Next Article in Journal
Carotenoids for Antiaging: Nutraceutical, Pharmaceutical, and Cosmeceutical Applications
Next Article in Special Issue
Hybrid Bis-(Imidazole/Benzimidazole)-Pyridine Derivatives with Antifungal Activity of Potential Interest in Medicine and Agriculture via Improved Efficiency Methods
Previous Article in Journal
Establishing Multi-Dimensional LC-MS Systems for Versatile Workflows to Analyze Therapeutic Antibodies at Different Molecular Levels in Routine Operations
Previous Article in Special Issue
Beyond Conventional Antifungals: Combating Resistance Through Novel Therapeutic Pathways
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Combating Antibiotic Resistance: Mechanisms, Multidrug-Resistant Pathogens, and Novel Therapeutic Approaches: An Updated Review

1
Botany and Microbiology Department, Faculty of Science, Tanta University, Tanta 31527, Egypt
2
Aquaculture Research, Alfred Wegener Institute (AWI)—Helmholtz Centre for Polar and Marine Research, Am Handelshafen, 27570 Bremerhaven, Germany
3
Applied and Analytical Microbiology Department, Faculty of Science, Ain Shams University, Cairo 11566, Egypt
4
Microbiology and Chemistry Department, Faculty of Science, Damanhour University, Damanhour 22514, Egypt
5
Microbiology and Biochemistry Program, Faculty of Science, Benha University-Obour Campus, Benha 13518, Egypt
6
Faculty of Pharmacy, Sinai University, Ismailia 41636, Egypt
7
Faculty of Medicine, Alexandria University, Alexandria 21526, Egypt
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2025, 18(3), 402; https://doi.org/10.3390/ph18030402
Submission received: 10 February 2025 / Revised: 6 March 2025 / Accepted: 10 March 2025 / Published: 12 March 2025

Abstract

:
The escalating global health crisis of antibiotic resistance, driven by the rapid emergence of multidrug-resistant (MDR) bacterial pathogens, necessitates urgent and innovative countermeasures. This review comprehensively examines the diverse mechanisms employed by bacteria to evade antibiotic action, including alterations in cell membrane permeability, efflux pump overexpression, biofilm formation, target site modifications, and the enzymatic degradation of antibiotics. Specific focus is given to membrane transport systems such as ATP-binding cassette (ABC) transporters, resistance–nodulation–division (RND) efflux pumps, major facilitator superfamily (MFS) transporters, multidrug and toxic compound extrusion (MATE) systems, small multidrug resistance (SMR) families, and proteobacterial antimicrobial compound efflux (PACE) families. Additionally, the review explores the global burden of MDR pathogens and evaluates emerging therapeutic strategies, including quorum quenching (QQ), probiotics, postbiotics, synbiotics, antimicrobial peptides (AMPs), stem cell applications, immunotherapy, antibacterial photodynamic therapy (aPDT), and bacteriophage. Furthermore, this review discusses novel antimicrobial agents, such as animal-venom-derived compounds and nanobiotics, as promising alternatives to conventional antibiotics. The interplay between clustered regularly interspaced short palindromic repeats (CRISPR) and CRISPR-associated proteins (Cas) in bacterial adaptive immunity is analyzed, revealing opportunities for targeted genetic interventions. By synthesizing current advancements and emerging strategies, this review underscores the necessity of interdisciplinary collaboration among biomedical scientists, researchers, and the pharmaceutical industry to drive the development of novel antibacterial agents. Ultimately, this comprehensive analysis provides a roadmap for future research, emphasizing the urgent need for sustainable and cooperative approaches to combat antibiotic resistance and safeguard global health.

Graphical Abstract

1. Introduction

Antibiotic resistance has emerged as a grave global public health concern, primarily attributed to the indiscriminate use of antibiotics in healthcare, veterinary care, aquaculture, and agriculture [1,2]. This trend has catalyzed the proliferation of antimicrobial-resistant (AMR) bacteria, challenging our ability to combat infectious diseases [2,3]. Despite efforts to develop new pharmaceutical solutions, infections caused by pathogens like Pseudomonas aeruginosa and multidrug-resistant Enterobacteriaceae have become increasingly impervious to existing antibiotics [4]. The World Health Organization ranks antibiotic resistance among the top 10 threats to global health [5]. This crisis underscores the urgent need for responsible antibiotic stewardship, novel therapeutic strategies, and continued research to curb the spread of AMR. In vitro antimicrobial susceptibility testing has been crucial in assessing resistance levels, with multidrug-resistant microbes emerging as a central concern [5].
Antibiotic resistance is a complex natural process influenced by environmental conditions, microbial community density, and widespread antibiotic use in healthcare, agriculture, and food production [6]. Antibiotic-resistant pathogens spread resistance genes through multiple horizontal gene transfer (HGT) mechanisms, including transformation, conjugation, and transduction. Additionally, outer membrane vesicles (OMVs) serve as vehicles for genetic exchange, further enhancing the dissemination of resistance traits among bacterial populations [7]. The discovery of penicillin in 1928 marked a turning point in drug development, leading to the current annual production of about 100,000 tonnes of antibiotics [8]. However, many bacterial species have evolved resistance to multiple drugs, resulting in multidrug resistance [9]. This evolution from the initial breakthrough of penicillin to the current crisis of widespread resistance highlights the urgent need for innovative strategies to combat antimicrobial resistance and safeguard public health. The adaptive nature of bacteria and the continued pressure from antibiotic use underscore the complexity of this challenge and the importance of developing new approaches to maintain effective treatments for infectious diseases.
Multidrug resistance (MDR) typically refers to bacteria resistant to three or more distinct antibiotic classes [10,11]. The human toll of antimicrobial resistance (AMR) is severe, causing over 700,000 deaths annually worldwide, with projections reaching 10 million by 2050 [5]. The economic impact is equally staggering, with AMR projected to cost USD 100 trillion annually by 2050 [12]. Global estimates indicate that antibiotic resistance poses a severe economic burden, particularly in low- and middle-income countries, due to extended hospital stays, higher treatment costs, and increased mortality. The total global cost of antimicrobial resistance is difficult to quantify precisely, but studies suggest that antibiotic resistance could cause up to 10 million deaths annually by 2050, significantly impacting healthcare expenses worldwide [13]. Additionally, the economic cost per antibiotic consumed varies by drug class and income level, underscoring the need for comprehensive policy interventions [14]. For a more in-depth global analysis, further data integration across multiple healthcare systems is necessary [15]. These alarming statistics underscore the urgent need for decisive action in combating antibiotic resistance to mitigate both the devastating loss of human life and the enormous economic burden on global healthcare systems.
The development of new antibiotics faces significant scientific, economic, and regulatory challenges that further complicate the fight against antimicrobial resistance [16]. Pharmaceutical companies have increasingly withdrawn from antibiotic research due to low financial returns and high development costs, with estimates suggesting that bringing a new antibiotic to market requires approximately USD 1 billion and 10–15 years of research [17,18]. The complex regulatory landscape, stringent safety requirements, and limited market incentives create substantial disincentives for pharmaceutical innovation [19,20]. Moreover, the rapid emergence of bacterial resistance mechanisms means that newly developed antibiotics quickly become less effective, further diminishing the economic viability of antibiotic research [21]. Scientifically, the complexity of bacterial adaptation, coupled with the increasing sophistication of resistance mechanisms, makes discovering novel antimicrobial compounds extremely challenging [22]. Traditional screening methods have yielded diminishing returns, necessitating more advanced approaches such as computational drug design, metagenomic exploration, and interdisciplinary strategies that integrate microbiology, genetics, and advanced computational techniques [23,24].
This comprehensive review examines the complex mechanisms of antibiotic resistance in multidrug-resistant bacteria, focusing on four key processes: limiting drug absorption, altering drug targets, drug inactivation, and active drug efflux. These mechanisms vary based on bacterial cell morphology and structure, affecting both acquired and innate resistance. This review explores strategies to mitigate resistance evolution and enhance antimicrobial efficacy, highlighting promising approaches such as quorum-sensing inhibitors (QSIs) and quorum quenchers (QQs). Novel findings include the susceptibility of some resistant strains to scorpion venom, the potential of DsbA protein blockade, and the efficacy of antibacterial photodynamic therapy against resistant Staphylococcus aureus. This review also discusses innovative approaches beyond traditional pharmacology, including using mesenchymal stem cells, immunotherapeutic cells, antimicrobial peptides, and probiotics, with the latter showing potential in reducing antibiotic resistance genes in the gastrointestinal environment.

2. Types and Resistance Mechanisms of Bacteria to Antibiotics

The resistance mechanisms of bacteria to antibiotics can be comprehensively categorized into three primary types: intrinsic, acquired, and adaptive resistance. Each type represents a distinct strategy employed by bacteria to survive antibiotic exposure.

2.1. Intrinsic Resistance

Intrinsic resistance is a natural and inherent characteristic of bacterial species, rooted in fundamental chromosomal elements. This type of resistance involves structural barriers like lipopolysaccharides (LPS) in Gram-negative bacteria, inherent antimicrobial resistance genes (ARGs), and naturally occurring defense mechanisms. Key features include beta-lactamase enzymes, efflux pumps that expel antimicrobial agents, and structural variations that prevent antibiotic penetration [25].

2.2. Acquired Resistance

Acquired resistance involves the horizontal transfer of genetic material between bacterial species, enabling a dynamic spread of resistance traits. This mechanism allows bacteria to acquire resistance through processes such as conjugation, transformation, and transduction. The potential for cross-species transmission of resistance genes makes this type of resistance particularly concerning, as it can rapidly extend antibiotic resistance beyond individual bacterial populations [25].

2.3. Adaptive Resistance

Adaptive resistance emerges from environmental pressures within microbial niches, characterized by dynamic cellular responses to antibiotic stress. These adaptive mechanisms involve modifications in cell permeability, biofilm formation, and metabolic adaptations that allow bacteria to survive in challenging environments. The ability to quickly modify cellular strategies provides bacteria with a significant survival advantage against antimicrobial treatments [25]. The specific mechanisms driving heightened resistances are multifaceted, including biofilm formation, the alteration of antibiotic binding sites, the production of enzyme-based antibiotic deactivation, reduced membrane permeability, and the active efflux of antimicrobial agents. When these resistance mechanisms converge and interact, they can synergistically contribute to enhanced drug resistance, significantly complicating treatment strategies [25]. Table 1 summarizes and provides a multidimensional view of bacterial resistance mechanisms.

3. Mechanisms of Antibiotic Resistance in Bacteria

Bacterial antibiotic resistance mechanisms include enzymatic modification or the inactivation of the antibiotic, the reduced binding affinity of the antibiotic to its bacterial molecular targets, horizontal gene transfer via plasmids carrying multidrug resistance (MDR)-associated genes between species, changes in the permeability of the bacterial cell surface, the overexpression of efflux pumps capable of recognizing and expelling a wide range of antibiotics with diverse mechanisms of action, and genetic variations such as polymorphisms or insertions in DNA sequences encoding transcriptional regulators (Figure 1) [33].
Efflux pumps are integral membrane transporters that play a critical role in bacterial pathogenesis, metabolism, and multidrug resistance [33]. By actively reducing the intracellular concentrations of external agents, such as antibiotics, disinfectants, and detergents, these pumps prevent such substances from reaching their intended biological targets. As a result, efflux transporters have become promising targets for the development of new inhibitors to combat MDR-associated infectious diseases [34].

3.1. Membrane Transport Systems in Antibiotic Resistance

Membrane transport systems play a crucial role in bacterial antibiotic resistance by actively expelling antimicrobial compounds from bacterial cells, thereby reducing their intracellular concentration to levels below the effective therapeutic threshold. These systems encompass several distinct superfamilies, each characterized by unique structural and functional attributes that contribute to antimicrobial resistance. Based on their sequence homology, substrate specificity, structural features, and energy sources, efflux transporters are categorized into five major superfamilies: the multidrug and toxic compound extrusion (MATE) superfamily, small multidrug resistance (SMR) superfamily, ATP-binding cassette (ABC) superfamily, proteobacterial antimicrobial compound efflux (PACE) superfamily, and resistance nodulation and cell division (RND) superfamily, along with the major facilitator superfamily (MFS) [34] (Figure 2).

3.1.1. ATP-Driven Efflux Pumps (ABC)

ATP-binding cassette (ABC) transporters are a major protein superfamily in bacteria, facilitating both nutrient uptake and xenobiotic efflux. These transporters consist of two transmembrane domains (TMDs) that form the substrate transport pathway and two nucleotide-binding domains (NBDs) that drive efflux via ATP hydrolysis. In Gram-negative bacteria, periplasmic binding proteins often aid in substrate recognition and transport initiation [35].
ABC transporters play a dual role in antibiotic resistance and virulence. Export systems actively pump out antibiotics, as seen in the MacAB-TolC system. A tripartite efflux system in Escherichia coli consists of MacA (membrane fusion protein), MacB (ATP-binding cassette transporter), and TolC (outer membrane channel), which contributes to macrolide resistance and virulence factor secretion in Enterobacteriaceae. Studies in Serratia marcescens have demonstrated that deleting macAB genes increases bacterial susceptibility to aminoglycosides and cationic antimicrobial peptides (CAPs), though with varying effects on different antibiotics [36].
The expression of ABC transporters is tightly regulated by substrate-responsive transcriptional regulators and two-component regulatory systems. Current research aims to develop ABC transporter inhibitors as antibiotic adjuvants, though challenges remain due to the redundancy and complex regulation of bacterial efflux systems [37].

