTreatment of gut-derived sepsis(a) Specifically act against bacterial pathogens
(a) Do not affect the natural bioflora
(a) Safer to use in humans
(1) They will be effective only if their favorable conditions exist
Hamsters and mice[150,154]
OMKO1MDR P. aeruginosaOuter membrane porin M (OprM) of the multidrug efflux systems MexAB and MexXY as a receptor-binding site(a) Specifically act against MDR P. aeruginosa
(a) Alters efflux pump mechanism to make the bacterium more susceptible to drugs
In vitro[154]
PBAB08 and PBAB25Acinetobacter baumanniiReduction of bacterial load, increase in serum IgE with a slight increase of GM-CSF, IL2, IL10, and IL17A(1) Inoculated in a cocktail and require properly set optimal conditionsmouse[163]
Mixture of three phagesCampylobacter jejuni and C. coliReduce bacterial colonization(1) Acquisition phage to resistancePoultry[143]
PlyF307 (phage lysin)MDR A. baumanniiLysing of bacterial cells(a) Inactive against eukaryotic cellsMouse[163]
Cpl-1 (phage endolysin)Streptococcus pneumoniaeReduced pulmonary bacterial counts and prevented bacteremia, systemic hypotension, and lactate increase as well as reduction of penicillin-susceptible pneumococci(1) Specific against pneumococciMouse[164]
PGHs (phage endolysins)MRSAThe peptidoglycan hydrolase enzyme targets the conserved regions and can destroy a wide range of mutant cell walls of bacteria(a) Active against mutant and resistant strains
(a) Also can clear static biofilms
In vitro and in Mouse[165]
PlyCD (prophage lysin)C. difficilePlyCD specifically targets the pathogenic C. difficile while not affecting other commensal bacteria(c) Phage lysins in combination with antibiotics more effective than antibiotics aloneEx vivo mouse colon model[166]
PlySs2 (phage endolysin)Streptococcus and Staphylococcus species, such as MRSALytic activity(a) High therapeutic potential compared to other lysinsMouse[167]
PlyC (phage endolysin)S. pyogenesLysins can cross the epithelial cell membrane to eradicate intracellular infections(a) Ability to traverse epithelial membranesmodel membranes[168]
PlyG (phage endolysin)Bacillus anthracisInterrupt vegetative cells; major advantage over antibiotics (attacking endospores)(a) Separate domains to recognize spores and vegetative cellsIn vitro[170]
Table 5. Different types of vaccines used to fight the bacterial multidrug resistance.
Table 5. Different types of vaccines used to fight the bacterial multidrug resistance.
AgentTarget Bacteria/DiseasesMode of Action/DescriptionNotes (Advantages (a), Limitations (l), Combination Strategy (c))In Vitro, In Vivo, Clinical Phase, Animal ModelRef
Vaccines
Tetra-subunit vaccineS. aureusComprising of two capsular polysaccharides and two virulence-associated proteins (ClfA and MntC)(c) Diminish the burden of the disease, thereby reducing use of antibioticsPhase 2b trial[175]
Three different
vaccines
C. difficileConstructed on C. difficile toxins A and B(1) More research is required for proper optimizations of toxin-based vaccines, including development and use of novel adjuvantsPhase 2 and 3 trials[189]
OprF/I fusion protein vaccineP. aeruginosaFounded on conserved outer membrane protein F/I fusion(a) Produce rapid immune response in healthy volunteersPhase 2/3 trials[175,190]
Vaccine NDV-3CandidaTargeting T cell target protein, Als3(a) Also protects from intravenous as well as skin and soft tissue infection with Staphylococcus aureusPhase 2 trials[184]
Antiresistance vaccinesMRSACloned internal region from transpeptidase domain from penicillin-binding protein (PBP2a) as DNA vaccine(a) More operational against drug-resistant strains by explicitly targeting resistant alleles of a conserved protein or by targeting proteins exclusively present in resistant clustersMouse[186]
Antiresistance vaccinesNeisseria meningitidisVaccination with purified recombinant PBP2 + passive immunization with anti-PBP2 rabbit IgG antibody(a) This vaccine candidate has a conserved region that is present in all strains of N. meningitidis and targeted by protective antibodiesMouse[175,188]
Table 6. Combination drug therapy to defeat the superbugs.
Table 6. Combination drug therapy to defeat the superbugs.
Combination Drug Therapy
AgentTarget Bacteria/
Diseases
Mode of Action/DescriptionNotes (Advantages (a), Limitations (l), Combination Strategy (c))Ref
Combination drug combination acting on diverse targets in different pathways
