Novel Strategies to Combat Antimicrobial Resistance: History
Please note this is an old version of this entry, which may differ significantly from the current revision.

Antimicrobial resistance (AMR) is one of the most important global public health problems. The imprudent use of antibiotics in humans and animals has resulted in the emergence of antibiotic-resistant bacteria. The dissemination of these strains and their resistant determinants could endanger antibiotic efficacy. Therefore, there is an urgent need to identify and develop novel strategies to combat antibiotic resistance. 

  • antimicrobial resistance
  • antibiotics alternatives
  • quorum sensing
  • prebiotics
  • probiotics

1. Small Molecules (SMs)

SMs are non-peptide organic molecules that are synthetic or obtained from natural product extracts. They have drug-like properties that can interact with biological molecules, including protein and nucleic acids, and can alter their normal functions. The low molecular weight (~200–500 Da) and high hydrophilicity of these molecules allow their effective absorption by both host and pathogen barriers [1][2]. SMs can be modified to enhance the qualities desired for specific applications, such as stability and solubility under adverse environmental conditions. These properties can be exploited to enhance the SMs’ antimicrobial efficacy and their mass applicability. High-throughput screening (HTS) of SM libraries is commonly used for the development of antibacterial drugs and identification of SM candidates that inhibit either bacterial growth in whole-cell assays or the activity of a main bacterial enzyme or protein [3]. Indeed, a cost-effective, cell-based HTS expedient approach has been recently developed to enhance anti-bacterial molecule discovery [4][5]. SM antimicrobials targeting bacterial membranes are highly desired because they have low potential for resistance development by pathogens, can potentiate the activity of many antibiotics, are effective against slow-growing bacteria and biofilms, and have high stability in serum and good tissue penetration [6]. They have been reported to be effective against several MDR bacteria, such as E. coli, P. aeruginosa, Enterococcus faecium, methicillin resistant S. aureus (MRSA), Klebsiella pneumoniae, Acinetobacter baumannii, and Enterobacter species [7][8][9]. Recently, SMs were used for the treatment of plant pathogens such as Xanthomonas spp., Erwinia tracheiphila, Acidovorax citrulli and Salmonella infection [10][11][12][13], as well as for the potentiation of antibiotics, which can help in reducing the resistance of the treated bacteria [14].

2. Quorum-Sensing/Antivirulence Inhibitors

Bacterial cells adapt to their surrounding environment and regulate their density and behavior via a cell-to-cell communication process named quorum sensing (QS). This process is mediated by bacterial secretion of extracellular signaling molecules called autoinducers (AIs) [15]. The bacteria produce and release AIs to coordinate their gene expression for survival as multicellular organisms. Additionally, AIs are also key regulators of biofilm formation, stress adaptation, secondary metabolite production, swarming motility, enzyme production, and virulence factor production [16][17]. Active transport or diffusion is used to release autoinducers into the environment to achieve efficient communication between bacterial cells [18]. As the bacterial population density rises, AIs build up in the environment, and, after this reaches a certain threshold, bacteria use them as extracellular signaling molecules to adjust their density and coordinate their gene expression [19].
QS systems are based on three fundamental concepts. (1) One is the bacterial cells’ density: at a high cell density, the cumulative generation of AIs results in local accumulation at a high concentration, which facilitates detection and response. However, at a low cell density, the AIs diffuse away, as they are present at concentrations below the detection threshold [20]. (2) Receptors generated in the cytoplasm or on the membrane are used to identify AIs. (3) The recognition of AIs leads to increased bacterial synthesis of AIs in addition to stimulation of gene expression required for cooperative behaviors [21].
Generally, the bacterial QS systems are classified into three types: (1) LuxI/LuxR–type QS, which is found in Gram-negative bacteria and uses acyl-homoserine lactones (AHL) as signaling molecules [22], (2) oligopeptide-two-component-type QS, which is found in Gram-positive bacteria and utilizes oligopeptides as signaling molecules, and (3) luxS-encoded autoinducer-2 (AI-2) QS, a general system, which is found in both Gram-negative and Gram-positive bacteria and uses AI-2 as signaling molecules [16]. The AIs are categorized into acylated homoserine lactones (AHLs), utilized by Gram-negative bacteria, oligopeptides, utilized by Gram-positive bacteria, and furanosyl borate diester, utilized by Gram-negative and Gram-positive bacteria. In addition, there are other signaling molecules of the QS system called autoinducer-3 (AI-3), which are utilized by P. aeruginosa and do not belong to any of the previous classes. This complex network of signals allows the bacterial community to react and adapt to different environments [23].
Interrupting the connection system between bacterial cells results in a reduction in bacterial biofilm formation and pathogenicity [23][24]. Therefore, many strategies have been developed to hinder this connection and control the QS-dependent bacterial infections [25]. The inactivation, blocking, or degradation of QS signal molecules refers to QS inhibition or quorum quenching (QQ) [26]. The perfect QS inhibitors (QSIs) are low-mass compounds with a great selectivity for the QS regulator and no deleterious side effects on the bacterium or a potential eukaryotic host. They must also be chemically stable and extremely efficient [23]. QSIs may be natural or artificial molecules. In fact, many anti-QS compounds are isolated from plants and microbes. 

3. Probiotics

Probiotics are microorganisms that live in a symbiotic relationship with the host. They provide health benefits and perform several biological functions when provided in adequate amounts. Probiotics were discovered and selected based on certain criteria, which ensure safety and effectiveness requirements [27]. The FAO/WHO have specified several parameters that should be assessed in vitro when selecting probiotics, such as safety, efficacy, cost effectiveness, function, and technological and physiological applications. The selected probiotics can be characterized by a lack of pathogenicity, tolerance to changes in the human gastrointestinal microenvironment, capacity for adherence to and colonization of the intestinal epithelium, antimicrobial activity, genetic and phenotypic stability, and immunomodulatory capabilities [28]. Several in vitro tests can be used to evaluate the efficacy of probiotics before starting the clinical trials, such as the agar spot test [29], the agar well diffusion assay [30], microdilution [31], antibiofilm analysis [32], 3D cell cultures, and use of human tissues and animal models [33][34][35].
Additionally, probiotics have been found to help with a variety of pathological conditions, including constipation, diarrhea, polycystic ovarian syndrome, ulcerative colitis, stress and anxiety, inflammatory bowel disease, breast cancer, and diabetes [36]. Probiotics are classified into four categories: (1) viable and active probiotics, (2) viable/non-active probiotics, in the forms of spores or vegetative cells, (3) dead/nonviable probiotics [37], and (4) next-generation probiotics [38]. The biological properties of probiotics have been extensively investigated, but only a few studies focused on their antimicrobial properties as novel antibiotic alternatives

4. Prebiotics

Prebiotics are defined as “non-digestible food materials that beneficially impact the host by selectively enhancing the growth and/or metabolism of bacterial species inhabiting the GIT, and thus improve the host health” [39]. Prebiotics are also defined as “any substrate preferentially consumed by host microorganisms that result in increasing the health benefit” [39]. Evidence indicates that prebiotics are promising alternatives in the medicinal and food industries. Prebiotics are characterized by (1) the ability to withstand the acidic environment during passage through the digestive tract (GIT) [40], (2) resistance to digestive enzymes but susceptibility to probiotic-hydrolyzing enzymes [41][42], (3) non-direct absorbance [43], (4) maintenance of gut microbial ecology [41], and (5) the ability to stimulate the host immune response [40]. Prebiotics are non-digestible oligosaccharides (fructans, inulins, xylans, galactans, and mannan), fibers (pectin, non-starch polysaccharides (such as β-glucan), xylooligosaccharides, andisomaltooligosaccharide), and seeds containing gums [44][45]. Human milk oligosaccharide is considered an endogenous source of prebiotics that increases the population of Bifidobacterium spp. in breastfeeding newborns, thereby enhancing their immunity [46]. To use prebiotics as alternatives to antibiotics, specific criteria must be met. Prebiotics should have a well-identified source chemical composition and structure, be a pure product, be at a suitable dose, and have been assessed in animal models or 3D cells to confirm their safety and beneficial health impact on the microflora [45]. When used as feed additives for livestock and poultry, prebiotics have shown an ability to improve host health and productivity via selective stimulation of proliferation and metabolism of the gut microbiota, such as Akkermansia spp., Christensenella spp., Propionibacterium spp., Faecalibacterium spp. and Roseburia spp., Lactobacillus spp., and Bifidobacterium spp. [47]. In the context of their benefits for human health, fermentation of the prebiotics konjac glucomannan hydrolysate and inulin in a batch culture of human feces has been associated with the production of short-chain fatty acids and proliferation of the genera Bifdobacterium, Lactobacillus and Enterococcus [48]. A meta-analysis of 64 studies reported that addition of dietary fibers stimulated Bifidobacterium spp. and Lactobacillus spp. resulting in an increase in the concentration of fecal short-chain fatty acids (SCFAs) in healthy adults [49]. Moreover, other studies have revealed that these bacteria play a key role in maintaining the composition of GIT microflora, enhancing the food and mineral absorption, and promoting the host defense system [50].
Prebiotics have also shown potential to eliminate harmful bacteria, such as Salmonella, Campylobacter, Clostridium and E. coli [51][52]; however, their mechanism of action remains to be elucidated. Fermentation of prebiotics by the gut’s resident microbiota and the subsequent production of SCFAs results in a reduction in the gut pH, which, in turn, make the condition unfavorable for the growth and colonization of invading pathogens [53]. It was reported that the activity of probiotic Bifidobacterium strains against C. difficile was significantly stimulated in the presence of five prebiotics (oligosaccharides) [54]. Similarly, the activity of Pediococcus acidilactici was enhanced against E. coli, Salmonella, E. fecalis and S. aureus in the presence of garlic and basil (natural prebiotics) [55]. Moreover, the use of the prebiotics mannan-oligosaccharides and fructooligosaccharides, as poultry feed supplements, reduced the colonization of Campylobacter and Salmonella in the GIT of poultry [52]. Supplementation of weaning pigs with prebiotic oligofructose resulted in a significant increase in the number of Bifidobacteria and Lactobacilli and a significant reduction in the number of clostridia, enterobacteria, and enterococci [56]. A reduction in disease severity was observed following treatment of patients with C. difficile-associated diarrhea with inulin and oligofructose [57].