3.1.2. Major Facilitator Superfamily (MFS)

The major facilitator superfamily (MFS) is the largest and most extensively characterized group of transmembrane secondary transport proteins in both prokaryotic and eukaryotic systems. MFS transporters mediate substrate translocation via passive transport mechanisms, including uniport (transport of a single substrate) and secondary active transport mechanisms, such as symport (co-transport of a substrate with a coupling ion in the same direction) and antiport (exchange of substrates in opposite directions). The substrates transported by MFS proteins are remarkably diverse, encompassing sugars, amino acids, metabolic intermediates, ions, and antimicrobials [38].
A chromosomally encoded MFS multidrug efflux pump, designated SA09310 gene, has been identified in Staphylococcus aureus. This 43-kDa protein features 12 transmembrane helices (TMHs), a hallmark structural characteristic of MFS transporters. SA09310 contains conserved amino acid sequence motifs specific to MFS transporters, confirming its classification within this family. Functionally, SA09310 has been implicated in resistance to a broad spectrum of antibiotics, including aminoglycosides, tetracyclines, macrolides, and chloramphenicol. Experimental evidence indicates that SA09310 mediates antibiotic resistance by actively exporting intracellular tetracycline from bacterial cells into the extracellular environment. This efflux mechanism reduces intracellular antibiotic concentrations, thereby preventing the drug from reaching its biological targets [39].
SCO4121, another MFS transporter gene, has been identified in Streptomyces coelicolor and characterized for its role in multidrug resistance (MDR). Overexpression studies of SCO4121, both in native and heterologous host systems, demonstrate its capacity to confer resistance to a variety of antibiotics, including ciprofloxacin and chloramphenicol. Conversely, the deletion of the SCO4121 gene significantly increases bacterial susceptibility to these drugs, highlighting its function as a major efflux transporter. In addition to antibiotic resistance, SCO4121 also confers tolerance to oxidative stress caused by hypochlorous acid (HOCl), a potent oxidizing agent. This dual role underscores the physiological importance of MFS transporters in bacterial survival under diverse environmental stresses. Notably, the expression of SCO4121 is transcriptionally regulated in a concentration-dependent manner by its substrate drugs, suggesting a sophisticated adaptive response mechanism in wild-type S. coelicolor cells [40]. Efflux pumps belonging to the MFS family, such as SA09310 and SCO4121, play a pivotal role in multidrug resistance by reducing intracellular antibiotic concentrations, thus allowing bacteria to survive in the presence of antimicrobial agents. The ability of MFS transporters to efflux diverse substrates highlights their evolutionary adaptation and functional versatility. Consequently, MFS efflux pumps represent promising targets for the development of efflux pump inhibitors (EPIs), which could restore the efficacy of existing antibiotics and help combat MDR in pathogenic bacteria.

3.1.3. Small Multidrug Resistance (SMR) Family

The small multidrug resistance (SMR) family comprises small multidrug-resistant proteins typically ranging from 100 to 140 amino acids in length. These proteins are characterized by their structural organization into four transmembrane α-helices embedded in bacterial membranes. SMR proteins are notable for their ability to confer resistance to a wide range of compounds, particularly quaternary ammonium compounds (QACs) and highly hydrophobic substances, due to their capacity to solubilize these substances in organic solvents.
Functionally, SMR proteins play a critical role in the synthesis and efflux of various lipophilic compounds, including antiseptics, detergents, antibiotics, and other drugs. The SMR family encompasses a diverse group of proteins encoded by genes located on both plasmids and bacterial chromosomes, demonstrating substantial structural and functional heterogeneity. This diversity enables SMR proteins to resist different classes of antibiotics, including β-lactams such as cephalosporins. The genetic basis for this resistance is attributed to the close association between SMR genes and antimicrobial resistance (AMR) genes within bacterial genomes, thereby enhancing the multidrug resistance capabilities of bacterial cells [41].
One noteworthy example of SMR functionality is the A1S_0710 gene, which encodes an SMR transporter protein in Acinetobacter baumannii. The experimental deletion of this gene has revealed a dual role for the encoded protein: its absence results in decreased motility and increased virulence, suggesting that its physiological importance extends beyond antimicrobial resistance [42]. This dual functionality highlights the complex roles SMR proteins play in bacterial survival and pathogenicity.
A prominent member of the SMR family is EmrE, a small multidrug transporter found in Escherichia coli. EmrE operates as a proton-driven efflux pump, utilizing the proton motive force (PMF) across the inner bacterial membrane to expel drug molecules. The stoichiometry of proton-to-drug exchange in EmrE is 2:1, wherein two protons are imported into the cell for each drug molecule exported [43]. The functional mechanism of EmrE revolves around a well-characterized binding pocket at the protein’s core, which includes two highly conserved glutamate residues (Glu14) located near the center of the transmembrane domain. This binding pocket interacts with two ligands simultaneously, facilitating their exchange. Recent structural and functional analyses have provided evidence for the existence of a peripheral ligand binding site, which may modulate the transport process without inducing significant conformational changes in the protein. The binding of substrates to this peripheral site has been hypothesized to influence proton translocation indirectly [43].
In addition to Glu14, emerging studies have highlighted the importance of less-characterized residues such as Histidine 110 (His110). Although distant from the core binding site, His110 may play a regulatory role in SMR transport activity and proton coupling. Its precise contribution remains a subject of ongoing investigation, with recent computational and experimental findings suggesting it influences the conformational dynamics of the protein [44].

3.1.4. Multidrug and Toxic Compound Extrusion (MATE) Family

The MATE family, first identified in 2008, was first described in relation to multidrug resistance in bacteria. One of the earliest characterized members, NorM from Neisseria gonorrhoeae, was identified as a multidrug efflux pump capable of exporting fluoroquinolones and other toxic compounds [45]. MATE transporters are secondary active transporters, primarily involved in the efflux of cationic substrates, and play a critical role in reducing bacterial susceptibility to a range of antimicrobial agents. These include ethidium bromide, berberine, acriflavin, norfloxacin, and tetraphenylphosphonium, all of which are substrates that can accumulate to toxic levels within bacterial cells if not extruded by these pumps [46].
The expression of MATE transporters is tightly regulated, and alterations in their expression profiles can significantly impact bacterial resistance mechanisms. For example, the overexpression of the MepA transporter in Staphylococcus aureus has been shown to confer resistance to tigecycline, an antibiotic frequently used in the treatment of infections caused by methicillin- and vancomycin-resistant S. aureus strains [47]. This overexpression enables the bacterium to actively expel tigecycline, thus reducing the intracellular drug concentration and rendering it ineffective in combating the infection [47]. Despite the structural similarities across the MATE superfamily, different members exhibit unique transport mechanisms that reflect their diverse physiological roles. For instance, NorM-VC, a MATE transporter found in Vibrio cholerae, and NorM-PS, found in Pseudomonas stutzeri, both share comparable structural properties, such as the presence of multiple transmembrane helices forming a central substrate-binding cavity. However, these transporters differ in their energization mechanisms. NorM-PS is driven by H+ electrochemical gradients, where the proton gradient is used to power the extrusion of substrates. In contrast, NorM-VC utilizes both Na+ and H+ gradients to facilitate the transport of substrates, which adds a level of complexity to its transport process and suggests functional divergence within the MATE family [48]. These differences in transport mechanisms reflect the evolutionary adaptability of the MATE family to diverse environmental conditions and substrates. The diverse range of MATE transporters underscores their critical role in bacterial survival, particularly in the context of antibiotic resistance. Targeting these transporters represents a promising strategy for the development of new antimicrobial agents, particularly efflux pump inhibitors (EPIs) that could enhance the efficacy of existing antibiotics by inhibiting MATE-mediated drug extrusion.

3.1.5. Proteobacterial Antimicrobial Compound Efflux (PACE) Family

The proteobacterial antimicrobial compound efflux (PACE) family represents a recently discovered group of bacterial membrane transport proteins that contribute to antimicrobial resistance. First identified in 2013, PACE family transporters have been found predominantly in Proteobacteria, particularly in clinically relevant pathogens such as Acinetobacter baumannii, Klebsiella pneumoniae, and Pseudomonas aeruginosa [49].
PACE family proteins typically consist of about 150–160 amino acid residues and are predicted to have four transmembrane α-helices (TMHs) [50]. Four amino acid residues, glutamic acid, asparagine, alanine, and aspartic acid, are conserved in the family by the alignment of 47 different PACE proteins from various bacterial species. Glutamic acid is found on Transmembrane Helix One. Amino acids are present at the periplasmic membrane boundary of Transmembrane Helix Four, and aspartic acid occurs in the cytoplasmic membrane boundary of Transmembrane Helix Four [49]. These conserved residues are thought to play crucial roles in the protein’s function, and are potentially involved in substrate binding or energy coupling mechanisms. PACE family transporters have been shown to confer resistance to a range of antimicrobial compounds, including chlorhexidine: a widely used biocide in healthcare settings; Acriflavine: an antiseptic agent; Proflavine: another antiseptic compound; and Benzalkonium: a quaternary ammonium compound used as a disinfectant [49]. The prototype member of this family, AceI (Acinetobacter chlorhexidine efflux protein I), was first identified in Acinetobacter baumannii. AceI has been demonstrated to resist chlorhexidine, a crucial antiseptic in clinical settings [51]. While the exact mechanism of PACE transporters remains under investigation, current evidence suggests they function as proton-dependent efflux pumps. This means they likely use the energy from the proton motive force to actively extrude antimicrobial compounds from the bacterial cell, thereby reducing intracellular drug concentrations and conferring resistance [50].
The discovery of PACE family transporters has significant implications for public health and clinical practice. Their ability to confer resistance to commonly used biocides like chlorhexidine poses challenges for infection control in healthcare settings [52]. Some PACE family members have been associated with multidrug-resistant (MDR) phenotypes in clinically important pathogens, complicating treatment options [49]. Understanding the structure and function of PACE transporters could lead to the development of novel inhibitors, potentially restoring the efficacy of certain antimicrobials [50].
As our knowledge of the PACE family expands, it is becoming increasingly clear that these transporters play a significant role in bacterial antimicrobial resistance. Continued research in this area is crucial for developing effective strategies to combat the growing threat of antibiotic-resistant infections.

3.1.6. Resistance Nodulation Cell Division (RND) Superfamily

The resistance nodulation cell division (RND) superfamily is a large group of proteins found in various organisms, including bacteria, archaea, and some eukaryotes. The RND superfamily proteins play a significant role in antibiotic resistance in bacteria. Many clinically relevant antibiotic resistance mechanisms in Gram-negative bacteria involve RND efflux systems, which contribute to multidrug resistance (MDR) by actively extruding a wide range of antibiotics and other antimicrobial agents from the bacterial cell. The RND superfamily comprises 12 transmembrane helices separated by two large loops, forming asymmetric trimers. The outer loop contains binding sites for exported ligands, while the transmembrane domain primarily functions as a channel for protons, providing energy for substrate translocation [53]. These pumps are crucial for bacterial survival, offering a spectrum of functions ranging from pH tolerance to resistance against an array of noxious agents, including toxic natural products, metal ions, bile acids, fatty acids, and host-derived antimicrobial peptides. Due to their diverse roles, including pH tolerance and resistance to various substances such as toxic natural products, metal ions, bile acids, fatty acids, and antimicrobial peptides produced by host cells, RND efflux pumps have garnered significant attention. These pumps are associated with outer membrane proteins (OMPs) and are facilitated by periplasmic adaptor proteins (PAPs). In human-associated bacteria like Salmonella enterica, Escherichia coli, and Pseudomonas aeruginosa, the characteristics of their RND multidrug efflux pumps have been extensively studied compared to other species [54].
In the realm of antimicrobial peptides, CASP (cationic antibiotic-sensitive peptide) emerges as a novel player, demonstrating unique properties in its interaction with bacterial efflux pumps. CASP exhibits a cyclic structure and possesses cationic attributes, allowing it to bind specifically within the periplasmic cleft region of MtrD, a component of the RND pump system. This interaction involves distinct and overlapping amino acid contact sites, distinguishing CASP from other cyclic peptides like colistin and linear cationic antibacterial agents derived from human peptide LL-37 (a human antimicrobial peptide derived from the cathelicidin precursor protein (hCAP18), involved in innate immunity). Although CASP may not sensitize Neisseria gonorrhoeae to novobiocin, an antibiotic substrate for the RND pump, it exhibits efficacy against select rod-shaped Gram-negative bacteria, showcasing its potential as a therapeutic agent with a unique mode of action [55].

3.2. Reduced Membrane Permeability

One of the fundamental mechanisms of antibiotic resistance involves the bacterial ability to reduce membrane permeability, limiting the intracellular concentration of antibiotics. Gram-negative bacteria, in particular, possess an outer membrane that acts as a selective barrier, preventing the entry of harmful compounds [56]. The primary route for hydrophilic antibiotics, such as β-lactams, fluoroquinolones, and aminoglycosides, is through porin proteins: channel-like structures embedded in the outer membrane. The mutations or downregulation of porins can significantly reduce antibiotic uptake, leading to resistance [57].
For instance, Pseudomonas aeruginosa exhibits reduced permeability to carbapenems due to mutations in the OprD porin, which normally facilitates the uptake of imipenem and meropenem [35]. Similarly, Escherichia coli strains resistant to cephalosporins have been found to downregulate OmpF and OmpC porins, leading to decreased drug influx. These changes often occur in conjunction with other resistance mechanisms, such as efflux pump overexpression or β-lactamase production, resulting in multidrug resistance.
Additionally, changes in the lipid composition of bacterial membranes can also alter permeability, affecting the diffusion of antibiotics. Studies on Acinetobacter baumannii indicate that modifications in lipopolysaccharide (LPS) structure can further enhance resistance by reducing membrane fluidity and increasing charge repulsion against cationic antibiotics [37].

3.3. Biofilm Formation

Biofilm formation is a critical survival strategy employed by bacteria to withstand antibiotic treatment and host immune responses. Biofilms are complex three-dimensional microbial communities embedded in a self-produced extracellular polymeric substance (EPS) [58]. The EPS matrix acts as a physical barrier that restricts antibiotic penetration, neutralizes antimicrobial agents, and facilitates the horizontal gene transfer of resistance elements among bacterial cells. Biofilm-associated bacteria exhibit significantly higher tolerance to antibiotics compared to planktonic (free-floating) cells. This resistance can be attributed to the following. Reduced Antibiotic Penetration: The dense biofilm matrix slows down the diffusion of antimicrobial agents, reducing their effectiveness. Altered Metabolic States: Bacteria within biofilms exist in a gradient of metabolic activity, with cells in the deeper layers becoming dormant [4]. These persister cells are highly tolerant to antibiotics that target actively growing bacteria. Quorum Sensing Regulation: Biofilm formation is controlled by quorum sensing, a cell-to-cell signaling mechanism that enhances stress resistance and antibiotic tolerance.
A study on Staphylococcus aureus and Pseudomonas aeruginosa biofilms revealed that exposure to sublethal antibiotic concentrations can trigger an adaptive response that strengthens biofilm defenses [59]. Moreover, bacterial biofilms are commonly associated with chronic infections, such as cystic fibrosis, catheter-related bloodstream infections, and implant-associated infections, making them a major clinical challenge.
Given the role of biofilms in antibiotic resistance, novel therapeutic strategies are being explored, including biofilm-disrupting agents like quorum-sensing inhibitors, EPS-degrading enzymes, and nanoparticle-based drug delivery systems [60].