Rifampicin (R), isoniazid (H), ethambutol (E), and pyrazinamide (Z)M. tuberculosisRifampicin (RNA polymerase inhibitor), isoniazid (enoyl reductase subunit of fatty acid synthase), ethambutol (an inhibitor of arabinosyl transferases involved in cell wall biosynthesis) and pyrazinamide (mechanism of action poorly understood)(a) Method highly effective since a bacterium may develop resistance by changing one of its targets, the combination drug strategy will still be effective against at least the other two pathways[147,192,193]
Drug combinations acting on diverse targets in the same pathways
Clavulanic acidGram-positive bacteriaDegrades the β-lactamase enzyme, allowing the drug to destroy these microorganisms(c) Use of a β-lactam antibiotic (amoxicillin) and β-lactamase enzyme inhibitor (clavulanic acid)[187]
Drug combination acting on a single target, but in multiple dimensions
Streptogramins Two active molecules that bind to the adjacent sites in the 50S ribosomal subunit near the peptidyl transferase center(c) Both of these molecules are used simultaneously; they show 10–100-fold more potency as compared to using a single molecule alone[194,195,196]
Table 7. Novel antibodies are used against bacteria with and/or without drug combination.
Table 7. Novel antibodies are used against bacteria with and/or without drug combination.
AgentTarget Bacteria/DiseasesMode of Action/DescriptionNotes (Advantages (a), Limitations (l), Combination Strategy (c))In Vitro, In Vivo, Clinical Phase, Animal ModelRef
Antibodies
A1102 (humanized mouse gal-III mAb)K. pneumoniae ST258Passive immunization with A1102 before infection with ST258 → infection prophylaxis(1) Efficacy and exact role for protection in vivo not understoodIn vitro and in vivo in experimental models (mice)[197]
O25b-specific mAb ASN-4MCR-1-positive colistin-resistant ST131-H30 strainOopsonophagocytosis, endotoxin neutralization, and complement-mediated killing(a) Multiple mechanisms of actionIn vitro and in vivo in experimental models (Murine models)[198]
MEDI3902P. aeruginosaBispecific antibody targeting the P. aeruginosa type III secretion (T3S) protein PcrV and Psl exopolysaccharide(c) In combination with drug therapy (antibiotics) deliver assistance when used alongside antibioticsIn vivo (mice)[199]
Table 8. Different types of probiotics used to combat bacteria.
Table 8. Different types of probiotics used to combat bacteria.
AgentTarget Bacteria/DiseasesMode of Action/DescriptionNotes (Advantages (a), Limitations (l), Combination Strategy (c))StatusRef.
Probiotics
Lactobacillus and BifidobacteriumE. coli, Salmonella, Helicobacter pylori, Listeria monocytogenes and rotavirusLessen occurrence, time period, and/or ruthlessness of antibiotic-linked gastroenteritis → enhancing efficacy of these antibiotics(c) Simultaneous utilization of probiotics with antibioticsIn vivo[204]
Lactobacillus acidophilus strainP. aeruginosaInhibit development of drug-resistant bacteria by secretion of antibacterial chemicals including lactic acid, hydrogen peroxide,
diminishing MDRs ability to colonize the body → reducing use of antibiotics
(a) Reduced use of antibiotics and development of MDR by providing protection against intrinsic resistance strainsIn vitro[20,21]
LactobacilliMDR K. pneumoniaeUsed in place of antibiotic therapy(1) Require identification of more strainsIn vitro[205]
Propionibacterium freudenreichii freudenreichii B3523 (PF) and P. freudenreichii shermanii B4327 (PS))MDR Salmonella heidelberg (SH)Used as an alternative to antibiotics for preventing SH infections(a) Non-host gastrointestinal tract-derived probioticTurkey poults[206]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Khan, A.A.; Manzoor, K.N.; Sultan, A.; Saeed, M.; Rafique, M.; Noushad, S.; Talib, A.; Rentschler, S.; Deigner, H.-P. Pulling the Brakes on Fast and Furious Multiple Drug-Resistant (MDR) Bacteria. Int. J. Mol. Sci. 2021, 22, 859. https://doi.org/10.3390/ijms22020859

AMA Style

Khan AA, Manzoor KN, Sultan A, Saeed M, Rafique M, Noushad S, Talib A, Rentschler S, Deigner H-P. Pulling the Brakes on Fast and Furious Multiple Drug-Resistant (MDR) Bacteria. International Journal of Molecular Sciences. 2021; 22(2):859. https://doi.org/10.3390/ijms22020859

Chicago/Turabian Style

Khan, Abid Ali, Khanzadi Nazneen Manzoor, Aamir Sultan, Maria Saeed, Mahrukh Rafique, Sameen Noushad, Ayesha Talib, Simone Rentschler, and Hans-Peter Deigner. 2021. "Pulling the Brakes on Fast and Furious Multiple Drug-Resistant (MDR) Bacteria" International Journal of Molecular Sciences 22, no. 2: 859. https://doi.org/10.3390/ijms22020859

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