5. Antimicrobial Peptides (AMPs)

AMPs are naturally produced by various immune cells and play a vital role in the innate immune systems of various animals, plants, and microorganisms [58]. AMPs, have a wide spectrum of antimicrobial activity against bacteria, fungi, viruses, and parasites [59]. In addition to their antimicrobial activities, AMPs possess biological functions, such as immune modulation, angiogenesis, antitumor activity, and wound healing [60][61]. AMPs are considered promising alternatives to antibiotics, due to many advantages; (1) they act on multiple target sites on the intracellular targets and plasma membranes of pathogenic bacteria, (2) they have potent killing activity against drug-resistant bacteria [59][62], (3) they are a component of the innate immune system, (4) their natural production by the host cells saves time and energy compared to antibody synthesis via acquired immunity, and (5) they reach the target sites faster than immunoglobulin [58]. AMPs are classified into several subgroups based on their amino acid sequences (10–100), the net charge of the peptide (+2 to +9), and their protein structure and sources [63]. These subgroups include (1) anionic AMPs, which consist of 5–70 amino acid residues and have a net charge range of −1 to −8 [64]. Their structural characteristics include α-helical peptides and cyclic cystine knots [64]. They use the negatively charged content of the microbial membrane to form salt bridges, leading to disruption of the microbial membrane [65]. (2) Cationic α-helical AMPs are ≤40 amino acids in length (50% hydrophobic in nature) and have a charge of +2 to +9 and the C-terminus amidated [66]. The structure of cationic α-helical AMPs is disordered in aqueous solutions [67]. They are capable of forming amphiphilic structures when interacting with target cells [68]. (3) Cationic AMPs, where the peptides consist of 2–8 cysteine residues forming 1–4 pairs of intramolecular disulfide bonds. These disulfide bonds play a crucial role in β-sheet AMP stabilization and biological functions [69]. (4) Extended cationic AMPs containing amino acids including tryptophan, arginine, proline, histidine and glycine and lacks the regular secondary structures [68]. (5) Fragments from antimicrobial proteins that have a broad-spectrum bactericidal effect, such as lysozyme [70]. The helix–loop–helix (HLH) region in the human and chicken lysozyme has a strong effect on the growth of pathogenic bacteria and fungi [71].
Many antimicrobial peptides, isolated from different sources, have shown activity against a wide variety of pathogenic bacteria. For example, magainin-2 (α-helix (TFE)) was originally isolated from frogs and has been shown to be active against P. gingivalis, F. nucleatum, P. intermedia, E. coli, and S. aureus [72]. Cecropin and cecropin P1 (α-helix structure) were isolated from silk moth and pig, respectively. These peptides inhibited the growth of E. coli ML-35p [73], S. aureus, B. subtilis, M. luteus, P. aeruginosa, S. Typhimurium, S. marcescens and E. coli [74]. In addition, apo-lactoferrin (α-helix structure), discovered in bovine and human PDB code (1BOL), inhibited the growth of E. coli O157:H7 [75]. Melittin (α-helix structure), extracted from bee, was active against S. salivarius, S. mitis, S. mutans, S. sanguinis, S. sobrinus, L. casei, and E. faecalis [76]. Temporin A and temporin L, extracted from frog, were active against MRSA [77], B. megaterium Bm11, S. aureus Cowan I, and E.coli D21 [78]. Moreover, buforin II and clavanin A (α-helix structure), discovered in extended toad and Styela clava, respectively, possess inhibitory activity against B. subtilis, S. aureus, E. coli, S. Typhimurium [79], E. coli ML35p and L. monocytogenes [80]. Furthermore, protegrin-1 (β-sheet structure), isolated from human and porcine, has demonstrated antagonistic activity against MRSA and P. aeruginosa [81]. Tachyplesin-I (β-sheet structure), discovered in horseshoe crab, was shown to inhibit the growth of S. Typhimurium [82]. Furthermore, hepcidin (β-structure), extracted from humans, has demonstrated capabilities to inhibit E. coli, S. aureus, and S. epidermidis [83]. Daptomycin (cyclic lipopeptide membrane), isolated from Streptomyces roseosporus, can kill MRSA [84] and nisin (lantibiotic), isolated from L. lactis, was shown to kill MRSA, S. pneumoniae, Enterococci and C. difficile [85]. NPSRQERR [P1], PDENK [P2], and VHTAPK [P3]), derived from L. rhamnosus GG, showed inhibitory activity against APEC in chicken [86].

6. Bacteriophages

Phages or bacterial viruses are obligate parasites that infect bacteria and archaea. Phages are classified according to their size, shape, type of nucleic acid and mechanism of action in the host bacterial cell [87]. The genomic sequences of phages range from a few thousand base pairs to 498 kilobase pairs in phage G, the biggest phage ever sequenced [88]. Some phages have wide host ranges; however, the majority of them have high host specificity [89]. Based on the last classification, bacteriophages are classified as virulent (e.g., lysis of the bacterial cells to release new phages) or temperate (e.g., incorporation of its genetic material in the host genome and the host changes its genetic characters) bacteriophages [87]. In vitro trials showed that phages have many advantages over antibiotics, including (1) high host specificity (phages can target one strain of bacteria without perturbing the human or animal gut microbiota, while antibiotics do not distinguish between pathogenic and beneficial bacteria); (2) cost effectiveness and time saving [90]; (3) inhibition of multi-drug-resistant bacteria while antibiotics increase them; (4) ease of delivery to the target site and ability to penetrate the blood–brain barrier [91]; (5) no antagonistic effect detected between phages when given as a cocktail (mixture of different phages); (6) that phages could prevent biofilm formation [92]; and (7) phages might be used as an alternative in antibiotic-allergic patients; however, very few reports discussed this [93]. Several studies have reported that phage therapy development might be a potential solution to bacterial antibiotic resistance as well as the treatment of numerous bacterial infectious illnesses [94]. As the poultry gut is considered the main reservoir of Campylobacter, most Campylobacter phages have been isolated from avian GIT. Many studies have demonstrated that administration of individual phages resulted in a significant decrease in the number of Campylobacter without altering the GIT microbiota [95]. The most multi-drug-resistant bacterial strains, such as E. faecium, S. aureus, K. pneumoniae, A. baumannii, P. aeruginosa, and Enterobacter spp, have been reported to cause serious human diseases. 
NPs are considered one of the potential alternative candidates of antibiotics for controlling multi-drug-resistant microorganisms [96]. They have demonstrated therapeutic potential due to their unique chemical and physical characteristics [97]. NPs have a tiny size (1–100 nm) with a large surface area to interact with target organisms [98]. They can be chemically or naturally synthesized from different sources with variable chemical structures that allow different chemical functionalities [99]. NPs exhibit antimicrobial activities through targeting critical active sites in pathogens, leading to partial or complete inhibition [99]. Organic or inorganic (metal and metal oxide) NPs can be synthesized from different sources. Inorganic NPs possess bactericidal activity against bacteria using multiple mechanisms and, therefore, they are denominated “nanobactericides”. The nanobactericides activity of inorganic NPs is attributed to (1) their tiny size [99], (2) formation of weak and nonspecific interactions with bacterial surfaces [100], (3) Van der Waals forces (distance-dependent interactions between atoms or molecules) [101], and (4) attachment through specific receptor–ligand bonds [102]. Therefore, the bacterial cells’ susceptibility to NPs depends on their structural components as well as their growth rate [103]. Gram-positive bacteria are more susceptible to NPs than Gram-negative bacteria. Gram-positive bacteria have a permeable and negatively charged cell wall, making them an easy target for NP penetration, while the non-porous cell walls of Gram-negative bacteria serve as penetration barriers against the NPs [104]. Moreover, bacteria with slow growth rates are more sensitive to NPs than those with rapid growth rates. This is due to different stress-response genes’ expressions in fast-growing bacteria [103].

8. Organic Acids (OAs)

Organic acids are widely used as antimicrobials in food processing and many industries [105]. Bacteria, fungi, and yeast play a critical role in the synthesis of organic acids during their lifecycle with high yields that can be achieved using cost-effective substrates. The bioproduction of OAs depends on many factors, such as the species of microorganism, inoculum size, substrate or carbon source, and environmental conditions (aeration, temperature, pH and agitation) [106]. Increasing the acidity by adding an acidulant or integrating natural fermentation is one of the commonly used methods to minimize and/or inhibit microbial growth. Using organic acids as an alternative to antibiotics depends on several factors, such as chemical formula structure, molecular weight, the value of the dissociation constant (pKa), minimum inhibitory concentration (MIC), nature of the microorganism, and exposure time to the food [107], where the pKa is an important criterion because of the undissociated part of the acid that is responsible for the antimicrobial effect.
Common OAs that are microbially produced and commercially used for microbial inhibition and food processing include (1) acetic acid, which is produced after fermentation of many substrates, such as glucose, lactose, and sucrose. This has the European code E260 and is used in the production of vinegar, stabilizer, flavor enhancer, and firming agent [106]. (2) Adipic acid is a crucial intermediate in the pathways of cyclic alkanes, long-chain aliphatic dicarboxylic acids and cyclic alcohols [108]. It is commonly used in the synthesis of polymers, plasticizers, nylon, clothing, automobile parts, and lubricant [109]. (3) Butyric acid is used in the fuel, plastic pharmaceutical, and textile industries. (4) Citric acid is used as a pH regulator, flavor enhancer, pharmaceutical reagent, and firming agent, in addition to its antimicrobial properties [110]. It is also used in soft candies, baked goods, gelatins, snacks, dairy products, and cheese warps as an antimicrobial and in the fuel industry. (5) Lactic acid is used in dairy products, biochemical processes, and the leather, pharmaceutical, textile, and biodegradable biopolymer industries [111]. (6) Malic acid, which is an intermediate compound in the tricarboxylic acid cycle, is naturally found in fruits including apricot, blackberry, cherry, mango, peach, and plum. It has been used in the food, water treatment, textile, metals, and pharmaceutical industries [112]. (7) Phenyl lactic acid naturally exists in honey and has an effective and broad microbial activity against bacteria, fungi, and yeast [113]. (8) Propionic acid naturally presents in apples, strawberries, grains, and cheese [114]. Adding propionic acid to chick diets was found to improve their growth, exert an antimicrobial effect in the intestine, and reduce the yellowness of the meat [115]. (9) Succinic acid is used in food preservation, perfume intermediates, herbicide production, and the plastics and textiles industries [116].