3.4. Target Site Modifications

Target site modification is a prevalent resistance mechanism where bacteria alter the molecular structures that antibiotics typically bind to, rendering the drugs ineffective. This occurs through genetic mutations, enzymatic modifications, or the acquisition of resistance genes via horizontal gene transfer [37].
Examples of target site modifications include the following. Fluoroquinolone Resistance: Bacteria develop resistance by acquiring mutations in the genes encoding DNA gyrase (gyrA) and topoisomerase IV (parC), the primary targets of fluoroquinolones. These mutations reduce drug binding affinity, leading to decreased susceptibility. β-Lactam Resistance: Changes in penicillin-binding proteins (PBPs) lower the binding affinity of β-lactam antibiotics. Methicillin-resistant Staphylococcus aureus (MRSA) expresses PBP2a, a modified enzyme encoded by the mecA gene, which retains transpeptidase activity even in the presence of β-lactams [37]. Macrolide Resistance: The methylation of 23S rRNA by erm genes prevents macrolide antibiotics, such as erythromycin and azithromycin, from binding to the ribosomal subunit, leading to resistance. Such modifications allow bacteria to continue essential cellular processes despite antibiotic presence. This type of resistance is particularly concerning in Acinetobacter baumannii and Klebsiella pneumoniae, where combined target modifications and other resistance mechanisms contribute to extensive drug resistance [35].

3.5. Enzymatic Degradation of Antibiotics

Bacteria can neutralize antibiotics through enzymatic degradation, a major mechanism contributing to resistance against β-lactams, aminoglycosides, and chloramphenicol. The production of these enzymes often spreads through plasmids and transposable elements, accelerating resistance dissemination.
Key classes of antibiotic-degrading enzymes include the following. β-Lactamases: These enzymes hydrolyze the β-lactam ring, rendering β-lactam antibiotics ineffective. The extended-spectrum β-lactamases (ESBLs) and carbapenemases (e.g., KPC, NDM, OXA-type) are particularly concerning as they confer resistance to a wide range of β-lactam antibiotics [36]. Aminoglycoside-Modifying Enzymes: Bacteria can enzymatically modify aminoglycosides (e.g., gentamicin, tobramycin) through acetylation, phosphorylation, or adenylation, reducing their binding affinity to ribosomal targets. Chloramphenicol Acetyltransferases (CATs): These enzymes inactivate chloramphenicol by acetylation, preventing it from inhibiting bacterial protein synthesis. Enzymatic degradation is a primary driver of resistance in Enterobacteriaceae and Pseudomonas aeruginosa. Given the increasing prevalence of β-lactamase-mediated resistance, combination therapies using β-lactamase inhibitors (e.g., clavulanic acid, avibactam) have been developed to restore β-lactam efficacy. However, the emergence of metallo-β-lactamases (MBLs), which are resistant to current inhibitors, presents an ongoing challenge [61].

4. Strategies to Combat Antibiotic Resistance

Antibiotic resistance remains a critical challenge in global healthcare, threatening the effectiveness of treatments and increasing the burden of resistant infections. While no single approach can completely eliminate this problem, various strategies can help mitigate its spread and impact. Antibiotic overuse is a major driver of resistance evolution, as epidemiological studies have shown a direct link between consumption and the emergence of resistant strains [4]. Resistance genes can transfer between bacterial species via horizontal gene transfer or arise naturally through mutations, which are processes exacerbated by inappropriate prescribing practices and suboptimal dosing [7]. These factors not only enhance pathogen virulence, but also accelerate the dissemination of resistance determinants.
Addressing these challenges requires a multifaceted approach that includes improving antibiotic stewardship, enhancing infection prevention measures, and advancing alternative treatments such as phage therapy and immunotherapies. Public awareness campaigns and education on the prudent use of antibiotics also play a vital role in mitigating resistance. The declining efficacy of conventional antibiotics has necessitated the investigation of alternative antimicrobial approaches. This section examines emerging interventions for combating antibiotic resistance, including quorum quenching mechanisms that disrupt bacterial communication networks, microbial-based therapies, stem cell applications, immunotherapeutic modalities, photodynamic antimicrobial techniques, and CRISPR-Cas systems and their relationship to resistance mechanisms, bacteriophage therapy, and bioactive compounds derived from animal venoms.

4.1. Quorum Quenching (QQ): Targeting Bacterial Communication

Quorum sensing (QS) is a bacterial cell-to-cell communication mechanism that regulates group behaviors, including biofilm formation, virulence factor production, and antibiotic resistance [62]. QS involves the synthesis, accumulation, and recognition of signaling molecules, which trigger changes in gene expression once a critical bacterial density is reached [63]. These coordinated processes enable bacteria to function collectively, enhancing their pathogenicity and resistance [64]. QS systems vary across bacterial types: Gram-negative bacteria rely on acyl-homoserine lactones (AHLs), Gram-positive bacteria use autoinducing peptides, and both groups can utilize autoinducer-2 molecules as a universal signaling system [65,66].
Quorum quenching (QQ) involves the inhibition of quorum sensing (QS) through chemical or enzymatic means, aiming to mitigate behaviors regulated by QS and thereby reduce bacterial pathogenicity (Figure 3). This approach also enhances the efficacy of antibiotic and phage treatments [67]. QS communication can be disrupted through the enzymatic breakdown of Acyl-homoserine lacton (AHL), a signal molecule in Gram-negative bacteria, by lactonases, acylases, and oxidoreductases. Alternatively, it can be achieved by employing small structural compounds that prevent the QS signal molecule from binding to its regulatory protein [68]. The high-density colony population can produce a sufficient amount of small-molecule signals to activate a variety of downstream cellular processes, including virulence and drug resistance mechanisms, and resist antibiotics [69]. Quorum sensing (QS) induces changes in several aspects to minimize bacterial antibiotic resistance. This includes suppressing, deactivating, and disrupting QS signaling molecule production [65]. Bacterial communities continuously produce extracellular polymeric substances (EPS), enhancing the durability of biofilm structures. QS quenching can disrupt the production and maintenance of the EPS matrix, leading to a loss of biofilm structural integrity and increased bacterial drug resistance [70]. Additionally, QS systems can alter the composition of cell membranes and suppress the expression of virulence, efflux pump, and antioxidant genes.
Improved membrane permeability and the suppression of efflux pump mechanisms enhance the ability of antibiotics to penetrate bacterial cells. Natural phytochemical constituents derived from Sambucus nigra and Matricaria chamomilla demonstrate significant quorum quenching (QQ) activity through specific molecular interference with bacterial communication systems [71]. These compounds interact with quorum sensing (QS) machinery by competitively binding to receptor proteins or degrading signaling molecules such as N-acyl homoserine lactones (AHLs) in Gram-negative bacteria and autoinducing peptides (AIPs) in Gram-positive species [72]. For instance, vanillin and trans-cinnamaldehyde from cinnamon bark directly inhibit signal molecule recognition and subsequent virulence gene transcription [73]. Similarly, salicylic acid disrupts the las and rhl QS systems in Pseudomonas aeruginosa by downregulating the production of signaling molecules and interfering with regulatory protein function [74]. These QQ mechanisms ultimately attenuate biofilm formation, virulence factor production, and pathogenicity without imposing selective pressure that would promote resistance development. The molecular specificity of these phytochemicals for QS components presents a promising therapeutic approach that targets bacterial virulence rather than viability, potentially circumventing conventional resistance mechanisms.

4.2. Probiotics, Postbiotics, Prebiotics, and Synbiotics

Microbial resistance to antibiotics has become a critical global health challenge, significantly complicating medical treatments [75]. In light of this issue, strategies that enhance the host’s immune system, particularly through the modulation of the gut and body microbiota, have gained increasing attention. Probiotics, postbiotics, and synbiotics represent promising approaches to restore microbial balance, alleviate infections, and reduce over-reliance on antibiotics [76] (Figure 4) (Table 2). These interventions are integral to maintaining or restoring health by targeting the host’s microbiota, thereby providing alternative therapeutic strategies.
Probiotics, which are live microorganisms, offer health benefits when administered in appropriate amounts. These beneficial bacteria can modulate the gut microbiota, enhance immune responses, and contribute to the overall well-being of the host [77]. Probiotics perform their beneficial effects through various mechanisms, including competitive exclusion, the production of antimicrobial substances, and the modulation of immune responses. They are capable of producing a wide range of bioactive metabolites, such as lactic and acetic acids, ethanol, carbon dioxide, hydrogen peroxide, and bacteriocins, all of which contribute to their antibacterial and immune-modulatory properties [75]. Furthermore, for probiotics to effectively colonize the host and exert beneficial effects, they must be resistant to digestive fluids, bile salts, sodium chloride, and acidic environments, ensuring their survival during gastrointestinal transit [78]. Probiotics must also meet rigorous safety and efficacy standards, supported by clinical evidence, to confirm their health benefits [79].
Postbiotics are the biologically active metabolites or non-viable bacterial products generated by probiotics during fermentation in the host. These metabolites enhance the efficacy of probiotics by exerting their biological effects, such as antimicrobial, anti-inflammatory, and immunomodulatory actions [80]. Postbiotics include a wide range of compounds, including cell wall fragments, proteins, peptides, and organic acids, which have been shown to reduce the risk of infections and improve gut health. Unlike probiotics, postbiotics do not require live microorganisms, making them a more stable and safer alternative in certain clinical settings. Their effectiveness stems from their ability to modulate the immune system, inhibit pathogen growth, and restore microbial balance.
Prebiotics play a crucial role in promoting the growth and activity of beneficial bacteria in the host’s gut. These indigestible food components support microbial health by providing substrates for beneficial microorganisms, thus enhancing their activity and facilitating the production of bioactive metabolites [81].
Synbiotics refer to the combination of prebiotics and probiotics, designed to provide synergistic effects by enhancing the survival and activity of beneficial microorganisms while simultaneously promoting their growth through prebiotic substrates. This combined approach maximizes the therapeutic potential of both probiotics and prebiotics, supporting a healthy gut microbiota, improving nutrient absorption, and enhancing the immune response [82]. Synbiotics are particularly effective in addressing various gastrointestinal and systemic conditions, as they not only replenish beneficial microbes, but also create an environment conducive to their long-term survival and activity [83].
Probiotics and synbiotics offer promising therapeutic alternatives to antibiotics, particularly in addressing chronic conditions such as gastrointestinal disorders, metabolic diseases, and infections associated with antimicrobial resistance. By competing for binding sites, nutrients, and space, probiotics prevent pathogenic bacteria from colonizing the gut, thus reducing the likelihood of infections. Furthermore, probiotics can directly disrupt bacterial biofilms, which are complex microbial communities that contribute to chronic infections and resistance to antibiotics [84]. By interfering with biofilm formation and limiting horizontal gene transfer, probiotics help reduce the spread of antibiotic resistance.
The immune-modulatory effects of probiotics and synbiotics also play a crucial role in maintaining host health. Probiotics stimulate the production of antimicrobial peptides, modulate gut-associated lymphoid tissue (GALT), and promote a balanced inflammatory response, which can help prevent excessive immune activation associated with conditions like inflammatory bowel disease (IBD) and allergies [85].
In summary, probiotics, postbiotics, and synbiotics offer valuable alternatives to traditional antibiotic therapies. These biologically active agents not only enhance immune function and restore microbial balance, but also help mitigate the risks associated with antibiotic resistance. Further research is essential to fully elucidate their mechanisms of action, therapeutic potentials, and specific clinical applications.
Table 2. Some examples of postbiotics, probiotics, and synbiotics and their action against antimicrobial resistance bacteria.
Table 2. Some examples of postbiotics, probiotics, and synbiotics and their action against antimicrobial resistance bacteria.
Prebiotics, Postbiotics, Probiotics, and SynbioticsActionAMR Bacteria References
Prebiotics
Fructooligosaccharides (FOS), Inulin, Galactooligosaccharides (GOS), PolydextroseStimulate beneficial microbial growth, improve gut health, enhance immunity, and reduce pathogen colonizationEscherichia coli, Clostridium difficile, Bacteroides fragilis, Streptococcus mutans[86,87]
Postbiotics
Short-chain fatty acids, enzymes, vitamins, extracellular polysaccharides, cell wall fragments and bacterial lysatesPositively affect immunity, are anti-inflammatory, antibacterial, anti-carcinogenic, and antibiofilmFusobacterium,
Clostridioides difficile, Escherichia coli, Streptococcus mutans, Porphyromonas gingivalis, Tannerella forsythia, Prevotella loescheii, and Salmonella spp.
[88]
Probiotics
Lactobacillus rhamnosus GG and Lactobacillus acidophilus produce bacteriocins such asrhamnosin and lactocinAre anticancer and alleviate intestinal damage, mucositis, and antibiofilmClostridium difficile,
Escherichia coli, and Pseudomonas aeruginosa
[89]
Bifidobacterium
produce bifidocin and lactic acid
Regulate inflammation and antimicrobial peptides, reduce antibiotic overuse, and improve gut healthPseudomonas aeruginosa and Clostridium difficile[90]
Leuconostoc mesenteroides MJM60376 and Leuconostoc mesenteroides LVBH107Antibiofilm and antimicrobial activitiesPorphyromonas gingivalis and Streptococcus mutans KCTC3065[91]
Synbiotic
Bifidobacterium lactis BL-99 with Fructooligosaccharide (FOS)Regulate intestinal microbiotaBilophila, Escherichia, and Shigella[92]
Lactobacillus paracasei VL8 and Mannan oligosaccharide (MOS)Positive role in foodborne pathogensSalmonella Typhimurium[93]
Lactobacillus rhamnosus GG and Fructooligosaccharide
(FOS)
Produces Bacteriocins like rhamnosin
Coordinating gut microbiotaShigella sonnei, Salmonella typhimurium, Klebsiella pneumoniae, and Clostridioides difficile[94]
Streptococcus thermophiles and Xylooligosaccharides (XOS)Positive role in colorectal cancerHelicobacter hepaticus, Helicobacter pylori, Escherichia coli, Enterococcus faecalis, and Streptococcus bovis[95]
Streptococcus thermophiles and Fructooligosaccharide (FOS)Overcoming diarrheaClostridioides difficile[96]

4.3. Stem Cells

Stem cells, particularly mesenchymal stem cells (MSCs), have emerged as promising agents in combating antimicrobial resistance (AMR). These cells possess innate antimicrobial properties and can be sourced from various origins, including induced pluripotent stem cells (iPSCs), adult tissues, fetal membranes, and embryonic stem cells (ESCs). The mechanisms by which MSCs combat antimicrobial resistance are twofold. First, they directly interact with cellular systems to inhibit bacterial growth, and second, they enhance the host’s innate immune system’s capacity to fight pathogens. This is achieved through multiple pathways, including the secretion of antimicrobial peptides and the ability to modify the surrounding cellular environment to impede microbial proliferation. Recent studies highlight MSCs’ role in eliminating AMR bacteria through biofilm degradation and the secretion of AMPs such as β-defensin-2, LL-37, and hepcidin [97] (Figure 5). These peptides disrupt bacterial cell walls and alter the microenvironment to inhibit bacterial growth. Furthermore, MSCs work synergistically with antibiotics to enhance their effectiveness in degrading biofilms, a key factor in bacterial resistance [97].
In wound healing and immune regulation, MSCs promote CXCR4 activity, a receptor essential for hematopoiesis and immune response modulation. CXCR4, a G-protein-coupled receptor (GPCR), interacts with its ligand CXCL12 to mediate immune cell activation. Observations suggest that increased CXCR4 expression enhances the extrafollicular B-cell response and early B-cell activation, further amplifying the antimicrobial effects [98,99]. This interplay of MSCs, AMPs, and CXCR4 forms a multi-pronged approach to tackling AMR by simultaneously boosting host immunity, disrupting bacterial biofilms, and enhancing antibiotic efficacy. The integration of MSC-based therapies with AMPs has shown great potential in developing innovative treatments for AMR infections (Figure 5).
AMPs represent a novel class of therapeutics against bacterial infections, offering an alternative to conventional antibiotics. Often referred to as peptide antibiotics, AMPs are small molecules (7–50 amino acid residues, <10 kDa) that are positively charged (+2 to +9). These peptides are integral to the innate immune system, targeting bacteria, fungi, and parasites [98]. AMPs are primarily stored in polymorphonuclear neutrophils (PMNs) and macrophages (MQ) and are produced by various cell types, including epithelial cells and MSCs. Among the well-studied AMPs is LL-37, identified in 1995 as a cathelicidin peptide derived from human bone marrow [100]. LL-37 exhibits potent antibacterial activity and plays a key role in host defense by promoting angiogenesis and tissue regeneration, attracting and activating mast cells through chemotaxis, stimulating cytokine and antibody production, enhancing keratinocyte survival to prevent cell death, and neutralizing bacterial lipopolysaccharides to mitigate inflammation (Figure 5).
Another notable AMP is hepcidin, synthesized by MSCs and essential for iron homeostasis. Hepcidin reduces intestinal iron absorption and mobilizes stored iron, restricting bacterial access to this vital nutrient during infections [100]. The development of resistance-free AMPs is a critical area of research. Techniques such as functional metagenomics, systematic gene overexpression, and laboratory evolution have been employed to engineer AMPs with reduced susceptibility to bacterial resistance. These modified AMPs are characterized by lower polar and positively charged amino acids and increased hydrophobicity, which enhance their antimicrobial activity while minimizing resistance development [100]. The combination of MSCs and AMPs offers a synergistic approach to combating AMR. MSCs not only produce AMPs like β-defensin, LL-37, and hepcidin, but also act as delivery systems that enhance AMP efficacy in targeted treatments. This dual-action approach has shown promise in treating chronic infections and biofilm-associated diseases.

4.4. Immunotherapeutic Approaches

Even bacteria that are susceptible to antibiotics can evade treatment by employing mechanisms to bypass the immune system. To address this challenge, immunotherapeutic approaches focus on leveraging and enhancing the immune system’s ability to combat bacterial infections. Adaptive lymphocytes, such as T and B cells, and innate immune cells, including innate lymphoid cells (ILCs) and natural killer (NK) cells, play a central role in regulating and eliminating infections. They achieve this by directly killing infected cells and secreting inflammatory molecules that stimulate or reinitiate bactericidal responses from myeloid cells. Importantly, a balance must be maintained to prevent excessive inflammation, which can cause tissue damage. This balance is mediated by pathways such as PD-1/PD-L1, which help to modulate immune activity. For instance, the upregulation of PD-1 in T cells during Mycobacterium tuberculosis infections is essential to avoid immune-mediated tissue damage [101].
Recent studies have explored the immunomodulatory potential of the Tim3-Gal9 axis in antimicrobial immunity. Tim3, a T cell immunoglobulin and mucin domain protein, serves as a negative regulator that suppresses effector TH1-type immune responses. While this suppression helps mitigate inflammation-related tissue damage, the overactivation of Tim3 can lead to T cell exhaustion. Galectin-9 (Gal9), a ligand of Tim3, interacts with Tim3 on TH1 cells, triggering macrophages to enhance their bactericidal activity. This process is mediated through the caspase–1-dependent secretion of IL-1β, a cytokine critical for antimicrobial responses. The coupling of Tim3 and Gal9 has been shown to increase antimicrobial immunity by both stimulating macrophages and balancing immune responses [102]. Another innovative method combines antibacterial sonodynamic therapy with antivirulence immunotherapy by using engineered nanovesicles to capture bacterial toxins and generate reactive oxygen species upon ultrasound activation [103].
Macrophages, a key component of the innate immune system, exhibit remarkable plasticity and can polarize into two distinct phenotypes in response to bacterial infections. The M1 phenotype, or classically activated macrophages, promotes pro-inflammatory responses, producing cytokines that attack pathogens and activate other immune cells. In contrast, the M2 phenotype, or alternatively activated macrophages, is associated with anti-inflammatory activities, tissue repair, and the resolution of inflammation. The therapeutic modulation of macrophage polarization is an emerging strategy in immunotherapy. Encouraging the M1 phenotype is crucial for effective pathogen clearance during the early stages of infection, while shifting macrophage activation toward the M2 phenotype can help resolve inflammation and repair tissue damage [104].

4.5. Antibacterial Photodynamic Therapy

Photodynamic therapy (PDT) emerged serendipitously in the early 1900s, revealing the profound potential of light-activated chemicals to induce cellular destruction [105]. This innovative therapeutic approach represents a sophisticated intersection of photochemistry, molecular biology, and targeted treatment strategies. The fundamental efficacy of PDT hinges on the intricate interaction between three critical components: a non-toxic photosensitizer (PS), specific wavelength light, and molecular oxygen within biological systems. The mechanism of PDT is a complex photochemical process that begins when a photosensitizer is exposed to light of a particular wavelength. Upon light activation, the photosensitizer transitions into an excited triplet state, initiating a cascade of molecular interactions. This exciting state becomes a powerful catalyst, transferring energy to surrounding oxygen molecules and generating highly reactive oxygen species (ROS). These unstable molecular entities, including singlet oxygen and hydroxyl radicals, possess remarkable destructive capabilities, which are capable of oxidizing and damaging critical biomolecules, cellular structures, and even drug-resistant pathogens (Figure 6) (Table 3). The generation of ROS introduces additional chemical dynamics, particularly in the presence of iodide anions. These interactions can produce unstable iodide radicals, forming triiodide and hydrogen peroxide, which further contribute to cellular damage mechanisms. However, the therapeutic potential of PDT is intricately linked to tissue oxygen concentrations, presenting both opportunities and challenges in clinical applications [106].
While PDT demonstrates significant promise, certain limitations currently constrain its widespread clinical implementation. Long absorption latency and extended exposure times can potentially diminish its practical utility. The therapy’s effectiveness is fundamentally determined by three interdependent factors: photosensitizer absorption characteristics, light energy properties, and intracellular oxygen content. Recent research has explored innovative strategies to enhance PDT’s antimicrobial potential, particularly in combating resistant bacterial strains. A compelling approach involves combining PDT with traditional antibiotics, creating synergistic therapeutic protocols. For instance, studies examining the combination of toluidine blue (TB)-based antibacterial photodynamic therapy (aPDT) with gentamicin (GEN) have yielded promising results.
In vitro and in vivo investigations demonstrated that the combined GEN and aPDT treatment could effectively inhibit the growth of both standard and multidrug-resistant Staphylococcus aureus for up to 15 h. This combined approach demonstrated remarkable capabilities in destroying bacterial cell envelopes and disrupting biofilm structures, highlighting the potential of integrated therapeutic strategies. The underlying mechanism of this combined approach relies heavily on ROS-induced oxidative stress. By generating reactive oxygen species through non-toxic photosensitizers and visible light, aPDT can fatally damage bacterial cells. Gentamicin, a traditional aminoglycoside antibiotic, complements this approach by inhibiting microbial ribosomal function, thereby accelerating pathogen elimination [107]. Notably, the oxidative stress generated by aPDT can attack critical bacterial envelope proteins and lipids, compromising membrane transport systems and cellular integrity. This multifaceted assault significantly reduces bacterial virulence and infectivity. The potential clinical implications are substantial, suggesting that such combined therapies could potentially reduce antibiotic dosages required for infection clearance while simultaneously promoting more efficient wound healing [107]. The oxidative mechanisms of PDT extend beyond immediate bacterial destruction. By reducing bacterial colonization, modulating inflammatory factors, and stimulating growth factor expression, these therapies represent a sophisticated approach to managing complex infectious challenges.
As research continues to evolve, photodynamic therapy stands at the forefront of innovative antimicrobial strategies, offering hope in an era increasingly challenged by antibiotic-resistant pathogens. The continued exploration of PDT’s mechanisms, refinement of photosensitizer technologies, and development of targeted therapeutic protocols promise exciting advancements in medical treatment paradigms.
Table 3. Photodynamic therapy: a comprehensive overview of photosensitizers, bacterial targets, and light sources.
Table 3. Photodynamic therapy: a comprehensive overview of photosensitizers, bacterial targets, and light sources.
PhotosensitizerPathogenic BacteriaLight TypeWavelength (NM)Key FindingsReference
Gentamicin Staphylococcus aureusBlue Light630–660Effective against standard and MDR strains[108]
Methylene blueEscherichia coliRed Light660–670Significant membrane damage[109]
Chlorin e6Methicillin-Resistant S. aureusRed Light665–685Enhanced ROS generation[110]
PhthalocyanineVancomycin-Resistant EnterococciFar-Red Light670–690Effective against resistant strains[111]
5-aminolevulinic acidCarbapenem-Resistant Acinetobacter baumanniiRed Light630–635Significant bacterial reduction[112]
Aluminum phthalocyanine chlorideMultidrug-Resistant EnterobacterNear-Infrared670–690Improved tissue penetration[113]