9. Essential Oils (EOs)

Essential oils are volatile, aromatic, and oily liquids extracted from plant parts, such as seeds, leaves, buds, twigs, flowers, bark, herbs, wood, fruits, and roots [117]. Plants generate EOs as a natural defense against pathogens and herbivore feeding by reducing the appetite of herbivores. As a result, the Department of Health and Human Services has designated EOs as safe antibacterial additives [118]. To date, about 3000 EOs have been recorded, 300 of which are economically valued in the pharmaceutical, agronomic, food, sanitary, cosmetics and perfume industries [119]. EOs are complex natural mixes that contain anywhere from 20 to 60 distinct components in various proportions. The antibacterial effects of EOs are dictated by their primary ingredients (85%), which include terpenes, terpenoids, and aromatic and aliphatic groups from different natural sources [120]. These groups are characterized by low molecular weights, which are limonene (31%) and α-phellandrene (36%) in Anethum graveolens leaf oil, d-limonene (over 80%) in citrus peel oils, α/β-thujone (57%) and camphor (24%) in Artemisia herba-alba oil, carvacrol (30%) and thymol (27%) in Origanum compactum oil, α-phellandrene (36%) and limonene (31%) in Anethum graveolens leaf oil, menthol (59%) and menthone (19%) in Mentha piperita oil, and carvone (58%) and d-limonene (37%) in Anethum graveolens seed oil [121].
Menthol, pulegone, linalool, thymol and camphor, extracted from Salvia lavandulifolia Lavandulaangustifolia, Mentha piperita, Mentha pulegium, and Satureja montana, respectively, have shown antagonistic effects against P. aeruginosa, S. pyogenes, S. mutans, S. sanguis, S. salivarius, and E. feacalis [122]. Thymol and carvacrol, extracted from many sources, such as Origanum compactum, Lavandula latifolia, Lavandula angustifolia, Rosmarinus officinalis, Origanum vulgare, Thymus vulgaris, and Thymus zygis chemotype thymol, have shown activity against S. aureus, C. hystoliticum, C. perfringens, E. coli O157:H7, S. Typhimurium, S. Enteritidis, and L. monocytogenes [122][123]. Linalool, linalyl acetate α-terpineol, β-caryophyllene and nerol, produced by Mentha citrata Ehrh, have shown inhibitory effects against P. aeruginosa, K. pneumoniae, E. coli (DH5α), E. coli (MTCC 723) and S. Typhimurium, S. aureus, S. epidermidis and S. mutans [124]. Additionally, E-anethole, linalool, 1,8-cineole, α-pinene, camphor, camphene, menthol, menthone, and limonene, produced by Ocimum basilicum, Rosmarinus officinalis, O. majorana, Mentha piperita, Thymus vulgaris, and Pimpinella anisum, have shown activity against C. perfringens [125]. Epilobium parviflorum, Salvia desoleana, S. sclarea, and Allium sativum were reported to produce palmitic acid, linoleic acid and α-linolenic acid, which have shown an ability to inhibit E. faecalis, S. aureus, P. aeruginosa, S. epidermidis, and E. coli [126]. Moreover, cinnamomum was reported to produce cinnamaldehyde, which was shown to inhibit E. coli, S. aureus, and S. Typhimurium [127]. Dipterocarpus gracilis was reported to produce elemicin and geranyl acetate, which were shown to suppress B. cereus and Proteus mirabilis [128].

10. Fecal Microbial Transplant (FMT)

FMT is a process of transferring processed fecal material from the intestine of a healthy donor to the intestine of a recipient patient [129]. Processed fecal matter can be administered to the recipient through several methods, such as a nasoduodenal tube [130], nasojejunal tube [131], colonoscopy, or retention enema [132]. Colonoscopy administration of fresh or frozen and thawed fecal matter from stool banks into the cecum and colon of C. difficile-infected children in Maryland resulted in complete resolution of CDI in recipients as well as reductions in both AMR and multidrug resistance genes. Moreover, FMT resulted in sustained elevations in alpha diversity post-FMT as well as significant changes in beta diversity, in addition to improving the biosynthetic pathways [132]. In another study performed in mice, when a combination of FMT and lytic phages was used for treatment of S. Typhimurium, a complete clearance of Salmonella, a reduction in inflammatory cytokines, and restoration of the intestinal microbial diversity was observed [133]. Additionally, patients with MRSA enteritis who were treated with FMT through nasointestinal tube, jejunostomy fistula tube or gastrostomy fistula tube had negative stool cultures for MRSA, and gut microbiota analysis also revealed that all recipients developed donor-related bacterial diversity [134]. FMT has also shown satisfactory results when given to patients infected with beta-lactamase-producing Enterobacteriaceae. Four weeks after the first FMT dose through nasoduodenal tube, decolonization was detected in 20% of the recipients, with recipients’ microbial composition showing a shift toward the donors’ microbial diversity [135]. Administration of fresh fecal matter using nasojejuneal tube into K. pneumoniae-infected patients stopped sepsis and resulted in elimination of K. pneumoniae, as demonstrated in the blood cultures and general improvement of the health status. Moreover, a restoration of microbial diversity was observed after 6 weeks of treatment [131]. Administration of fresh fecal matter using nasoduodenal gastroscope into the third part of the duodenum of chronically infected hepatitis B patients (CHB) also resulted in suppression of the hepatitis B virus as well as clearance of the HBeAg [130]. In addition to its use in the treatment of microbial infections, FMT has also shown potential to improve non-infectious GIT conditions, such as ulcerative colitis, where it resulted in a lowered pediatric ulcerative colitis activity index (PUCAI) following completion of the FMT treatment course [136]. Similar observations were made for other inflammatory bowel diseases [137]. Moreover, several experiments were undertaken to assess the safety and efficacy of FMT in treating other extra-intestinal disorders, such as obesity and metabolic disorders [138] and psoriatic arthritis [139].

11. Vaccines

Vaccines are preparations used to stimulate the body’s immune response against diseases by exploiting the ability of the human immune system to respond to, and remember, the antigens of pathogens. Several vaccines have been developed to make a revolutionary change in the world, such as fowl (avian) cholera, anthrax, polio, norovirus, rift valley fever, and rabies vaccines [140][141][142]. Vaccines play a pivotal role in reducing the need for antibiotics and controlling the emergence of AMR bacterial strains [143]. Vaccines reduce the burden of antimicrobial resistance through disease prevention and thus reducing the use of antibiotics [144][145]. This occurs as a vaccine curbing the ability of the pathogen to establish a foothold in the host, by conferring immunity against these pathogens, thus minimizing the chances of some bacterial mutations and the development and spread of resistant genes to other bacteria [146][147]. For instance, a 67% reduction in the circulation of penicillin-resistant invasive pneumococcal strains was demonstrated in a group of children that received pneumococcal conjugate vaccine 9 (PCV9) compared to controls in South Africa [148]. Conjugate vaccines combine weak antigens with strong antigens (which serve as the carriers) to increase the response of the body to the weaker antigen. In this context, the typhoid conjugate vaccine (TCV) has been introduced in children in order to protect them extensively from drug-resistant S. Typhi [149]. It has been observed that TCV can avert 44% of typhoid cases, of which 35% are resistant to antibiotics [150]. Salmonellosis, caused by Salmonella spp., is one of the most common zoonotic diseases associated with consumption of dairy and beef [151]. S. enterica serotype Dublin, which infects cattle and can be shed in milk, colostrum, and feces, also poses a threat to public health. S. Dublin causes bloodstream infections in humans, with a relatively high case fatality [152]. Data from the CDC (Centers for Disease Control and Prevention) showed that Salmonella Dublin infections caused more hospitalization during 1996–2004. Additionally, a higher percentage of Salmonella isolates were resistant to more than seven classes of antimicrobial drugs during 2005–2013 (50.8%) compared to only 2.4% during the period 1996–2004. Resistant S. enterica causes at least 100,000 foodborne human infections annually [153]. A commercial modified-live Salmonella Dublin vaccine (EnterVene-d) is approved by the USDA for use in calves, but vaccination does not reduce the likelihood of contamination or the risk to public health; it only improves clinical outcome [154]. C. perfringens enterotoxin (CPE) and Shiga-toxin-producing E. coli (STEC) are also common causes of food poisoning. Research has been conducted on possible development of a vaccine for CPE and STEC in the form of a bivalent food poisoning vaccine. The bivalent vaccine uses a fused protein (Stx2B-C-CPE) consisting of the B subunit of E. coli Shiga toxin 2 fused to CPE to enhance its antigenicity [155]. Two other extremely important bacteria, C. jejuni and C. perfringens, have been the subject of many vaccine studies in poultry [156][157][158][159]. The efficacy and commercial potential of these vaccines has been described and reviewed in detail elsewhere [160]. It is worth mentioning that the conserved N-glycan heptasaccharide conjugated to GlycoTag, or fused to the E. coli lipopolysaccharide core, has shown tremendous potential to reduce C. jejuni colonization in the gastrointestinal tract of chickens by up to 10 log10 [161]. Despite the demonstrated efficacy of this vaccine, its commercialization remains murky.
Vaccination can cause indirect effects on infections. While resistance is a predictable outcome of antibiotic use, resistance to vaccines is very rare [162]. Vaccines are administered prophylactically, whereas antibiotics are administered only once symptoms have begun to show. Thus, by the time antibiotics are administered, there are possibly already millions of copies of the pathogen, raising the probability of mutation occurring. Vaccines prevent the pathogen from gaining a foothold and multiplying in the first place. In many cases, the use of vaccines has globally eradicated some diseases, while decreases of 95% in the incidence of diseases like diphtheria, tetanus, and pertussis have been observed [163]. Much progress has been made in bacterial vaccine development. Bacterial vaccines can and should help address the global AMR problem. It is reasonable to believe that reductions in MDR infections as well as the prevention thereof can be achieved using bacterial vaccines. Attention should also be paid to the role of veterinary bacterial vaccines to reduce antibiotic use in animals, especially food-producing animals. The role of bacterial vaccines is set to expand dramatically in response to the crisis of AMR and MDR [164]. Although vaccines against major AMR pathogens are still missing, predictions of the impact of vaccines against AMR hint that vaccines could have a significant impact in controlling resistance [165].