4.6. Relationship Between CRISPR CAS (Adaptive Immunity of Bacteria) and Antibiotic Resistance

CRISPR-Cas (clustered regularly interspaced short palindromic repeats and their associated Cas proteins) are an adaptive immune system found in bacteria that helps protect them from invasive genetic material such as bacteriophages and plasmids [114]. It acts as a natural barrier to horizontal gene transfer, a common mechanism for spreading antibiotic resistance [115]. This suggests that CRISPR-Cas may play a role in limiting the acquisition and dissemination of antibiotic resistance genes among bacterial populations. Moreover, unlike traditional antibiotics, which often lack specificity and harm beneficial bacteria, CRISPR-Cas systems can directly and selectively target antibiotic resistance genes (ARGs) and eliminate pathogenic bacteria without affecting other bacterial species in complex bacterial populations. This is because CRISPR-Cas systems are guided by RNA molecules complementary to specific DNA sequences, allowing for the precise targeting of ARGs (Figure 7) [116].
The specificity of CRISPR-Cas systems makes them a promising tool for combating antibiotic resistance. By targeting and eliminating ARGs, CRISPR-Cas can restore the effectiveness of traditional antibiotics and improve the treatment of bacterial infections. CRISPR-Cas systems can be engineered to target multiple ARGs simultaneously, making them effective against bacteria that have developed resistance to multiple antibiotics [117].
Research is ongoing to develop CRISPR-Cas systems for clinical use. These systems have the potential to revolutionize the treatment of antibiotic-resistant infections and improve global health outcomes.
One of the main factors contributing to the growth of multidrug resistance among bacterial pathogens is the acquisition of antibiotic resistance (ABR) genes through horizontal gene transfer (HGT). CRISPR-Cas, as an adaptive immune system, comprises CRISPR loci and CRISPR-associated (cas) genes present in around 30–40% of bacteria [118] between repeating sequences loci. Serving as a kind of “memory” for the immune system, any sequence that matches this immunological record is cut and destroyed by the spacer sequences’ cognate Cas enzymes. By doing so, the cell can protect itself from bacteriophage attack and prevent the entry of additional and possibly expensive Mobile Genetic Elements (MGEs) [118].
Acinetobacter baumannii utilizes the I-Fb CRISPR-Cas system to impede the spread of antibiotic resistance genes. This CRISPR-Cas system effectively halts the horizontal transfer of genetic elements and is highly conserved. The removal of any component of the CRISPR-Cas system resulted in the decreased outer membrane permeability of AB43 strains, aligning with previous findings on membrane permeability [119].
Recent research has demonstrated a correlation between the presence of the CRISPR-Cas system and a reduction in antibiotic resistance genes, along with a significant decrease in virulence factors. Conversely, the absence of the CRISPR locus was found to be linked to multidrug resistance. Furthermore, an inverse relationship exists between the prevalence of bacterial resistance and the existence of the CRISPR-Cas system [115].
Additionally, targeting the quorum-sensing regulator lasR mRNA has shown promise in reducing bacterial pathogenicity in Pseudomonas aeruginosa, particularly when utilizing the I-F CRISPR-Cas system. Interestingly, the QS synthase gene abaI contains a region that partially matches with spacer 101 of the CRISPR system, specifically, nucleotides 29 to 39 in abaI and its repeats. This suggests a potential mechanism for the CRISPR-mediated regulation of quorum sensing and pathogenicity in P. aeruginosa. Targeting the quorum-sensing regulator lasR mRNA has been shown to reduce bacterial pathogenicity in Pseudomonas aeruginosa, particularly when utilizing the I-F CRISPR-Cas system. Interestingly, the QS synthase gene abaI contains a region that partially matches with spacer 101 of the CRISPR system, specifically, nucleotides 29 to 39 in abaI and its repeats. Eliminating any component of the CRISPR-Cas system renders the strain highly resistant to the majority of the medicines tested. The I-Fb CRISPR-Cas system may also target and degrade abaI (QS synthase) mRNA, thereby reducing biological features and genes associated with drug resistance due to the decreased level of AHLs. Furthermore, Cas3 cleavage activity was found to be crucial in controlling the degradation of abaI mRNA [119].
The number of CRISPR loci was found to be significantly higher in gentamicin-, teicoplanin-, erythromycin-, and tetracycline-susceptible strains. Conversely, a smaller number of CRISPR loci were identified in strains positive for vancomycin (vanA), tetracycline (tetM), macrolides (ermB), aminoglycosides (aac6’-aph(2”)), streptogramins (aadE), and streptothricin (ant(6)), indicating a negative correlation between CRISPR-Cas loci and antibiotic resistance [120].
The majority of the Staphylococcus coagulans isolates that were methicillin-susceptible had Type IIC CRISPR-Cas systems and no genes for antibiotic resistance. This suggests that the Type IIC CRISPR–Cas system in S. coagulans strains may hinder the acquisition of antibiotic-resistant genes (ARG) containing mobile genomic elements [121].
Furthermore, Enterococci isolates exhibited CRISPR-Cas components in 63.5% (54/85) of cases. The study revealed a significant inverse relationship between the presence of CRISPR-Cas elements and the proportion of E. faecalis isolates resistant to antibiotics such as vancomycin, ampicillin, chloramphenicol, erythromycin, rifampin, teicoplanin, tetracycline, imipenem, tigecycline, and trimethoprim-sulfamethoxazole. Moreover, compared to CRISPR-Cas negative E. faecium strains, CRISPR-Cas positive E. faecium isolates exhibited significantly lower rates of resistance to vancomycin, ampicillin, chloramphenicol, erythromycin, teicoplanin, and tetracycline [122].
In E. faecalis, the consensus repeats sequences of the CRISPR1 and CRISPR2 loci are identical, suggesting a functional connection between the two loci. An orphan CRISPR2 locus alone cannot effectively defend the genome against mobile genomic elements; it requires the support of CRISPR1-cas. Interestingly, CRISPR3, rather than CRISPR1, was associated with the absence of antibiotic resistance. The availability of the CRISPR3-mutant T11, which acquired cas9 (cas9 + CRISPR3) for interference, indicates that CRISPR3-cas is active for sequence-specific genome defense. Deleting just two loci can significantly weaken the genome’s ability to protect itself from clinically mobile genomic fragments [120].
Figure 7 illustrates how CRISPR-Cas systems can potentially function as antibiotics. The Cas9 enzyme, which can cut DNA in a specific location guided by an RNA molecule, is expressed along with a guide RNA designed to target a particular DNA sequence. The targeted sequence could be on a plasmid or the main bacterial chromosome. The cutting of the plasmid or chromosome by Cas9 can lead to the degradation of the genetic material, resulting in cell death or loss of antibiotic resistance encoded by genes on the targeted DNA. Therefore, the CRISPR-Cas9 system can selectively degrade DNA in bacterial cells, providing an antibacterial mechanism.

4.7. Bacteriophages and Their Role in Antibiotic Resistance and Sensitivity

Bacteriophages, or phages, are viruses that specifically infect and replicate within bacterial cells. These natural predators of bacteria have garnered significant attention for their potential role in combating antibiotic-resistant infections. Phage therapy and antibiotic resistance phage therapy, the use of bacteriophages to treat bacterial infections, have emerged as a promising alternative to traditional antibiotics, particularly in the face of growing antimicrobial resistance [123]. Phages can selectively target and kill specific bacterial strains, including those resistant to multiple antibiotics. This specificity helps preserve the host’s beneficial gut microbiome, unlike broad-spectrum antibiotics that can disrupt the natural microbial balance [124]. Several studies have highlighted the ability of phages to overcome antibiotic resistance. For example, a review by Dufour et al. [125] discussed the successful use of phage therapy to treat infections caused by multidrug-resistant Pseudomonas aeruginosa, Acinetobacter baumannii, and Klebsiella pneumoniae. The authors noted that phages can employ various mechanisms to combat resistance, such as directly targeting resistant determinants or disrupting resistance pathways. The combination of phages and antibiotics, known as phage-antibiotic synergy (PAS), has emerged as a particularly effective strategy against antibiotic-resistant bacteria [124]. By using phages and antibiotics concurrently, researchers have observed enhanced antimicrobial activity and the ability to overcome resistance mechanisms.
A study by Oechslin et al. [126] demonstrated the potential of PAS in treating acute pneumonia caused by multidrug-resistant P. aeruginosa in a mouse model. The combination of a phage cocktail and the antibiotic ciprofloxacin resulted in a significant reduction in bacterial load and improved survival compared to either treatment alone. The synergistic effects of phages and antibiotics can be attributed to several mechanisms. Phages can sensitize bacteria to antibiotics by disrupting cell membranes, altering gene expression, or inducing the formation of persisted cells that are more susceptible to antibiotics [127]. Conversely, antibiotics can increase phage adsorption to bacterial cells or enhance phage replication, leading to a more effective antimicrobial response [124]. While phages offer promising solutions for combating antibiotic resistance, it is important to consider their potential role in the spread of resistance genes. Phages can act as vectors for horizontal gene transfer, allowing for the dissemination of resistance determinants across bacterial populations [128]. Bacteriophages can acquire resistance genes from their bacterial hosts and, subsequently, transfer these genes to other bacteria through transduction, a process in which phages package and deliver bacterial genetic material [129]. This exchange of genetic material can contribute to the emergence and spread of antibiotic-resistant strains, underscoring the need for careful consideration and monitoring when employing phage therapy.
Bacteriophages have emerged as a promising alternative to traditional antibiotics, with the potential to overcome antibiotic resistance through various mechanisms. The synergistic use of phages and antibiotics has shown particular promise in combating multidrug-resistant infections. However, the role of phages in the dissemination of resistance genes highlights the importance of a balanced and well-informed approach to phage therapy. Continued research and the careful implementation of phage-based interventions will be crucial in harnessing the full potential of these natural bacterial predators in the fight against antimicrobial resistance.

4.8. Animal Venoms

Although animal venom is extremely costly and deadly, it offers many interesting possibilities for the creation of biotherapeutics. The “peptides”, which have demonstrated a wide range of biological activities, potential site particularity, and involvement in regulating biological mechanisms, are one of the main ingredients of scorpion venom. In particular, these peptides have drawn increased interest in developing antibiotic-resistant solutions [130]. Animal venoms, which have been utilized for decades in traditional medicine to treat infections and viruses, are one example of the many bioactive substances found in nature [131]. Antimicrobial peptides (AMP) are a promising option for novel therapeutic alternatives in these kinds of instances. AMPs, which are antimicrobial agents with an average length of less than 100 amino acids, are derived from the venoms of several animals, including scorpions, bees, spiders, and snakes [132]. AMPs are short cationic amphipathic peptides found in venoms. Based on their structure, they are categorized into three groups: (1) peptides with cysteine residues and disulfide bridges; (2) peptides without cysteine residues that include members with amphipathic α-helix; and (3) peptides with proline and glycine residues that are abundant in the second group. Three to four disulfide bridges combine to generate cysteine-containing AMPs [133].

4.8.1. Scorpion Venom

Scorpion venom has shown promise in combating antimicrobial resistance (AMR) in bacteria. The venom contains peptides known as α-toxins and β-toxins that disrupt vital bacterial processes, making them effective against multidrug-resistant (MDR) pathogens.
α-Toxins, or sodium channel toxins, can modulate the gating properties of voltage-gated sodium channels in bacterial cell membranes, leading to membrane depolarization and the disruption of essential cellular processes. Several studies have demonstrated the antimicrobial activity of α-toxins against MDR bacteria, including methicillin-resistant Staphylococcus aureus (MRSA) and extended-spectrum beta-lactamase (ESBL)-producing Gram-negative bacteria [134].
Similarly, β-toxins, or potassium channel toxins, target bacterial potassium channels, interfering with membrane potential regulation and metabolic pathways. Recent research has shown the efficacy of β-toxins against MDR bacteria, such as Acinetobacter baumannii, Klebsiella pneumoniae, and Pseudomonas aeruginosa [134]. Interestingly, the combination of α-toxins and β-toxins can produce synergistic effects, enhancing their antimicrobial potency against MDR bacteria by simultaneously disrupting multiple cellular processes [134].
Furthermore, scorpion venom peptides have demonstrated insecticidal activities, with Meucin-49 from Mesobuthus eupeus exhibiting broad-spectrum action against both Gram-positive and Gram-negative bacteria, as well as activity against the bacterial symbionts of pea aphids (Acyrthosiphon pisum), a serious agricultural pest [135]. Additionally, peptides from the venom of the Australian scorpion Urodacus yaschenkoi, such as UyCT1, UyCT3, and UyCT5, have been shown to effectively inhibit Acinetobacter baumannii and other MDR infections [136]. The BmKn-22 peptide, a modified version of the parental BmKn-2 scorpion venom peptide, has also demonstrated the ability to inhibit the development of and destroy pre-formed Pseudomonas aeruginosa biofilms [137]. Overall, the unique mechanisms of action and the potential for synergistic effects make scorpion-venom-derived compounds a promising avenue for combating antimicrobial resistance.

4.8.2. Bee Venom

One of the many bee products high in bioactive compounds is bee venom, a naturally occurring material with a range of actions against various disease etiology [138]. Bee venom is a complex mixture of bioactive compounds, including peptides, enzymes, and other small molecules. Among these components, melittin and apamin peptides have gained significant attention due to their potential antimicrobial properties and ability to modulate immune responses [139] (Table 4). Bv and its main constituents, PLA2 and melittin, were used to combat oral infections found to be the root causes of tooth decay. Between 20 and 40 µg/mL is the minimum inhibitory concentration (MIC) of the BV against Lactobacillus casei, Enterococcus faecalis, Streptococcus mitis, Streptococcus sanguinis, Streotococcus sobrinus, and Streptococcus salivarius [140].
Bee venom has been shown to possess significant antibacterial activity against a range of multidrug-resistant bacteria. Hegazi et al. also demonstrated the antibacterial activity of bee venom against various pathogenic bacterial strains [141]. Fadl further confirmed this, showing that bee venom has a strong potential effect against MDR isolates, including Gram-negative and Gram-positive bacteria. Another study highlighted the potential of melittin, a component of bee venom, as an antimicrobial agent, particularly in treating MRSA infections [142]. These studies collectively suggest that bee venom and its components could be valuable in the fight against multidrug-resistant bacteria [143].

4.8.3. Snake Venom

Snake venom contains a variety of antimicrobial proteins and peptides, including defensins and cathelicidins, which have shown potential for developing alternative antimicrobial agents. Defensins are small cationic peptides active against a broad spectrum of bacteria, including Gram-positive and Gram-negative bacteria. They are typically 30–45 amino acids in length and have a conserved cysteine-rich motif [144]. Cathelicidins are another type of antimicrobial peptide produced by various organisms, including snakes. They are typically larger than defensins, ranging from 12 to 80 amino acids in length, and have a more variable structure, but contain a conserved cathelin domain [145].
These antimicrobial peptides have been observed in various snake species, including the common night adder, gaboon adder, and black mamba, among others [146]. One important class of AMPs found in snake venoms is cathelicidins, such as cathelicidin-NA from the Chinese cobra and cathelicidin-BF from the Bungarus fasciatus snake, which exhibit broad-spectrum antimicrobial activity against bacteria, fungi, and viruses [132].
Additionally, crotamines, cationic peptides isolated from the venom of the South American rattlesnake Crotalus durissus terrificus, have demonstrated potent antimicrobial activity against various bacterial strains, including multidrug-resistant pathogens [147]. Crotamines have also been found to inhibit the development of Trypanosoma cruzi, the causative agent of Chagas disease [148]. Other components of snake venom, such as phospholipase A2 (PLA2) enzymes and L-amino acid oxidases (LAAOs), have also exhibited antimicrobial potential. PLA2s can disrupt bacterial cell membranes, while LAAOs can generate oxidative stress and damage bacterial cells [149,150].
Furthermore, snake venom metalloproteases, like zinc-dependent metalloproteinases (SVMPs), have shown activity against Gram-positive and Gram-negative bacteria, including multidrug-resistant strains, by disrupting cell membranes and degrading essential proteins [151]. Overall, the diverse array of antimicrobial compounds found in snake venom represents a promising avenue for the development of novel therapeutic agents to combat antimicrobial resistance.