12. Antibodies

Antibodies, also known as immunoglobulins, are the most diverse set of proteins [166]. They have two major functions: antigen binding and effector functions [167]. Most of these effector functions are induced via the constant Fc (fragment crystallizable region, the tail region of an antibody) of the antibody, which can interact with complementary proteins and specialized Fc-receptors. This can activate or inhibit pathways, depending on the type of receptor [168]. Therapeutic antibacterial monoclonal antibodies (mAbs) are gaining traction as an alternative in treating infectious diseases [145]. Monoclonal antibodies could offer more effective ways of addressing antibiotic resistance and bacterial infections due to their superb specificity, by which they target conserved pathways. This allows for fewer off-target effects and less selective pressure for cross-resistance to other mAbs or antibiotics. Monoclonal antibodies also do not harm the beneficial microbiome [169][170] In 1879, Amil von Behring and Shibasaburo Kitasato were the first to develop antibodies called antitoxins that target specific toxins. Blood-serum-containing antitoxin was directly injected to convey immunity to diphtheria in humans [171][172]. The toxin/antitoxin approach provided a steady treatment against numerous pathogens, such as Haemophilus influenzae, Neisseria meningitides, Corynebacterium diphtheria, Clostridium tetani, S. pneumonia, and Group A Streptococcus. However, the antitoxin approach exhibits heterogeneity between lots, allergic reactions, and a limited spectrum, eventually leading to its replacement by antibiotics in 1930 [173]. Antibiotic production peaked for the next 80 years because of their safe application and their ease of formulation and manufacture [174]. However, due to the development of the hybridoma technology and recent advances in mAb engineering, awareness has shifted back to antibacterial mAbs [175].

13. Conclusions and Future Perspectives

The alarming rise in the emergence and spread of AMR and the associated global impact necessitate an urgent intervention of alternatives to combat the growing threat of antibiotic-resistant bacteria. Beside their potential adverse events, the inappropriate prescription or dispensing of antibiotics for humans and their irrational use in animal agriculture are among the factors contributing to the growing incidence of AMR in humans. As such, averting AMR could be achieved by focusing on two aspects: one is to implement antimicrobial stewardship programs through promoting prudent use of antibiotics in healthcare and agricultural settings, and the other is to develop effective antimicrobial alternatives to substitute antibiotics in animal food production. In fact, developed countries, including Europe and North America, have taken steps to ban the use of sub-therapeutic doses of antibiotics as growth promoters in livestock and poultry production; however, such steps have yet to be implemented in developing countries. Thus, a global solution is crucial to tackle AMR, as the world has become increasingly interconnected. Research efforts have been made to limit AMR in both humans and animals by exploring various interventions, including SMs, QSIs, probiotics, prebiotics, phage therapy, nanoparticles, EOs, AMPs, OAs, FMT, vaccines, and immune-based strategies, as potential replacements for antibiotics. Despite the promising role of most of these strategies in promoting host immunity and in antagonizing a range of human and animal pathogens, their variable effects, combined with their limited spectrum, safety concerns, and poor efficacy, are among the potential limitations to their use. Nonetheless, the exuberant development of molecular technologies may improve the efficacy of existing strategies and reduce their limitations. For instance, the recent breakthroughs in CRISPR-Cas9-based genome editing offer a revolutionary platform for designing safe and effective vaccines. Likewise, computational molecular biology has directed vaccine development towards genome-based reverse vaccinology approaches, a process of analyzing the whole genome sequence for identification of novel target antigens. The processes of 16S rRNA next-generation sequencing and bioinformatics analysis have enabled the identification of bacterial strains at species level. This technology can be utilized to not only identify novel probiotic species, but also to develop a consortium of beneficial microbes, which may offer a safer and acceptable alternative to FMT. With millions of people travelling around the world and the uncontrollable spread of AMR, holistic AMR control requires global solidarity to expand and implement robust antimicrobial stewardship programs in both medical and veterinary practices.

This entry is adapted from the peer-reviewed paper 10.3390/antibiotics12020274

References

  1. Leeson, P.D.; Springthorpe, B. The influence of drug-like concepts on decision-making in medicinal chemistry. Nat. Rev. Drug Discov. 2007, 6, 881–890.
  2. Hong-Geller, E.; Micheva-Viteva, S. Small molecule screens to identify inhibitors of infectious disease. In Drug Discovery; El Shelmy, H.A., Ed.; InTech: London, UK, 2013; pp. 157–175.
  3. Selin, C.; Stietz, M.S.; Blanchard, J.E.; Gehrke, S.S.; Bernard, S.; Hall, D.G.; Brown, E.D.; Cardona, S.T. A Pipeline for Screening Small Molecules with Growth Inhibitory Activity against Burkholderia cenocepacia. PLoS ONE 2015, 10, e0128587.
  4. Kumar, A.; Drozd, M.; Pina-Mimbela, R.; Xu, X.; Helmy, Y.A.; Antwi, J.; Fuchs, J.R.; Nislow, C.; Templeton, J.; Blackall, P.J.; et al. Novel Anti-Campylobacter Compounds Identified Using High Throughput Screening of a Pre-selected Enriched Small Molecules Library. Front. Microbiol. 2016, 7, 405.
  5. Helmy, Y.A.; Kathayat, D.; Ghanem, M.; Jung, K.; Closs, G., Jr.; Deblais, L.; Srivastava, V.; El-Gazzar, M.; Rajashekara, G. Identification and characterization of novel small molecule inhibitors to control Mycoplasma gallisepticum infection in chickens. Vet. Microbiol. 2020, 247, 108799.
  6. Mingeot-Leclercq, M.-P.; Decout, J.-L. Bacterial lipid membranes as promising targets to fight antimicrobial resistance, molecular foundations and illustration through the renewal of aminoglycoside antibiotics and emergence of amphiphilic aminoglycosides. MedChemComm 2016, 7, 586–611.
  7. Garg, S.K.; Singh, O.; Juneja, D.; Tyagi, N.; Khurana, A.S.; Qamra, A.; Motlekar, S.; Barkate, H. Resurgence of Polymyxin B for MDR/XDR Gram-Negative Infections: An Overview of Current Evidence. Crit. Care Res. Pract. 2017, 2017, 3635609.
  8. Hubbard, A.T.; Barker, R.; Rehal, R.; Vandera, K.A.; Harvey, R.D.; Coates, A.R. Mechanism of Action of a Membrane-Active Quinoline-Based Antimicrobial on Natural and Model Bacterial Membranes. Biochemistry 2017, 56, 1163–1174.
  9. Hart, E.M.; Mitchell, A.M.; Konovalova, A.; Grabowicz, M.; Sheng, J.; Han, X.; Rodriguez-Rivera, F.P.; Schwaid, A.G.; Malinverni, J.C.; Balibar, C.J. A small-molecule inhibitor of BamA impervious to efflux and the outer membrane permeability barrier. Proc. Natl. Acad. Sci. USA 2019, 116, 21748–21757.
  10. Vrisman, C.M.; Deblais, L.; Helmy, Y.A.; Johnson, R.; Rajashekara, G.; Miller, S.A. Discovery and Characterization of Low-Molecular Weight Inhibitors of Erwinia tracheiphila. Phytopathology 2020, 110, 989–998.
  11. Srivastava, V.; Deblais, L.; Kathayat, D.; Rotondo, F.; Helmy, Y.A.; Miller, S.A.; Rajashekara, G. Novel Small Molecule Growth Inhibitors of Xanthomonas spp. Causing Bacterial Spot of Tomato. Phytopathology 2021, 111, 940–953.
  12. Lu, Y.; Deblais, L.; Rajashekara, G.; Miller, S.A.; Helmy, Y.A.; Zhang, H.; Wu, P.; Qiu, Y.; Xu, X. High-throughput screening reveals small molecule modulators inhibitory to Acidovorax citrulli. Plant Pathol. 2020, 69, 818–826.
  13. Deblais, L.; Vrisman, C.; Kathayat, D.; Helmy, Y.A.; Miller, S.A.; Rajashekara, G. Imidazole and Methoxybenzylamine Growth Inhibitors Reduce Salmonella Persistence in Tomato Plant Tissues. J. Food Prot. 2019, 82, 997–1006.
  14. Kathayat, D.; Antony, L.; Deblais, L.; Helmy, Y.A.; Scaria, J.; Rajashekara, G. Small Molecule Adjuvants Potentiate Colistin Activity and Attenuate Resistance Development in Escherichia coli by Affecting pmrAB System. Infect. Drug Resist. 2020, 13, 2205–2222.
  15. Gao, X.; Li, C.; He, R.; Zhang, Y.; Wang, B.; Zhang, Z.-H.; Ho, C.-T. Research advances on biogenic amines in traditional fermented foods: Emphasis on formation mechanism, detection and control methods. Food Chem. 2023, 405, 134911.
  16. Miller, M.B.; Bassler, B.L. Quorum sensing in bacteria. Annu. Rev. Microbiol. 2001, 55, 165–199.
  17. Preda, V.G.; Săndulescu, O. Communication is the key: Biofilms, quorum sensing, formation and prevention. Discoveries 2019, 7, e100.
  18. Kose-Mutlu, B.; Ergon-Can, T.; Koyuncu, I.; Lee, C.-H. Quorum quenching for effective control of biofouling in membrane bioreactor: A comprehensive review of approaches, applications, and challenges. Environ. Eng. Res. 2019, 24, 543–558.
  19. Williams, P.; Cámara, M. Quorum sensing and environmental adaptation in Pseudomonas aeruginosa: A tale of regulatory networks and multifunctional signal molecules. Curr. Opin. Microbiol. 2009, 12, 182–191.
  20. Kaplan, H.B.; Greenberg, E.P. Diffusion of autoinducer is involved in regulation of the Vibrio fischeri luminescence system. J. Bacteriol. 1985, 163, 1210–1214.
  21. Seed, P.C.; Passador, L.; Iglewski, B.H. Activation of the Pseudomonas aeruginosa lasI gene by LasR and the Pseudomonas autoinducer PAI: An autoinduction regulatory hierarchy. J. Bacteriol. 1995, 177, 654–659.
  22. Fuqua, C.; Greenberg, E.P. Listening in on bacteria: Acyl-homoserine lactone signalling. Nat. Rev. Mol. Cell Biol. 2002, 3, 685–695.
  23. Papenfort, K.; Bassler, B.L. Quorum sensing signal-response systems in Gram-negative bacteria. Nat. Rev. Microbiol. 2016, 14, 576–588.
  24. Chan, W.C.; Coyle, B.J.; Williams, P. Virulence Regulation and Quorum Sensing in Staphylococcal Infections: Competitive AgrC Antagonists as Quorum Sensing Inhibitors. J. Med. Chem. 2004, 47, 4633–4641.
  25. Singh, V.K.; Kavita, K.; Prabhakaran, R.; Jha, B. Cis-9-octadecenoic acid from the rhizospheric bacterium Stenotrophomonas maltophilia BJ01 shows quorum quenching and anti-biofilm activities. Biofouling 2013, 29, 855–867.
  26. Andrei, S.; Droc, G.; Stefan, G. FDA approved antibacterial drugs: 2018–2019. Discoveries 2019, 7, e102.
  27. Silva, D.R.; de Cássia Orlandi Sardi, J.; de Souza Pitangui, N.; Roque, S.M.; da Silva, A.C.B.; Rosalen, P.L. Probiotics as an alternative antimicrobial therapy: Current reality and future directions. J. Funct. Foods 2020, 73, 104080.
  28. de Melo Pereira, G.V.; de Oliveira Coelho, B.; Júnior, A.I.M.; Thomaz-Soccol, V.; Soccol, C.R. How to select a probiotic? A review and update of methods and criteria. Biotechnol. Adv. 2018, 36, 2060–2076.
  29. Tharmaraj, N.; Shah, N.P. Antimicrobial effects of probiotics against selected pathogenic and spoilage bacteria in cheese-based dips. Int. Food Res. J. 2009, 16, 261–276.
  30. Parente, E.; Brienza, C.; Moles, M.; Ricciardi, A. A comparison of methods for the measurement of bacteriocin activity. J. Microbiol. Methods 1995, 22, 95–108.
  31. Adimpong, D.B.; Nielsen, D.S.; Sørensen, K.I.; Derkx, P.M.; Jespersen, L. Genotypic characterization and safety assessment of lactic acid bacteria from indigenous African fermented food products. BMC Microbiol. 2012, 12, 75.
  32. Flemming, H.-C.; Wingender, J.; Szewzyk, U.; Steinberg, P.; Rice, S.A.; Kjelleberg, S. Biofilms: An emergent form of bacterial life. Nat. Rev. Microbiol. 2016, 14, 563–575.
  33. Helmy, Y.A.; Closs, G., Jr.; Jung, K.; Kathayat, D.; Vlasova, A.; Rajashekara, G. Effect of Probiotic E. coli Nissle 1917 Supplementation on the Growth Performance, Immune Responses, Intestinal Morphology, and Gut Microbes of Campylobacter jejuni Infected Chickens. Infect. Immun. 2022, 90, e0033722.
  34. Helmy, Y.A.; Kassem, I.I.; Rajashekara, G. Immuno-modulatory effect of probiotic E. coli Nissle 1917 in polarized human colonic cells against Campylobacter jejuni infection. Gut Microbes 2021, 13, 1–16.
  35. Helmy, Y.A.; Kassem, I.I.; Kumar, A.; Rajashekara, G. In vitro evaluation of the impact of the probiotic E. coli Nissle 1917 on Campylobacter jejuni’s invasion and intracellular survival in human colonic cells. Front. Microbiol. 2017, 8, 1588.
  36. Kechagia, M.; Basoulis, D.; Konstantopoulou, S.; Dimitriadi, D.; Gyftopoulou, K.; Skarmoutsou, N.; Fakiri, E.M. Health benefits of probiotics: A review. Int. Sch. Res. Not. 2013, 2013, 481651.
  37. Zendeboodi, F.; Khorshidian, N.; Mortazavian, A.M.; da Cruz, A.G. Probiotic: Conceptualization from a new approach. Curr. Opin. Food Sci. 2020, 32, 103–123.
  38. Saarela, M.H. Safety aspects of next generation probiotics. Curr. Opin. Food Sci. 2019, 30, 8–13.
  39. Gibson, G.R.; Hutkins, R.; Sanders, M.E.; Prescott, S.L.; Reimer, R.A.; Salminen, S.J.; Scott, K.; Stanton, C.; Swanson, K.S.; Cani, P.D.; et al. Expert consensus document: The International Scientific Association for Probiotics and Prebiotics (ISAPP) consensus statement on the definition and scope of prebiotics. Nat. Rev. Gastroenterol. Hepatol. 2017, 14, 491–502.
  40. Manning, T.S.; Gibson, G.R. Prebiotics. Best Pract. Res. Clin. Gastroenterol. 2004, 18, 287–298.
  41. Mohanty, D.; Misra, S.; Mohapatra, S.; Sahu, P.S. Prebiotics and synbiotics: Recent concepts in nutrition. Food Biosci. 2018, 26, 152–160.
  42. Davani-Davari, D.; Negahdaripour, M.; Karimzadeh, I.; Seifan, M.; Mohkam, M.; Masoumi, S.J.; Berenjian, A.; Ghasemi, Y. Prebiotics: Definition, Types, Sources, Mechanisms, and Clinical Applications. Foods 2019, 8, 92.
  43. Sarangi, N.R.; Babu, L.K.; Kumar, A.; Pradhan, C.R.; Pati, P.K.; Mishra, J.P. Effect of dietary supplementation of prebiotic, probiotic, and synbiotic on growth performance and carcass characteristics of broiler chickens. Vet. World 2016, 9, 313–319.
  44. Pourabedin, M.; Zhao, X. Prebiotics and gut microbiota in chickens. FEMS Microbiol. Lett. 2015, 362, fnv122.