4.8.4. Spider Venom

Spiders are known to produce a diverse array of venom peptides, including a family of cationic peptides called insect-selective cystine-knot (ICK) peptides. These peptides have garnered significant interest due to their potential as antimicrobial agents, particularly against multidrug-resistant (MDR) bacteria (Table 4). ICK peptides are characterized by their unique knotted structure, formed by the interlocking of disulfide bridges. This structural motif confers remarkable stability and resistance to proteolytic degradation, making them attractive candidates for therapeutic applications [132]. Several studies have demonstrated the antimicrobial activity of ICK peptides against a range of MDR bacteria, including Gram-positive and Gram-negative pathogens. Here are some examples of ICK peptides from spiders and their potential as antimicrobial agents. Wang et al. [152] isolated and characterized lycosin-II from the venom of the spider Lycosa singoriensis, which displayed potent bacteriostatic effects on drug-resistant bacterial strains. Similarly, Abreu et al. [153] identified gomesin with potential antimicrobial activity in the venom of the Brazilian spider Acanthoscurria gomesiana. These findings suggest that spider venom peptides have the potential to be developed into novel antibiotics. Wu et al. [154] further support this, noting the diverse pharmacological activities of spider venom peptides, including their potential as anticancer and antinociceptive agents.
Table 4. Selected AMPs from animal venom showing their antimicrobial activity.
Table 4. Selected AMPs from animal venom showing their antimicrobial activity.
Venom SpeciesPeptides’ NameAmino Acid ResiduesAntimicrobial ActivityMICReference
Scorpion (Hoffmannihadrurus aztecus) Hadurin41 amino acid-long AMPAntimicrobial activity was mainly detected against Escherichia coli, Serratia marscencens, and Enterococcus cloacae Lower than 10 µm[155]
Scorpion (Pandinus imperator) Pandinin-144 amino acidsAntimicrobial activity was mainly detected against Enterococcus faecalis, Bacillus subtilis, Staphylococcus aureus, and Staphylococcus epidermidis1.3 µm, 5.2 µm, 2.6 µm, and 5.2 µm, respectively, according to those species[156]
Scorpion (Pandinus imperator)Pandinin-2, 24 amino acid residuesAntimicrobial activity was mainly detected against Enterococcus faecalis, Bacillus subtilis, Staphylococcus aureus, and Staphylococcus epidermidis strains and Mycobacterium tuberculosis2.4 µm, 4.8 µm, 2.4 µm and 4.8 µm, respectively, according to those species[157]
Scorpion (Mesobuthus martensii)Bmkbpp47 amino acid residuesAntimicrobial activity was mainly detected against Gram-negative bacteria2.3 to 68.2 µm for different
strains
[158]
Scorpion (Vaejovis punctatus)Vpamp1.0 and vpamp2.019 to 25 amino acid residuesAntimicrobial activity was mainly detected against Gram-positive and Gram-negative bacteria2.5 to 15 µm and 2.5 to 24 µm[159]
Scorpion (Tityus serrulatus)Tsap-217 amino acid residuesAntimicrobial activity was mainly detected against Staphylococcus aureus5 µm[160]
Scorpion (Scorpiops tibetanus)CT214 amino acid residuesInhibits mainly Gram-positive bacteria, especially Staphylococcus aureus
Effective against methicillin-resistant bacterial strains
6.25 μg/mL[161]
Bee (Apis mellifera)Melittin26 amino acidsBroad-spectrum antimicrobial activity against Gram-positive and Gram-negative bacteria, as well as fungi0.5–100 μg/mL against various bacteria[162]
Bee (Apis mellifera)Apidaecin18 amino acidsPotent activity against Gram-negative bacteria, particularly Escherichia coli and Salmonella spp.1.25–5 μg/mL against E. coli[163]
Snake (Bothrops atrox)Cathelicidin-NA34 amino acidsAntimicrobial activity against Gram-positive and Gram-negative bacteria, as well as fungi0.25–128 µg/mL[164]
Snake (Naja atra)Cathelicidin-NA34 amino acidsBacillus anthracis0.29 µg/mL[165]
Snake (Ophiophagus hannah)Cathelicidin-NA-P. Aeruginosa3.25 µm[166]
Snake (Crotalus durissus terrificus)Crotamine42 amino acidsCitrobacter freundii, B. Subtilis, and Micrococcus luteus-[144]
Spider (Acanthoscurria paulensis)Gomesin18 amino acidsPotent antimicrobial activity against Gram-positive and Gram-negative bacteria, as well as fungi0.8–6.4 μm against various bacteria[153]
Spider (Cupiennius salei)Cupiennin 1a35 amino acidsAntimicrobial activity against Gram-positive and Gram-negative bacteria, as well as fungi0.5–4 μm against various bacteria[167]
Spider (Psalmopoeus cambridgei)Psalmopeotoxin I 28 amino acidsAntimicrobial activity against Gram-positive and Gram-negative bacteria, as well as fungi0.4–12.5 μm against various bacteria[168]
In the future, a promising strategy to mitigate the cytotoxic effects of traditional antibiotics on eukaryotic cells involves combining low quantities of fast-killing scorpion antimicrobial peptides (AMPs) with traditional antibiotics. This innovative approach aims to address the challenge of multidrug-resistant (MDR) infections by leveraging the unique properties of scorpion-derived peptides. By utilizing AMPs from animal venom, which are known for their potent antimicrobial activity against a wide range of pathogens, alongside conventional antibiotics, it may be possible to achieve synergistic effects that enhance the efficacy of treatment while minimizing adverse effects on host cells.

4.8.5. Potential Side Effects of Venom-Based Antimicrobial Agents

While venom-derived compounds show promise in combating antibiotic resistance, their potential side effects on humans must be considered. Many venom peptides exhibit cytotoxicity toward mammalian cells, which can lead to hemolysis, neurotoxicity, or immunogenic responses. Bee venom, for instance, has shown promising antimicrobial effects, but may cause allergic reactions and cytotoxicity in high doses [140]. Similarly, lionfish venom contains bioactive peptides with antimicrobial properties, but its toxic side effects, including hemolysis and pain, remain concerns [169]. Additionally, spider venom peptides demonstrate antibacterial and anti-inflammatory effects, but require further investigation into their safety profiles [170]. Therefore, while venoms hold the potential for combating antibiotic resistance, their adverse effects must be carefully evaluated for clinical applications.
To mitigate toxicity risks, researchers are employing peptide modifications to enhance selectivity toward bacterial membranes while minimizing harm to human cells. Chemical modifications, such as amino acid substitutions and structural optimization, have been shown to improve antimicrobial activity while reducing cytotoxicity [171]. Additionally, nanocarrier-based delivery systems, including self-assembled peptide nanostructures, have been developed to control drug release and minimize off-target effects [172]. These approaches collectively improve the clinical viability of antimicrobial peptides while addressing safety concerns. Additionally, the successful development of venom-derived drugs like Captopril and Ziconotide demonstrates that, with rigorous clinical testing, venoms can be safely harnessed for therapeutic applications [151]. Further studies are essential to evaluate the clinical safety profile of venom-based antibiotics before their therapeutic application.

4.9. Nanobiotics

Nanoparticles (NPs) have emerged as a promising platform for the controlled delivery of antibiotics and other therapeutic agents due to their unique properties and versatility. NPs possess several advantages, including their ability to encapsulate and protect drugs from degradation, enhance solubility and bioavailability, and target specific sites or cells within the body. Additionally, some NPs exhibit inherent antimicrobial properties, making them valuable tools in combating bacterial infections. Nanobiotics offer a promising approach to combat antimicrobial resistance (AMR) by providing several advantages over conventional antibiotics, which are summarized in Table 5.
There are four main categories of NPs: inorganic (metallic NPs, silica NPs, and Mesoporous silica NPs), organic (polymeric NPs, liposomes, and dendrimers), carbon-based (carbon nanotubes, Graphene NPs, and Fullerenes), and composite (metal, polymer, and ceramic) [177]. The efficacy of these NPs in drug delivery and antimicrobial applications depends on various factors, including their size, shape, surface chemistry, drug loading capacity, release kinetics, and targeting specificity. Additionally, the inherent antimicrobial properties of certain NPs, such as silver and copper, can be exploited to enhance their therapeutic potential against bacterial infections. Nanobiotics lead to an increase in the medications’ bioavailability in the intended tissue, preventing drug deterioration and reducing the harmful effects on patients. Antibiotics can increase their antibacterial effectiveness by loading them into nanocarriers, according to several studies. When delivering novel antibacterial modalities to bacteria, “nanobiotics” refer to the use of intentionally generated pure antibiotics with a size range of ≤100 nm or antibiotic molecules enclosed with man-made nanoparticles (NPs). These techniques present special chances to fight infections caused by planktonic and multidrug-resistant biofilms [178].
The chemical composition of antibiotic-loaded nanocarriers may raise concerns about immunological action, cytotoxicity, biocompatibility, and biodegradability, even if they may minimize the dosage needed to eliminate bacterial resistance [178].
Compared to conventional antibiotics, antimicrobial nanoparticles (NPs) provide numerous exceptional advantages in overcoming resistance and reducing expenses [179]. In order to improve the pharmacokinetic characteristics of antibiotics and reduce their adverse effects, a variety of nanosized drug carriers are currently available [180]. The utilization of antibacterial nanoscale materials holds significant potential in addressing life-threatening infections. These materials are strategically engineered to enhance biofilm penetration and disrupt bacterial systems effectively. To safely manage superficial infections and infectious disorders, the preference lies in employing metals at the nanoscale, with metal oxides, organic nanoparticles, and nanocomposites all possessing robust antibacterial properties. The diverse chemical compositions and inherent characteristics of these antibacterial nanomaterials, commonly referred to as nanobiotics, offer versatile strategies for combating targeted bacteria [181]. When compared to using antibiotics in bulk, “antibiotic nanocarriers” based on liposomal, solid/lipid, terpenoid, polymeric, dendrimeric, and inorganic materials have demonstrated positive results in improving the total efficacy of antibiotics [182]. NPs can either actively or passively target disease areas. Because infected locations have higher nanocarrier permeability than uninfected tissue, on-target accumulation results are improved. It is also possible to functionalize ligands (such as antibodies) on nanocarrier surfaces that bind to sick tissues or microbes as receptors (such as antigens). Active targeting or ligand-mediated targeting refers to the latter strategy.
Additionally, ligand-conjugated nanocarriers have the potential to enhance intracellular infection treatment by improving cell absorption. Targeted drug delivery systems have advanced quickly in recent years, primarily in the treatment of cancer and tuberculosis caused by Mycobacterium tuberculosis [174,183]. The conjugation was more effective than the antibiotic in its free form. Most importantly, the targeted therapy seems to affect intracellular bacteria that the antibiotic would have missed and left “hidden,” acting as a latent cause of recurrent sickness [184].
Because inorganic NPs are composed of the inorganic oxides of Ag, Mn, Al, Ti, Se, Au, Si, or Cu, their size, shape, solubility, and stability vary. Their antibacterial efficacy is also influenced by aggregation behavior, pH, temperature, reduction time, and reducing agent concentration, all of which determine their properties. To impede respiration, including bacterial cell lysis and inducing inflammatory responses, AgNPs, for instance, adhere to cell membranes, interact with membrane proteins, raise membrane porosity, and enter and promote ROS formation [184]. Since ancient times, silver has demonstrated immense promise for use as an antibacterial and antiseptic agent. Because of the antibacterial effect of bacterial AgNPs, there may be a special way to minimize the development of antibiotic resistance by using fewer drugs overall [185]. AgNPs act as a delivery system for ampicillin (amp) antibiotic with a spherical shape and 4 nm and coated with citrate that can load 1.06 × 10−6 of the amp to target Pseudomonas aeruginosa, Escherichia coli, and Vibrio cholerae to target their cell walls and β-Lactamase species [186].
Because of strong electrostatic forces that cause intracellular loss and cell death, gold nanoparticles (AuNPs) are known to have bactericidal effects [187]. AuNPs accumulate on the surface of cells and have bactericidal effects. The antibacterial properties of Au nanocrystals are facet-dependent and include bacterial membrane damage, the suppression of cellular enzyme activity, and energy metabolism [188]. AuNPs with 4–5 nm and spherical size represent an example of delivering the vancomycin (Van) antibiotic. The phenyl group of Van attaches to AuNP via the Au-S bond, and each AuNP links with approximately 31 Van on its surface to target Escherichia coli and Vancomycin-resistant Enterococci, and targets their cell membranes [189].
For instance, iron oxide NPs (FeONPs) and DNA hybridization are combined to increase the bacterial 16S ribosomal RNA gene capture [190] due to strong electrostatic forces, cytoplasmic leakage, cell death, and gold nanoparticles. Because CaF2 NPs stick to tooth surfaces and release fluoride ions continuously, they have been demonstrated to have fatal effects on Streptococcus mutans. This promotes remineralization and inhibits pathogenic S. mutans [191]. Because of their hydrophilic/hydrophobic individualities, most organic NPs, including liposomes, polymeric materials, and micelles, are biocompatible and rapidly opsonize [181].
Liposomes and lipid nanoparticles are phospholipid bilayers in spherical vesicle form. The medication can be released into bacteria when they fuse with the microbial membrane. They have been transformed into a medication delivery system to combat illnesses caused by biofilms using antimicrobial medicines. By reducing recurrent infections, their special qualities—such as target specificity, low toxicity, and the capacity to fuse biofilm matrix/cell membrane—improve the effectiveness of antibiotics [192].
Drug carriers are polymeric nanoparticles (NPs), such as nanospheres and nanocapsules. They improve the effectiveness of targeting and are chemically and physically stable [193]. One such medication delivery method that shields antibiotics from deterioration and can attach to biofilm constituents is lipid-based surface-functionalized PLGA [194].
Quantum dots (QDs) that are photoexcited are metallic in appearance, but they have a type of semiconductor core made of zinc or cadmium. The capacity of QDs to break down bacterial cell walls or membranes, produce free radicals, bind with genetic material, and prevent energy synthesis is what gives them their antibacterial effect. The antibacterial activity of transferrin-modified silver QDs combined with zinc and rifampicin is significantly higher than that of the zinc and rifampicin complexes [195]. The redox potential of photogenerated charge carriers that interact with the bacterial environment has been found to suppress the growth of multiple-drug-resistant (MDR) clinical isolates (S. typhimurium, methicillin-resistant Staphylococcus aureus, Klebsiella pneumoniae, and carbapenem-resistant Escherichia coli) via photoexcited QDs [196]. Factors such as the poor targeting of antibiotics to infection sites, increased dosing frequencies and side effects, the spread of resistance to currently used antibiotic medicines, the slow development rate of newer antibacterials, and the possibility of resistance to future new antimicrobial drugs all highlight the need to follow novel approaches for managing microbial infections [197].

5. Potential Limitations of Combatting Antibiotic Resistance Strategies

Despite the promising potential of the strategies outlined in Section 4.1, Section 4.2, Section 4.3, Section 4.4, Section 4.5, Section 4.6, Section 4.7, Section 4.8 and Section 4.9, several critical limitations impede their clinical implementation and widespread adoption. Many of these approaches remain predominantly in preclinical or early clinical investigation phases, with insufficient validation through randomized controlled trials and longitudinal studies [198]. The evolutionary adaptability of bacterial pathogens presents another significant concern, as selective pressures may drive the development of resistance mechanisms even against these novel therapies. For instance, bacteria could potentially evolve modified quorum-sensing receptors, bacteriophage resistance, or mechanisms to neutralize antimicrobial peptides from animal venoms [199]. Economic considerations further constrain practical applications, as the high costs associated with the production, purification, and quality control of biologics-based therapies (e.g., bacteriophages, stem cells, and venom-derived peptides) make them potentially inaccessible in resource-limited settings [200]. Additionally, regulatory frameworks remain underdeveloped for many of these novel approaches, creating uncertainty regarding safety assessment, standardization protocols, and approval pathways [201]. These multifaceted challenges necessitate concurrent advances in regulatory science, cost-effective manufacturing processes, and innovative clinical trial designs to facilitate the translation of promising laboratory findings into clinically viable alternatives to conventional antibiotics.

6. Conclusions

The antibiotic resistance crisis, fueled by the overuse of antibiotics and stagnation in new drug development, remains a critical global health challenge. This review delved into the complex resistance mechanisms of multidrug-resistant pathogens and highlighted alternative strategies, including quorum-sensing inhibitors, probiotics, antimicrobial peptides, venoms, nanobiotics, bacteriophages, CRISPR-Cas systems, immunotherapy, and photodynamic therapy. The significance of the environment as both a reservoir for resistance genes and a source of novel antimicrobials was also underscored. Despite these promising alternatives, significant challenges remain. The high costs associated with developing and commercializing new therapies, the need for rigorous regulatory approvals, and the potential for resistance evolution against novel treatments pose major hurdles. Moreover, the complexity of microbial ecosystems necessitates further research to understand the long-term efficacy and ecological impact of these interventions. Future efforts should focus on enhancing the scalability and affordability of novel antimicrobials, optimizing combination therapies to prevent resistance, and fostering global collaborations to implement sustainable antibiotic stewardship programs.
Addressing this crisis demands a unified interdisciplinary effort. Collaboration among healthcare providers, researchers, policymakers, and the public is essential to develop innovative solutions, promote awareness, and implement sustainable strategies. By advancing these approaches, we can safeguard the efficacy of existing antibiotics, ensure the development of effective treatments, and protect public health for current and future generations.

Author Contributions

M.E.E., N.K.B., Y.A., A.A.Z., H.H.M., M.M.A., N.A.M., S.M., and A.M.A. collected data and wrote a draft. M.E.E. contributed significantly to the analysis and manuscript review; M.E.E. reviewed the manuscript and performed the analysis with productive discussions. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained in the paper.