  45. Solis-Cruz, B.; Hernandez-Patlan, D.; Hargis, B.M.; Tellez, G. Use of prebiotics as an alternative to antibiotic growth promoters in the poultry industry. In Prebiotics and Probiotics-Potential Benefits in Nutrition and Health; IntechOpen: London, UK, 2019.
  46. Rajendran, S.R.C.K.; Okolie, C.L.; Udenigwe, C.C.; Mason, B. Structural features underlying prebiotic activity of conventional and potential prebiotic oligosaccharides in food and health. J. Food Biochem. 2017, 41, e12389.
  47. Cunningham, M.; Azcarate-Peril, M.A.; Barnard, A.; Benoit, V.; Grimaldi, R.; Guyonnet, D.; Holscher, H.D.; Hunter, K.; Manurung, S.; Obis, D.; et al. Shaping the Future of Probiotics and Prebiotics. Trends Microbiol. 2021, 29, 667–685.
  48. Connolly, M.L.; Lovegrove, J.A.; Tuohy, K.M. Konjac glucomannan hydrolysate beneficially modulates bacterial composition and activity within the faecal microbiota. J. Funct. Foods 2010, 2, 219–224.
  49. So, D.; Whelan, K.; Rossi, M.; Morrison, M.; Holtmann, G.; Kelly, J.T.; Shanahan, E.R.; Staudacher, H.M.; Campbell, K.L. Dietary fiber intervention on gut microbiota composition in healthy adults: A systematic review and meta-analysis. Am. J. Clin. Nutr. 2018, 107, 965–983.
  50. Wilson, B.; Whelan, K. Prebiotic inulin-type fructans and galacto-oligosaccharides: Definition, specificity, function, and application in gastrointestinal disorders. J. Gastroenterol. Hepatol. 2017, 32, 64–68.
  51. Micciche, A.C.; Foley, S.L.; Pavlidis, H.O.; McIntyre, D.R.; Ricke, S.C. A Review of Prebiotics Against Salmonella in Poultry: Current and Future Potential for Microbiome Research Applications. Front. Vet. Sci. 2018, 5, 191.
  52. Kim, S.A.; Jang, M.J.; Kim, S.Y.; Yang, Y.; Pavlidis, H.O.; Ricke, S.C. Potential for Prebiotics as Feed Additives to Limit Foodborne Campylobacter Establishment in the Poultry Gastrointestinal Tract. Front. Microbiol. 2019, 10, 91.
  53. Elshaghabee, F.M.F.; Rokana, N. Mitigation of antibiotic resistance using probiotics, prebiotics and synbiotics. A review. Environ. Chem. Lett. 2022, 20, 1295–1308.
  54. Kondepudi, K.K.; Ambalam, P.; Nilsson, I.; Wadström, T.; Ljungh, A. Prebiotic-non-digestible oligosaccharides preference of probiotic bifidobacteria and antimicrobial activity against Clostridium difficile. Anaerobe 2012, 18, 489–497.
  55. Koruri, S.S.; Chowdhury, R.; Bhattacharya, P. Potentiation of functional and antimicrobial activities through synergistic growth of probiotic Pediococcus acidilactici with natural prebiotics (garlic, basil). In Microbes in the Spotlight: Recent Progress in the Understanding of Beneficial and Harmful Microorganisms; Brown Walker Press: Irvine, CA, USA, 2016; pp. 219–224.
  56. Bomba, A.; Nemcová, R.; Gancarcíková, S.; Herich, R.; Guba, P.; Mudronová, D. Improvement of the probiotic effect of micro-organisms by their combination with maltodextrins, fructo-oligosaccharides and polyunsaturated fatty acids. Br. J. Nutr. 2002, 88, S95–S99.
  57. Lewis, S.; Burmeister, S.; Brazier, J. Effect of the prebiotic oligofructose on relapse of Clostridium difficile-associated diarrhea: A randomized, controlled study. Clin. Gastroenterol. Hepatol. 2005, 3, 442–448.
  58. Zhang, Q.-Y.; Yan, Z.-B.; Meng, Y.-M.; Hong, X.-Y.; Shao, G.; Ma, J.-J.; Cheng, X.-R.; Liu, J.; Kang, J.; Fu, C.-Y. Antimicrobial peptides: Mechanism of action, activity and clinical potential. Mil. Med. Res. 2021, 8, 48.
  59. Lei, J.; Sun, L.; Huang, S.; Zhu, C.; Li, P.; He, J.; Mackey, V.; Coy, D.H.; He, Q. The antimicrobial peptides and their potential clinical applications. Am. J. Transl. Res. 2019, 11, 3919–3931.
  60. Mookherjee, N.; Anderson, M.A.; Haagsman, H.P.; Davidson, D.J. Antimicrobial host defence peptides: Functions and clinical potential. Nat. Rev. Drug Discov. 2020, 19, 311–332.
  61. Pfalzgraff, A.; Brandenburg, K.; Weindl, G. Antimicrobial Peptides and Their Therapeutic Potential for Bacterial Skin Infections and Wounds. Front. Pharmacol. 2018, 9, 281.
  62. Mwangi, J.; Hao, X.; Lai, R.; Zhang, Z.-Y. Antimicrobial peptides: New hope in the war against multidrug resistance. Zool. Res. 2019, 40, 488–505.
  63. Di Somma, A.; Moretta, A.; Canè, C.; Cirillo, A.; Duilio, A. Antimicrobial and Antibiofilm Peptides. Biomolecules 2020, 10, 652.
  64. Dennison, S.R.; Harris, F.; Mura, M.; Phoenix, D.A. An Atlas of Anionic Antimicrobial Peptides from Amphibians. Curr. Protein Pept. Sci. 2018, 19, 823–838.
  65. Almarwani, B.; Phambu, N.; Hamada, Y.Z.; Sunda-Meya, A. Interactions of an Anionic Antimicrobial Peptide with Zinc(II): Application to Bacterial Mimetic Membranes. Langmuir 2020, 36, 14554–14562.
  66. Teixeira, V.; Feio, M.J.; Bastos, M. Role of lipids in the interaction of antimicrobial peptides with membranes. Prog. Lipid Res. 2012, 51, 149–177.
  67. Gennaro, R.; Zanetti, M. Structural features and biological activities of the cathelicidin-derived antimicrobial peptides. Biopolymers 2000, 55, 31–49.
  68. Lewies, A.; Wentzel, J.F.; Jacobs, G.; Du Plessis, L.H. The Potential Use of Natural and Structural Analogues of Antimicrobial Peptides in the Fight against Neglected Tropical Diseases. Molecules 2015, 20, 15392–15433.
  69. Koehbach, J.; Craik, D.J. The Vast Structural Diversity of Antimicrobial Peptides. Trends Pharmacol. Sci. 2019, 40, 517–528.
  70. Starling, S. Innate immunity: A new way out for lysozyme. Nat. Rev. Gastroenterol. Hepatol. 2017, 14, 567.
  71. Ibrahim, H.R.; Thomas, U.; Pellegrini, A. A helix-loop-helix peptide at the upper lip of the active site cleft of lysozyme confers potent antimicrobial activity with membrane permeabilization action. J. Biol. Chem. 2001, 276, 43767–43774.
  72. Genco, C.A.; Maloy, W.L.; Kari, U.P.; Motley, M. Antimicrobial activity of magainin analogues against anaerobic oral pathogens. Int. J. Antimicrob. Agents 2003, 21, 75–78.
  73. Silvestro, L.; Weiser, J.N.; Axelsen, P.H. Antibacterial and antimembrane activities of cecropin A in Escherichia coli. Antimicrob. Agents Chemother. 2000, 44, 602–607.
  74. Pillai, A.; Ueno, S.; Zhang, H.; Lee, J.M.; Kato, Y. Cecropin P1 and novel nematode cecropins: A bacteria-inducible antimicrobial peptide family in the nematode Ascaris suum. Biochem. J. 2005, 390, 207–214.
  75. Franco, I.; Pérez, M.D.; Castillo, E.; Calvo, M.; Sánchez, L. Effect of high pressure on the structure and antibacterial activity of bovine lactoferrin treated in different media. J. Dairy Res. 2013, 80, 283–290.
  76. Leandro, L.F.; Mendes, C.A.; Casemiro, L.A.; Vinholis, A.H.; Cunha, W.R.; de Almeida, R.; Martins, C.H. Antimicrobial activity of apitoxin, melittin and phospholipase A2; of honey bee (Apis mellifera) venom against oral pathogens. An. Acad. Bras. Cienc. 2015, 87, 147–155.
  77. Wang, W.; Yang, W.; Du, S.; Xi, X.; Ma, C.; Wang, L.; Zhou, M.; Chen, T. Bioevaluation and Targeted Modification of Temporin-FL From the Skin Secretion of Dark-Spotted Frog (Pelophylax nigromaculatus). Front. Mol. Biosci. 2021, 8, 707013.
  78. Brancaccio, D.; Pizzo, E.; Cafaro, V.; Notomista, E.; De Lise, F.; Bosso, A.; Gaglione, R.; Merlino, F.; Novellino, E.; Ungaro, F.; et al. Antimicrobial peptide Temporin-L complexed with anionic cyclodextrins results in a potent and safe agent against sessile bacteria. Int. J. Pharm. 2020, 584, 119437.
  79. Park, C.B.; Yi, K.S.; Matsuzaki, K.; Kim, M.S.; Kim, S.C. Structure-activity analysis of buforin II, a histone H2A-derived antimicrobial peptide: The proline hinge is responsible for the cell-penetrating ability of buforin II. Proc. Natl. Acad. Sci. USA 2000, 97, 8245–8250.
  80. Lee, I.H.; Cho, Y.; Lehrer, R.I. Effects of pH and salinity on the antimicrobial properties of clavanins. Infect. Immun. 1997, 65, 2898–2903.
  81. Steinberg, D.A.; Hurst, M.A.; Fujii, C.A.; Kung, A.H.; Ho, J.F.; Cheng, F.C.; Loury, D.J.; Fiddes, J.C. Protegrin-1: A broad-spectrum, rapidly microbicidal peptide with in vivo activity. Antimicrob. Agents Chemother. 1997, 41, 1738–1742.
  82. Dai, J.G.; Xie, H.W.; Jin, G.; Wang, W.G.; Zhang, Y.; Guo, Y. Preliminary study on high-level expression of tandem-arranged tachyplesin-encoding gene in Bacillus subtilis Wb800 and its antibacterial activity. Mar. Biotechnol. 2009, 11, 109–117.
  83. Michels, K.; Nemeth, E.; Ganz, T.; Mehrad, B. Hepcidin and Host Defense against Infectious Diseases. PLoS Pathog. 2015, 11, e1004998.
  84. Blaskovich, M.A.T.; Hansford, K.A.; Gong, Y.; Butler, M.S.; Muldoon, C.; Huang, J.X.; Ramu, S.; Silva, A.B.; Cheng, M.; Kavanagh, A.M.; et al. Protein-inspired antibiotics active against vancomycin- and daptomycin-resistant bacteria. Nat. Commun. 2018, 9, 22.
  85. Shin, J.M.; Gwak, J.W.; Kamarajan, P.; Fenno, J.C.; Rickard, A.H.; Kapila, Y.L. Biomedical applications of nisin. J. Appl. Microbiol. 2016, 120, 1449–1465.
  86. Kathayat, D.; Closs, G.; Helmy Yosra, A.; Lokesh, D.; Ranjit, S.; Rajashekara, G. Peptides Affecting the Outer Membrane Lipid Asymmetry System (MlaA-OmpC/F) Reduce Avian Pathogenic Escherichia coli (APEC) Colonization in Chickens. Appl. Environ. Microbiol. 2021, 87, e00567-21.
  87. Engelkirk, P.G.; Duben-Engelkirk, J.; Fader, R.C. Burton’s Microbiology for the Health Sciences; Jones & Bartlett Publishers: Burlington, MA, USA, 2020.
  88. Sabour, P.M.; Griffiths, M.W. Bacteriophages in the Control of Food-and Waterborne Pathogens; American Society for Microbiology Press: Washington, DC, USA, 2010.
  89. Romero-Calle, D.; Guimarães Benevides, R.; Góes-Neto, A.; Billington, C. Bacteriophages as alternatives to antibiotics in clinical care. Antibiotics 2019, 8, 138.
  90. Golkar, Z.; Bagasra, O.; Pace, D.G. Bacteriophage therapy: A potential solution for the antibiotic resistance crisis. J. Infect. Dev. Ctries. 2014, 8, 129–136.
  91. Wittebole, X.; De Roock, S.; Opal, S.M. A historical overview of bacteriophage therapy as an alternative to antibiotics for the treatment of bacterial pathogens. Virulence 2014, 5, 226–235.
  92. Azeredo, J.; Sutherland, I.W. The use of phages for the removal of infectious biofilms. Curr. Pharm. Biotechnol. 2008, 9, 261–266.
  93. Ligonenko, O.V.; Borysenko, M.; Digtyar, I.; Ivashchenko, D.; Zubakha, A.; Chorna, I.; Shumeyko, I.; Storozhenko, O.; Gorb, L.; Ligonenko, O.O. Application of bacteriophages in complex of treatment of a shot-gun wounds of soft tissues in the patients, suffering multiple allergy for antibiotics. Klin. Khirurhiia 2015, 10, 65–66.
  94. El-Shibiny, A.; El-Sahhar, S. Bacteriophages: The possible solution to treat infections caused by pathogenic bacteria. Can. J. Microbiol. 2017, 63, 865–879.
  95. Kashoma, I.P.; Srivastava, V.; Rajashekara, G. Advances in Vaccines for Controlling Campylobacter in Poultry. In Food Safety in Poultry Meat Production; Venkitanarayanan, K., Thakur, S., Ricke, S.C., Eds.; Springer International Publishing: Cham, Switzerland, 2019; pp. 191–210.
  96. Lee, N.-Y.; Ko, W.-C.; Hsueh, P.-R. Nanoparticles in the treatment of infections caused by multidrug-resistant organisms. Front. Pharmacol. 2019, 10, 1153.
  97. Hemeg, H.A. Nanomaterials for alternative antibacterial therapy. Int. J. Nanomed. 2017, 12, 8211.
  98. World Health Organization. Antimicrobial Resistance: Global Report on Surveillance; World Health Organization: Geneva, Switzerland, 2014.
  99. Slavin, Y.N.; Asnis, J.; Häfeli, U.O.; Bach, H. Metal nanoparticles: Understanding the mechanisms behind antibacterial activity. J. Nanobiotechnol. 2017, 15, 65.
  100. Li, H.; Chen, Q.; Zhao, J.; Urmila, K. Enhancing the antimicrobial activity of natural extraction using the synthetic ultrasmall metal nanoparticles. Sci. Rep. 2015, 5, 11033.
  101. Armentano, I.; Arciola, C.R.; Fortunati, E.; Ferrari, D.; Mattioli, S.; Amoroso, C.F.; Rizzo, J.; Kenny, J.M.; Imbriani, M.; Visai, L. The interaction of bacteria with engineered nanostructured polymeric materials: A review. Sci. World J. 2014, 2014, 410423.
  102. Gao, W.; Thamphiwatana, S.; Angsantikul, P.; Zhang, L. Nanoparticle approaches against bacterial infections. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2014, 6, 532–547.
  103. Khan, S.T.; Musarrat, J.; Al-Khedhairy, A.A. Countering drug resistance, infectious diseases, and sepsis using metal and metal oxides nanoparticles: Current status. Colloids Surf. B Biointerfaces 2016, 146, 70–83.
  104. Zaidi, S.; Misba, L.; Khan, A.U. Nano-therapeutics: A revolution in infection control in post antibiotic era. Nanomedicine 2017, 13, 2281–2301.
  105. Dittoe, D.K.; Ricke, S.C.; Kiess, A.S. Organic Acids and Potential for Modifying the Avian Gastrointestinal Tract and Reducing Pathogens and Disease. Front. Vet. Sci. 2018, 5, 216.
  106. Coban, H.B. Organic acids as antimicrobial food agents: Applications and microbial productions. Bioprocess Biosyst. Eng. 2020, 43, 569–591.
  107. Davidson, P.M.; Sofos, J.N.; Branen, A.L. Antimicrobials in Food; CRC Press: Boca Raton, FL, USA, 2005.
  108. Polen, T.; Spelberg, M.; Bott, M. Toward biotechnological production of adipic acid and precursors from biorenewables. J. Biotechnol. 2013, 167, 75–84.
  109. Blach, P.; Böstrom, Z.; Franceschi-Messant, S.; Lattes, A.; Perez, E.; Rico-Lattes, I. Recyclable process for sustainable adipic acid production in microemulsions. Tetrahedron 2010, 66, 7124–7128.
  110. Ciriminna, R.; Meneguzzo, F.; Delisi, R.; Pagliaro, M. Citric acid: Emerging applications of key biotechnology industrial product. Chem. Cent. J. 2017, 11, 22.
  111. Abdel-Rahman, M.A.; Tashiro, Y.; Sonomoto, K. Lactic acid production from lignocellulose-derived sugars using lactic acid bacteria: Overview and limits. J. Biotechnol. 2011, 156, 286–301.
  112. Marques, C.; Sotiles, A.R.; Farias, F.O.; Oliveira, G.; Mitterer-Daltoé, M.L.; Masson, M.L. Full physicochemical characterization of malic acid: Emphasis in the potential as food ingredient and application in pectin gels. Arab. J. Chem. 2020, 13, 9118–9129.
  113. Rodríguez-Pazo, N.; Salgado, J.M.; Cortés-Diéguez, S.; Domínguez, J.M. Biotechnological production of phenyllactic acid and biosurfactants from trimming vine shoot hydrolyzates by microbial coculture fermentation. Appl. Biochem. Biotechnol. 2013, 169, 2175–2188.
  114. Hashemi, S.M.B.; Roohi, R. Kinetic models for production of propionic acid by Propionibacter freudenrechii subsp. shermanii and Propionibacterium freudenreichii subsp. freudenreichii in date syrup during sonication treatments. Biocatal. Agric. Biotechnol. 2019, 17, 613–619.
  115. Galli, G.M.; Aniecevski, E.; Petrolli, T.G.; da Rosa, G.; Boiago, M.M.; Simões, C.A.D.P.; Wagner, R.; Copetti, P.M.; Morsch, V.M.; Araujo, D.N.; et al. Growth performance and meat quality of broilers fed with microencapsulated organic acids. Anim. Feed. Sci. Technol. 2021, 271, 114706.
  116. Kuenz, A.; Hoffmann, L.; Goy, K.; Bromann, S.; Prüße, U. High-level production of succinic acid from crude glycerol by a wild type organism. Catalysts 2020, 10, 470.
  117. Brenes, A.; Roura, E. Essential oils in poultry nutrition: Main effects and modes of action. Anim. Feed. Sci. Technol. 2010, 158, 1–14.
  118. Stefanakis, M.K.; Touloupakis, E.; Anastasopoulos, E.; Ghanotakis, D.; Katerinopoulos, H.E.; Makridis, P. Antibacterial activity of essential oils from plants of the genus Origanum. Food Control 2013, 34, 539–546.
  119. Djilani, A.; Dicko, A.J.N. The therapeutic benefits of essential oils. Nutr. Well-Being Health 2012, 7, 155–179.
  120. Chouhan, S.; Sharma, K.; Guleria, S. Antimicrobial Activity of Some Essential Oils-Present Status and Future Perspectives. Medicines 2017, 4, 58.
  121. Shaaban, H.A.E.; El-Ghorab, A.H.; Shibamoto, T. Bioactivity of essential oils and their volatile aroma components: Review. J. Essent. Oil Res. 2012, 24, 203–212.
  122. Nikolić, M.; Jovanović, K.K.; Marković, T.; Marković, D.; Gligorijević, N.; Radulović, S.; Soković, M. Chemical composition, antimicrobial, and cytotoxic properties of five Lamiaceae essential oils. Ind. Crops Prod. 2014, 61, 225–232.
  123. Kachur, K.; Suntres, Z. The antibacterial properties of phenolic isomers, carvacrol and thymol. Crit. Rev. Food Sci. Nutr. 2020, 60, 3042–3053.
  124. Verma, S.K.; Goswami, P.; Verma, R.S.; Padalia, R.C.; Chauhan, A.; Singh, V.R.; Darokar, M.P. Chemical composition and antimicrobial activity of bergamot-mint (Mentha citrata Ehrh.) essential oils isolated from the herbage and aqueous distillate using different methods. Ind. Crops Prod. 2016, 91, 152–160.
  125. Radaelli, M.; da Silva, B.P.; Weidlich, L.; Hoehne, L.; Flach, A.; da Costa, L.A.; Ethur, E.M. Antimicrobial activities of six essential oils commonly used as condiments in Brazil against Clostridium perfringens. Braz. J. Microbiol. 2016, 47, 424–430.
  126. Bajer, T.; Šilha, D.; Ventura, K.; Bajerová, P. Composition and antimicrobial activity of the essential oil, distilled aromatic water and herbal infusion from Epilobium parviflorum Schreb. Ind. Crops Prod. 2017, 100, 95–105.
  127. Zhang, Y.; Liu, X.; Wang, Y.; Jiang, P.; Quek, S. Antibacterial activity and mechanism of cinnamon essential oil against Escherichia coli and Staphylococcus aureus. Food Control 2016, 59, 282–289.
  128. El Kolli, M.; Laouer, H.; El Kolli, H.; Akkal, S.; Sahli, F. Chemical analysis, antimicrobial and anti-oxidative properties of Daucus gracilis essential oil and its mechanism of action. Asian Pac. J. Trop. Biomed. 2016, 6, 8–15.
  129. Gupta, S.; Allen-Vercoe, E.; Petrof, E.O. Fecal microbiota transplantation: In perspective. Ther. Adv. Gastroenterol. 2016, 9, 229–239.
  130. Chauhan, A.; Kumar, R.; Sharma, S.; Mahanta, M.; Vayuuru, S.K.; Nayak, B.; Kumar, S. Fecal microbiota transplantation in Hepatitis B e antigen-positive chronic Hepatitis B patients: A pilot study. Dig. Dis. Sci. 2021, 66, 873–880.
  131. Ueckermann, V.; Hoosien, E.; De Villiers, N.; Geldenhuys, J. Fecal microbial transplantation for the treatment of persistent multidrug-resistant Klebsiella pneumoniae infection in a critically ill patient. Case Rep. Infect. Dis. 2020, 2020, 8462659.
  132. Hourigan, S.K.; Ahn, M.; Gibson, K.M.; Pérez-Losada, M.; Felix, G.; Weidner, M.; Leibowitz, I.; Niederhuber, J.E.; Sears, C.L.; Crandall, K.A. Fecal transplant in children with Clostridioides difficile gives sustained reduction in antimicrobial resistance and potential pathogen burden. Open Forum Infect. Dis. 2019, 6, ofz379.
  133. Wang, X.; Xing, Y.; Ji, Y.; Xi, H.; Liu, X.; Yang, L.; Lei, L.; Han, W.; Gu, J. The Combination of Phages and Faecal Microbiota Transplantation Can Effectively Treat Mouse Colitis Caused by Salmonella enterica Serovar Typhimurium. Front. Microbiol. 2022, 13, 944495.
  134. Wei, Y.; Gong, J.; Zhu, W.; Guo, D.; Gu, L.; Li, N.; Li, J. Fecal microbiota transplantation restores dysbiosis in patients with methicillin resistant Staphylococcus aureus enterocolitis. BMC Infect. Dis. 2015, 15, 265.
  135. Singh, R.; de Groot, P.F.; Geerlings, S.E.; Hodiamont, C.J.; Belzer, C.; Berge, I.; de Vos, W.M.; Bemelman, F.J.; Nieuwdorp, M. Fecal microbiota transplantation against intestinal colonization by extended spectrum beta-lactamase producing Enterobacteriaceae: A proof of principle study. BMC Res Notes 2018, 11, 190.
  136. Mańkowska-Wierzbicka, D.; Stelmach-Mardas, M.; Gabryel, M.; Tomczak, H.; Skrzypczak-Zielińska, M.; Zakerska-Banaszak, O.; Sowińska, A.; Mahadea, D.; Baturo, A.; Wolko, Ł. The effectiveness of multi-session FMT treatment in active ulcerative colitis patients: A pilot study. Biomedicines 2020, 8, 268.
  137. Paramsothy, S.; Paramsothy, R.; Rubin, D.T.; Kamm, M.A.; Kaakoush, N.O.; Mitchell, H.M.; Castaño-Rodríguez, N. Faecal microbiota transplantation for inflammatory bowel disease: A systematic review and meta-analysis. J. Crohn’s Colitis 2017, 11, 1180–1199.
  138. Zhang, Z.; Mocanu, V.; Cai, C.; Dang, J.; Slater, L.; Deehan, E.C.; Walter, J.; Madsen, K.L. Impact of fecal microbiota transplantation on obesity and metabolic syndrome—A systematic review. Nutrients 2019, 11, 2291.
  139. Kragsnaes, M.S.; Kjeldsen, J.; Horn, H.C.; Munk, H.L.; Pedersen, F.M.; Holt, H.M.; Pedersen, J.K.; Holm, D.K.; Glerup, H.; Andersen, V. Efficacy and safety of faecal microbiota transplantation in patients with psoriatic arthritis: Protocol for a 6-month, double-blind, randomised, placebo-controlled trial. BMJ Open 2018, 8, e019231.
  140. Greenwood, B. The contribution of vaccination to global health: Past, present and future. Philos. Trans. R. Soc. B-Biol. Sci. 2014, 369, 20130433.
  141. Elaish, M.; Ngunjiri, J.M.; Ali, A.; Xia, M.; Ibrahim, M.; Jang, H.; Hiremath, J.; Dhakal, S.; Helmy, Y.A.; Jiang, X.; et al. Supplementation of inactivated influenza vaccine with norovirus P particle-M2e chimeric vaccine enhances protection against heterologous virus challenge in chickens. PLoS ONE 2017, 12, e0171174.
  142. Fawzy, M.; Helmy, Y.A. The One Health Approach is Necessary for the Control of Rift Valley Fever Infections in Egypt: A Comprehensive Review. Viruses 2019, 11, 139.
  143. Hoelzer, K.; Bielke, L.; Blake, D.P.; Cox, E.; Cutting, S.M.; Devriendt, B.; Erlacher-Vindel, E.; Goossens, E.; Karaca, K.; Lemiere, S.; et al. Vaccines as alternatives to antibiotics for food producing animals. Part 2: New approaches and potential solutions. Vet. Res. 2018, 49, 70.
  144. Klugman, K.P.; Black, S. Impact of existing vaccines in reducing antibiotic resistance: Primary and secondary effects. Proc. Natl. Acad. Sci. USA 2018, 115, 12896–12901.
  145. Micoli, F.; Bagnoli, F.; Rappuoli, R.; Serruto, D. The role of vaccines in combatting antimicrobial resistance. Nat. Rev. Microbiol. 2021, 19, 287–302.
  146. Kennedy, D.A.; Read, A.F. Why does drug resistance readily evolve but vaccine resistance does not? Proc. Biol. Sci. 2017, 284, 20162562.
  147. Sihvonen, R.; Siira, L.; Toropainen, M.; Kuusela, P.; Patari-Sampo, A. Streptococcus pneumoniae antimicrobial resistance decreased in the Helsinki Metropolitan Area after routine 10-valent pneumococcal conjugate vaccination of infants in Finland. Eur. J. Clin. Microbiol. Infect. Dis. 2017, 36, 2109–2116.
  148. Klugman, K.P.; Madhi, S.A.; Huebner, R.E.; Kohberger, R.; Mbelle, N.; Pierce, N.; Grp, V.T. A trial of a 9-valent pneumococcal conjugate vaccine in children with and those without HIV infection. N. Engl. J. Med. 2003, 349, 1341–1348.
  149. WHO. Pakistan First Country to Introduce New Typhoid Vaccine into Routine Immunization Programme; WHO: Geneva, Switzerland, 2019.
  150. Kaufhold, S.; Yaesoubi, R.; Pitzer, V.E. Predicting the Impact of Typhoid Conjugate Vaccines on Antimicrobial Resistance. Clin. Infect. Dis. 2019, 68, S96–S104.
  151. House, J.K.; Ontiveros, M.M.; Blackmer, N.M.; Dueger, E.L.; Fitchhorn, J.B.; McArthur, G.R.; Smith, B.P. Evaluation of an autogenous Salmonella bacterin and a modified live Salmonella serotype Choleraesuis vaccine on a commercial dairy farm. Am. J. Vet. Res. 2001, 62, 1897–1902.
  152. Harvey, R.R.; Friedman, C.R.; Crim, S.M.; Judd, M.; Barrett, K.A.; Tolar, B.; Folster, J.P.; Griffin, P.M.; Brown, A.C. Epidemiology of Salmonella enterica Serotype Dublin Infections among Humans, United States, 1968–2013. Emerg. Infect. Dis. 2017, 23, 1493–1501.
  153. Nair, D.V.T.; Venkitanarayanan, K.; Kollanoo, J.A. Antibiotic-Resistant Salmonella in the Food Supply and the Potential Role of Antibiotic Alternatives for Control. Foods 2018, 7, 167.
  154. Cummings, K.J.; Rodriguez-Rivera, L.D.; Capel, M.B.; Rankin, S.C.; Nydam, D.V. Short communication: Oral and intranasal administration of a modified-live Salmonella Dublin vaccine in dairy calves: Clinical efficacy and serologic response. J. Dairy Sci. 2019, 102, 3474–3479.
  155. Hosomi, K.; Hinenoya, A.; Suzuki, H.; Nagatake, T.; Nishino, T.; Tojima, Y.; Hirata, S.-I.; Matsunaga, A.; Kondoh, M.; Yamasaki, S.; et al. Development of a bivalent food poisoning vaccine: Augmented antigenicity of the C-terminus of Clostridium perfringens enterotoxin by fusion with the B subunit of Escherichia coli Shiga toxin 2. Int. Immunol. 2019, 31, 91–100.
  156. Taha-Abdelaziz, K.; Alkie, T.N.; Hodgins, D.C.; Yitbarek, A.; Shojadoost, B.; Sharif, S. Gene expression profiling of chicken cecal tonsils and ileum following oral exposure to soluble and PLGA-encapsulated CpG ODN, and lysate of Campylobacter jejuni. Vet. Microbiol. 2017, 212, 67–74.
  157. Taha-Abdelaziz, K.; Yitbarek, A.; Alkie, T.N.; Hodgins, D.C.; Read, L.R.; Weese, J.S.; Sharif, S. PLGA-encapsulated CpG ODN and Campylobacter jejuni lysate modulate cecal microbiota composition in broiler chickens experimentally challenged with C. jejuni. Sci. Rep. 2018, 8, 12076.
  158. Alizadeh, M.; Shojadoost, B.; Boodhoo, N.; Astill, J.; Taha-Abdelaziz, K.; Hodgins, D.C.; Kulkarni, R.R.; Sharif, S. Necrotic enteritis in chickens: A review of pathogenesis, immune responses and prevention, focusing on probiotics and vaccination. Anim. Health Res. Rev. 2021, 22, 147–162.
  159. Taha-Abdelaziz, K.; Hodgins, D.C.; Alkie, T.N.; Quinteiro-Filho, W.; Yitbarek, A.; Astill, J.; Sharif, S. Oral administration of PLGA-encapsulated CpG ODN and Campylobacter jejuni lysate reduces cecal colonization by Campylobacter jejuni in chickens. Vaccine 2018, 36, 388–394.
  160. Taha-Abdelaziz, K.; Singh, M.; Sharif, S.; Sharma, S.; Kulkarni, R.R.; Alizadeh, M.; Yitbarek, A.; Helmy, Y.A. Intervention Strategies to Control Campylobacter at Different Stages of the Food Chain. Microorganisms 2023, 11, 113.
  161. Nothaft, H.; Perez-Muñoz, M.E.; Gouveia, G.J.; Duar, R.M.; Wanford, J.J.; Lango-Scholey, L.; Panagos, C.G.; Srithayakumar, V.; Plastow, G.S.; Coros, C.; et al. Coadministration of the Campylobacter jejuni N-Glycan-Based Vaccine with Probiotics Improves Vaccine Performance in Broiler Chickens. Appl. Environ. Microbiol. 2017, 83, e01523-17.
  162. Buchy, P.; Ascioglu, S.; Buisson, Y.; Datta, S.; Nissen, M.; Tambyah, P.A.; Vong, S. Impact of vaccines on antimicrobial resistance. Int. J. Infect. Dis. 2020, 90, 188–196.
  163. Mishra, R.P.; Oviedo-Orta, E.; Prachi, P.; Rappuoli, R.; Bagnoli, F. Vaccines and antibiotic resistance. Curr. Opin. Microbiol. 2012, 15, 596–602.
  164. Poolman, J.T. Expanding the role of bacterial vaccines into life-course vaccination strategies and prevention of antimicrobial-resistant infections. NPJ Vaccines 2020, 5, 84.
  165. 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.
  166. Singh, A.; Chaudhary, S.; Agarwal, A.; Verma, A.S. Antibodies. In Animal Biotechnology; Elsevier: Amsterdam, The Netherlands, 2014; pp. 265–287.
  167. Liddell, E. Antibodies. In The Immunoassay Handbook; Elsevier: Amsterdam, The Netherlands, 2013; pp. 245–265.
  168. van Erp, E.A.; Luytjes, W.; Ferwerda, G.; van Kasteren, P.B. Fc-Mediated Antibody Effector Functions During Respiratory Syncytial Virus Infection and Disease. Front. Immunol. 2019, 10, 548.
  169. DiGiandomenico, A.; Sellman, B.R. Antibacterial monoclonal antibodies: The next generation? Curr. Opin. Microbiol. 2015, 27, 78–85.
  170. McConnell, M.J. Where are we with monoclonal antibodies for multidrug-resistant infections? Drug Discov. Today 2019, 24, 1132–1138.
  171. Cavaco, M.; Castanho, M.A.R.B.; Neves, V. The Use of Antibody-Antibiotic Conjugates to Fight Bacterial Infections. Front. Microbiol. 2022, 13, 835677.
  172. Yamada, T. Therapeutic monoclonal antibodies. Keio J. Med. 2011, 60, 37–46.
  173. Reichert, J.M.; Dewitz, M.C. Ooutlook—Anti-infective monoclonal antibodies: Perils and promise of development. Nat. Rev. Drug Discov. 2006, 5, 191–195.
  174. Chan, C.E.Z.; Chan, A.H.Y.; Hanson, B.J.; Ooi, E.E. The use of antibodies in the treatment of infectious diseases. Singap. Med. J. 2009, 50, 663–672.
  175. Kohler, G.; Milstein, C. Continuous cultures of fused cells secreting antibody of predefined specificity (Reprinted from Nature, vol 256, 1975). J. Immunol. 2005, 174, 2453–2455.
More
This entry is offline, you can click here to edit this entry!
Video Production Service