Acknowledgments

The authors acknowledge the support of the Open Access publication fund of Alfred-Wegener-Institut Helmholtz-Zentrum für Polar-und Meeresforschung.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhao, X.; Chen, Z.; Yang, T.; Jiang, M.; Wang, J.; Cheng, Z.; Yang, M.; Zhu, J.; Zhang, T.; Li, H.; et al. Glutamine Promotes Antibiotic Uptake to Kill Multidrug-Resistant Uropathogenic Bacteria. Sci. Transl. Med. 2021, 13, eabj0716. [Google Scholar] [CrossRef] [PubMed]
  2. Islam, M.A.; Islam, M.R.; Khan, R.; Amin, M.B.; Rahman, M.; Hossain, M.I.; Ahmed, D.; Asaduzzaman, M.; Riley, L.W. Prevalence, Etiology and Antibiotic Resistance Patterns of Community-Acquired Urinary Tract Infections in Dhaka, Bangladesh. PLoS ONE 2022, 17, e0274423. [Google Scholar] [CrossRef] [PubMed]
  3. Safdar, N.; Zahra, A.; Komal, S.; Ul Ain, Q.; Mazhar, M. The Trend of Antibiotic Resistance In the Era of COVID-19 and Its Trailing. SSRN J. 2023. [Google Scholar] [CrossRef]
  4. El-Sapagh, S.; El-Shenody, R.; Pereira, L.; Elshobary, M. Unveiling the Potential of Algal Extracts as Promising Antibacterial and Antibiofilm Agents against Multidrug-Resistant Pseudomonas aeruginosa: In Vitro and In Silico Studies Including Molecular Docking. Plants 2023, 12, 3324. [Google Scholar] [CrossRef]
  5. Osei Sekyere, J.; Kerdsin, A.; Chopjitt, P.; Wendling, C.C. Editorial: Community Series—Characterization of Mobile Genetic Elements Associated with Acquired Resistance Mechanisms, Volume II. Front. Microbiol. 2023, 14, 1230730. [Google Scholar] [CrossRef]
  6. Günther, T.; Kramer-Schadt, S.; Fuhrmann, M.; Belik, V. Environmental Factors Associated with the Prevalence of ESBL/AmpC-Producing Escherichia coli in Wild Boar (Sus scrofa). Front. Vet. Sci. 2022, 9, 980554. [Google Scholar] [CrossRef]
  7. Lee, A.R.; Park, S.B.; Kim, S.W.; Jung, J.W.; Chun, J.H.; Kim, J.; Kim, Y.R.; Lazarte, J.M.S.; Jang, H.B.; Thompson, K.D.; et al. Membrane Vesicles from Antibiotic-Resistant Staphylococcus aureus Transfer Antibiotic-Resistance to Antibiotic-Susceptible Escherichia coli. J. Appl. Microbiol. 2022, 132, 2746–2759. [Google Scholar] [CrossRef]
  8. Terefinko, D.; Caban, M.; Motyka-Pomagruk, A.; Babinska, W.; Pohl, P.; Jamroz, P.; Cyganowski, P.; Sledz, W.; Lojkowska, E.; Stepnowski, P.; et al. Removal of Clinically Significant Antibiotics from Aqueous Solutions by Applying Unique High-Throughput Continuous-Flow Plasma Pencil and Plasma Brush Systems. Chem. Eng. J. 2023, 452, 139415. [Google Scholar] [CrossRef]
  9. Zhou, Y.; Zhong, Z.; Hu, S.; Wang, J.; Deng, Y.; Li, X.; Chen, X.; Li, X.; Tang, Y.; Li, X.; et al. A Survey of Helicobacter pylori Antibiotic-Resistant Genotypes and Strain Lineages by Whole-Genome Sequencing in China. Antimicrob. Agents Chemother. 2022, 66, e02188-21. [Google Scholar] [CrossRef]
  10. Alam, M.M.; Islam, N.; Hossain Hawlader, M.D.; Ahmed, S.; Wahab, A.; Islam, M.; Uddin, K.R.; Hossain, A. Prevalence of Multidrug Resistance Bacterial Isolates from Infected Wound Patients in Dhaka, Bangladesh: A Cross-Sectional Study. Int. J. Surg. Open 2021, 28, 56–62. [Google Scholar] [CrossRef]
  11. Nocera, F.P.; Ambrosio, M.; Fiorito, F.; Cortese, L.; De Martino, L. On Gram-Positive- and Gram-Negative-Bacteria-Associated Canine and Feline Skin Infections: A 4-Year Retrospective Study of the University Veterinary Microbiology Diagnostic Laboratory of Naples, Italy. Animals 2021, 11, 1603. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, H.; Wang, H.; Yu, X.; Zhou, H.; Li, B.; Chen, G.; Ye, Z.; Wang, Y.; Cui, X.; Zheng, Y.; et al. Impact of Antimicrobial Stewardship Managed by Clinical Pharmacists on Antibiotic Use and Drug Resistance in a Chinese Hospital, 2010–2016: A Retrospective Observational Study. BMJ Open 2019, 9, e026072. [Google Scholar] [CrossRef] [PubMed]
  13. Pulingam, T.; Parumasivam, T.; Gazzali, A.M.; Sulaiman, A.M.; Chee, J.Y.; Lakshmanan, M.; Chin, C.F.; Sudesh, K. Antimicrobial Resistance: Prevalence, Economic Burden, Mechanisms of Resistance and Strategies to Overcome. Eur. J. Pharm. Sci. 2021, 170, 106103. [Google Scholar] [CrossRef] [PubMed]
  14. Shrestha, P.; Cooper, B.S.; Coast, J.; Oppong, R.; Do Thi Thuy, N.; Phodha, T.; Celhay, O.; Guerin, P.J.; Wertheim, H.; Lubell, Y. Enumerating the Economic Cost of Antimicrobial Resistance per Antibiotic Consumed to Inform the Evaluation of Interventions Affecting Their Use. Antimicrob. Resist. Infect. Control 2018, 7, 98. [Google Scholar] [CrossRef]
  15. Li, T.; Wang, Z.; Guo, J.; de la Fuente-Nunez, C.; Wang, J.; Han, B.; Tao, H.; Liu, J.; Wang, X. Bacterial Resistance to Antibacterial Agents: Mechanisms, Control Strategies, and Implications for Global Health. Sci. Total Environ. 2022, 860, 160461. [Google Scholar] [CrossRef]
  16. Hosny, S.; Elshobary, M.E.; El-Sheekh, M.M. Unleashing the Power of Microalgae: A Pioneering Path to Sustainability and Achieving the Sustainable Development Goals. Environ. Sci. Pollut. Res. 2025, 1–31. [Google Scholar] [CrossRef]
  17. Theuretzbacher, U.; Bush, K.; Harbarth, S.; Paul, M.; Rex, J.H.; Tacconelli, E.; Thwaites, G.E. Critical Analysis of Antibacterial Agents in Clinical Development. Nat. Rev. Microbiol. 2020, 18, 286–298. [Google Scholar] [CrossRef]
  18. Outterson, K. Estimating the Appropriate Size of Global Pull Incentives For Antibacterial Medicines: Study Examines Global Antibacterial Pull Incentives. Health Aff. 2021, 40, 1758–1765. [Google Scholar] [CrossRef]
  19. Årdal, C.; Balasegaram, M.; Laxminarayan, R.; McAdams, D.; Outterson, K.; Rex, J.H.; Sumpradit, N. Antibiotic Development—Economic, Regulatory and Societal Challenges. Nat. Rev. Microbiol. 2020, 18, 267–274. [Google Scholar] [CrossRef]
  20. Blind, K. The Overall Impact of Economic, Social and Institutional Regulation on Innovation: An Update. In Handbook of Innovation and Regulation; Edward Elgar Publishing: Northampton, MA, USA, 2023; pp. 230–262. [Google Scholar]
  21. Pierce, J.D.; Shen, Q.; Cintron, S.A.; Hiebert, J.B. Post-COVID-19 Syndrome. Nurs. Res. 2022, 71, 164–174. [Google Scholar] [CrossRef]
  22. Smith, W.P.; Wucher, B.R.; Nadell, C.D.; Foster, K.R. Bacterial Defences: Mechanisms, Evolution and Antimicrobial Resistance. Nat. Rev. Microbiol. 2023, 21, 519–534. [Google Scholar] [CrossRef] [PubMed]
  23. Sakagianni, A.; Koufopoulou, C.; Koufopoulos, P.; Kalantzi, S.; Theodorakis, N.; Nikolaou, M.; Paxinou, E.; Kalles, D.; Verykios, V.S.; Myrianthefs, P. Data-Driven Approaches in Antimicrobial Resistance: Machine Learning Solutions. Antibiotics 2024, 13, 1052. [Google Scholar] [CrossRef] [PubMed]
  24. Abushaheen, M.A.; Muzaheed; Fatani, A.J.; Alosaimi, M.; Mansy, W.; George, M.; Acharya, S.; Rathod, S.; Divakar, D.D.; Jhugroo, C.; et al. Antimicrobial Resistance, Mechanisms and Its Clinical Significance. Disease-a-Month 2020, 66, 100971. [Google Scholar] [CrossRef] [PubMed]
  25. Tavares-Carreon, F.; De Anda-Mora, K.; Rojas-Barrera, I.C.; Andrade, A. Serratia marcescens Antibiotic Resistance Mechanisms of an Opportunistic Pathogen: A Literature Review. PeerJ 2023, 11, e14399. [Google Scholar] [CrossRef]
  26. Lade, H.; Kim, J.-S. Molecular Determinants of β-Lactam Resistance in Methicillin-Resistant Staphylococcus aureus (MRSA): An Updated Review. Antibiotics 2023, 12, 1362. [Google Scholar] [CrossRef]
  27. Pérez-Varela, M.; Tierney, A.R.; Dawson, E.; Hutcheson, A.R.; Tipton, K.A.; Anderson, S.E.; Haldopoulos, M.E.; Song, S.; Tomlinson, B.R.; Shaw, L.N. Stochastic Activation of a Family of TetR Type Transcriptional Regulators Controls Phenotypic Heterogeneity in Acinetobacter baumannii. PNAS Nexus 2022, 1, pgac231. [Google Scholar] [CrossRef]
  28. Esau, V.C.-Z. Structure-Activity Relationship Between Klebsiella pneumoniae β-Lactamase CTX-M-15 and Selected β-Lactam Antibiotics: Evaluating the Binding Site Promiscuity. Master’s Thesis, University of the Witwatersrand, Johannesburg, South Africa, 2023. [Google Scholar]
  29. Miller, W.R.; Murray, B.E.; Rice, L.B.; Arias, C.A. Vancomycin-Resistant Enterococci: Therapeutic Challenges in the 21st Century. Infect. Dis. Clin. 2016, 30, 415–439. [Google Scholar]
  30. Lewis, K. New Approaches to Antimicrobial Discovery. Biochem. Pharmacol. 2017, 134, 87–98. [Google Scholar] [CrossRef]
  31. Singh, R.K.; Prasad, A.; Maurya, J.; Prasad, M. Regulation of Small RNA-Mediated High Temperature Stress Responses in Crop Plants. Plant Cell Rep. 2022, 41, 765–773. [Google Scholar] [CrossRef]
  32. Rodríguez-Núñez, K.; Cortés-Monroy, A.; Serey, M.; Ensari, Y.; Davari, M.D.; Bernal, C.; Martinez, R. Modulating Substrate Specificity of Rhizobium Sp. Histamine Dehydrogenase Through Protein Engineering for Food Quality Applications. Molecules 2023, 28, 3748. [Google Scholar] [CrossRef]
  33. Kakoullis, L.; Papachristodoulou, E.; Chra, P.; Panos, G. Mechanisms of Antibiotic Resistance in Important Gram-Positive and Gram-Negative Pathogens and Novel Antibiotic Solutions. Antibiotics 2021, 10, 415. [Google Scholar] [CrossRef] [PubMed]
  34. De Gaetano, G.V.; Lentini, G.; Famà, A.; Coppolino, F.; Beninati, C. Antimicrobial Resistance: Two-Component Regulatory Systems and Multidrug Efflux Pumps. Antibiotics 2023, 12, 965. [Google Scholar] [CrossRef]
  35. Scoffone, V.C.; Trespidi, G.; Barbieri, G.; Arshad, A.; Israyilova, A.; Buroni, S. The Evolution of Antimicrobial Resistance in Acinetobacter baumannii and New Strategies to Fight It. Antibiotics 2025, 14, 85. [Google Scholar] [CrossRef] [PubMed]
  36. Zakaria, N.F.S.; Yahya, M.F.Z.R.; Jamil, N.M. Multiple Bacterial Strategies to Survive Antibiotic Pressure: A Review. Preprints 2023. [Google Scholar] [CrossRef]
  37. Belay, W.Y.; Getachew, M.; Tegegne, B.A.; Teffera, Z.H.; Dagne, A.; Zeleke, T.K.; Abebe, R.B.; Gedif, A.A.; Fenta, A.; Yirdaw, G.; et al. Mechanism of Antibacterial Resistance, Strategies and Next-Generation Antimicrobials to Contain Antimicrobial Resistance: A Review. Front. Pharmacol. 2024, 15, 1444781. [Google Scholar] [CrossRef]
  38. Henderson, R.K.; Fendler, K.; Poolman, B. Coupling Efficiency of Secondary Active Transporters. Curr. Opin. Biotechnol. 2019, 58, 62–71. [Google Scholar] [CrossRef]
  39. Li, D.; Ge, Y.; Wang, N.; Shi, Y.; Guo, G.; Zou, Q.; Liu, Q. Identification and Characterization of a Novel Major Facilitator Superfamily Efflux Pump, SA09310, Mediating Tetracycline Resistance in Staphylococcus aureus. Antimicrob. Agents Chemother. 2023, 67, e01696-22. [Google Scholar] [CrossRef]
  40. Nag, A.; Mehra, S. A Major Facilitator Superfamily (MFS) Efflux Pump, SCO4121, from Streptomyces coelicolor with Roles in Multidrug Resistance and Oxidative Stress Tolerance and Its Regulation by a MarR Regulator. Appl. Environ. Microbiol. 2021, 87, e02238-20. [Google Scholar] [CrossRef] [PubMed]
  41. Vrancianu, C.O.; Gheorghe, I.; Czobor, I.B.; Chifiriuc, M.C. Antibiotic Resistance Profiles, Molecular Mechanisms and Innovative Treatment Strategies of Acinetobacter baumannii. Microorganisms 2020, 8, 935. [Google Scholar] [CrossRef]
  42. Pérez-Varela, M.; Corral, J.; Aranda, J.; Barbé, J. Roles of Efflux Pumps from Different Superfamilies in the Surface-Associated Motility and Virulence of Acinetobacter baumannii ATCC 17978. Antimicrob. Agents Chemother. 2019, 63, e02190-18. [Google Scholar] [CrossRef]
  43. Thomas, N.E. The Proton/Drug Coupling Mechanism of EmrE; The University of Wisconsin-Madison: Madison, WI, USA, 2021; ISBN 979-8-5442-4690-9. [Google Scholar]
  44. Teng, D.; Voth, G.A. Ligand Binding by the Small Multidrug-Resistant Transporter EmrE. Biophys. J. 2023, 122, 400a. [Google Scholar] [CrossRef]
  45. Guérin, F.; Galimand, M.; Tuambilangana, F.; Courvalin, P.; Cattoir, V. Overexpression of the Novel MATE Fluoroquinolone Efflux Pump FepA in Listeria monocytogenes Is Driven by Inactivation of Its Local Repressor FepR. PLoS ONE 2014, 9, e106340. [Google Scholar] [CrossRef] [PubMed]
  46. Miyauchi, H.; Moriyama, S.; Kusakizako, T.; Kumazaki, K.; Nakane, T.; Yamashita, K.; Hirata, K.; Dohmae, N.; Nishizawa, T.; Ito, K.; et al. Structural Basis for Xenobiotic Extrusion by Eukaryotic MATE Transporter. Nat. Commun. 2017, 8, 1633. [Google Scholar] [CrossRef]
  47. Kim, J.; Cater, R.J.; Choy, B.C.; Mancia, F. Structural Insights into Transporter-Mediated Drug Resistance in Infectious Diseases. J. Mol. Biol. 2021, 433, 167005. [Google Scholar] [CrossRef] [PubMed]
  48. Stephen, J.; Lekshmi, M.; Ammini, P.; Kumar, S.H.; Varela, M.F. Membrane Efflux Pumps of Pathogenic Vibrio Species: Role in Antimicrobial Resistance and Virulence. Microorganisms 2022, 10, 382. [Google Scholar] [CrossRef]
  49. Hassan, K.A.; Liu, Q.; Elbourne, L.D.H.; Ahmad, I.; Sharples, D.; Naidu, V.; Chan, C.L.; Li, L.; Harborne, S.P.D.; Pokhrel, A.; et al. Pacing across the Membrane: The Novel PACE Family of Efflux Pumps Is Widespread in Gram-Negative Pathogens. Res. Microbiol. 2018, 169, 450–454. [Google Scholar] [CrossRef]
  50. Javed, W. Study of the Conformational States of a Bacterial Multidrug ABC Transporter BmrA. Ph.D. Thesis, Université Grenoble Alpes, Saint-Martin-d’Hères, France, 2020. [Google Scholar]
  51. Migliaccio, A.; Esposito, E.P.; Bagattini, M.; Berisio, R.; Triassi, M.; De Gregorio, E.; Zarrilli, R. Inhibition of AdeB, AceI, and AmvA Efflux Pumps Restores Chlorhexidine and Benzalkonium Susceptibility in Acinetobacter baumannii ATCC 19606. Front. Microbiol. 2022, 12, 790263. [Google Scholar] [CrossRef]
  52. Garneau-Tsodikova, S.; Labby, K.J. Mechanisms of Resistance to Aminoglycoside Antibiotics: Overview and Perspectives. Medchemcomm 2016, 7, 11–27. [Google Scholar] [CrossRef]
  53. Nikaido, H. RND Transporters in the Living World. Res. Microbiol. 2018, 169, 363–371. [Google Scholar] [CrossRef]
  54. Yamasaki, S.; Zwama, M.; Yoneda, T.; Hayashi-Nishino, M.; Nishino, K. Drug Resistance and Physiological Roles of RND Multidrug Efflux Pumps in Salmonella enterica, Escherichia coli and Pseudomonas aeruginosa: This Article Is Part of the Antimicrobial Efflux Collection. Microbiology 2023, 169, 001322. [Google Scholar] [CrossRef]
  55. Lyu, M.; Ayala, J.C.; Chirakos, I.; Su, C.-C.; Shafer, W.M.; Yu, E.W. Structural Basis of Peptide-Based Antimicrobial Inhibition of a Resistance-Nodulation-Cell Division Multidrug Efflux Pump. Microbiol. Spectr. 2022, 10, e02990-22. [Google Scholar] [CrossRef] [PubMed]
  56. Ghai, I. Electrophysiological Insights into Antibiotic Translocation and Resistance: The Impact of Outer Membrane Proteins. Membranes 2024, 14, 161. [Google Scholar] [CrossRef] [PubMed]
  57. Ghai, I. A Barrier to Entry: Examining the Bacterial Outer Membrane and Antibiotic Resistance. Appl. Sci. 2023, 13, 4238. [Google Scholar] [CrossRef]
  58. Elshobary, M.E.; Ebaid, R.; Alquraishi, M.; Ende, S.S. Synergistic Microalgal Cocultivation: Boosting Flocculation, Biomass Production, and Fatty Acids Profile of Nannochloropsis oculata and Phaeodactylum tricornutum. Biomass Bioenergy 2025, 193, 107595. [Google Scholar] [CrossRef]
  59. Ahsan, A.; Thomas, N.; Barnes, T.J.; Subramaniam, S.; Loh, T.C.; Joyce, P.; Prestidge, C.A. Lipid Nanocarriers-Enabled Delivery of Antibiotics and Antimicrobial Adjuvants to Overcome Bacterial Biofilms. Pharmaceutics 2024, 16, 396. [Google Scholar] [CrossRef]
  60. Parvin, N.; Joo, S.W.; Mandal, T.K. Nanomaterial-Based Strategies to Combat Antibiotic Resistance: Mechanisms and Applications. Antibiotics 2025, 14, 207. [Google Scholar] [CrossRef]
  61. Chaudhari, R.; Singh, K.; Kodgire, P. Biochemical and Molecular Mechanisms of Antibiotic Resistance in Salmonella spp. Res. Microbiol. 2022, 174, 103985. [Google Scholar] [CrossRef]
  62. Mahdally, N.H.; George, R.F.; Kashef, M.T.; Al-Ghobashy, M.; Murad, F.E.; Attia, A.S. Staquorsin: A Novel Staphylococcus aureus Agr-Mediated Quorum Sensing Inhibitor Impairing Virulence In Vivo Without Notable Resistance Development. Front. Microbiol. 2021, 12, 700494. [Google Scholar] [CrossRef]
  63. Gbian, D.L.; Omri, A. The Impact of an Efflux Pump Inhibitor on the Activity of Free and Liposomal Antibiotics Against Pseudomonas aeruginosa. Pharmaceutics 2021, 13, 577. [Google Scholar] [CrossRef]
  64. Lopez-Fernandez, M.; Westmeijer, G.; Turner, S.; Broman, E.; Ståhle, M.; Bertilsson, S.; Dopson, M. Thiobacillus as a Key Player for Biofilm Formation in Oligotrophic Groundwaters of the Fennoscandian Shield. npj Biofilms Microbiomes 2023, 9, 41. [Google Scholar] [CrossRef]
  65. Li, Y.; Li, X.; Hao, Y.; Liu, Y.; Dong, Z.; Li, K. Biological and Physiochemical Methods of Biofilm Adhesion Resistance Control of Medical-Context Surface. Int. J. Biol. Sci. 2021, 17, 1769–1781. [Google Scholar] [CrossRef] [PubMed]
  66. Xu, Y.; Zeng, C.; Wen, H.; Shi, Q.; Zhao, X.; Meng, Q.; Li, X.; Xiao, J. Discovery of AI-2 Quorum Sensing Inhibitors Targeting the LsrK/HPr Protein–Protein Interaction Site by Molecular Dynamics Simulation, Virtual Screening, and Bioassay Evaluation. Pharmaceuticals 2023, 16, 737. [Google Scholar] [CrossRef] [PubMed]
  67. Fatima, M.; Amin, A.; Alharbi, M.; Ishtiaq, S.; Sajjad, W.; Ahmad, F.; Ahmad, S.; Hanif, F.; Faheem, M.; Khalil, A.A.K. Quorum Quenchers from Reynoutria japonica in the Battle against Methicillin-Resistant Staphylococcus aureus (MRSA). Molecules 2023, 28, 2635. [Google Scholar] [CrossRef]
  68. Shi, Q.; Wen, H.; Xu, Y.; Zhao, X.; Zhang, J.; Li, Y.; Meng, Q.; Yu, F.; Xiao, J.; Li, X. Virtual Screening–Based Discovery of AI-2 Quorum Sensing Inhibitors That Interact with an Allosteric Hydrophobic Site of LsrK and Their Functional Evaluation. Front. Chem. 2023, 11, 1185224. [Google Scholar] [CrossRef]
  69. Zhao, X.; Yu, Z.; Ding, T. Quorum-Sensing Regulation of Antimicrobial Resistance in Bacteria. Microorganisms 2020, 8, 425. [Google Scholar] [CrossRef]
  70. Yin, L.; Zhang, P.-P.; Wang, W.; Tang, S.; Deng, S.-M.; Jia, A.-Q. 3-Phenylpropan-1-Amine Enhanced Susceptibility of Serratia Marcescens to Ofloxacin by Occluding Quorum Sensing. Microbiol. Spectr. 2022, 10, e01829-22. [Google Scholar] [CrossRef]
  71. Křížkovská, B.; Hoang, L.; Brdová, D.; Klementová, K.; Szemerédi, N.; Loučková, A.; Kronusová, O.; Spengler, G.; Kaštánek, P.; Hajšlová, J.; et al. Modulation of the Bacterial Virulence and Resistance by Well-Known European Medicinal Herbs. J. Ethnopharmacol. 2023, 312, 116484. [Google Scholar] [CrossRef]
  72. Patel, K.; Panchal, R.; Sakariya, B.; Gevariya, M.; Raiyani, R.; Soni, R.; Goswami, D. Combatting Antibiotic Resistance by Exploring the Promise of Quorum Quenching in Targeting Bacterial Virulence. Microbe 2025, 6, 100224. [Google Scholar] [CrossRef]
  73. Qin, X.; Vila-Sanjurjo, C.; Singh, R.; Philipp, B.; Goycoolea, F.M. Screening of Bacterial Quorum Sensing Inhibitors in a Vibrio Fischeri LuxR-Based Synthetic Fluorescent E. Coli Biosensor. Pharmaceuticals 2020, 13, 263. [Google Scholar] [CrossRef]
  74. Mirpour, M.; Zahmatkesh, H. Ketoprofen Attenuates Las/Rhl Quorum-Sensing (QS) Systems of Pseudomonas aeruginosa: Molecular and Docking Studies. Mol. Biol. Rep. 2024, 51, 133. [Google Scholar] [CrossRef]
  75. Saeed, A.; Ali, H.; Yasmin, A.; Baig, M.; Ullah, A.; Kazmi, A.; Ahmed, M.A.; Albadrani, G.M.; El-Demerdash, F.M.; Bibi, M.; et al. Unveiling the Antibiotic Susceptibility and Antimicrobial Potential of Bacteria from Human Breast Milk of Pakistani Women: An Exploratory Study. BioMed Res. Int. 2023, 2023, 6399699. [Google Scholar] [CrossRef] [PubMed]
  76. Lin, Y.-C.; Chen, E.H.-L.; Chen, R.P.-Y.; Dunny, G.M.; Hu, W.-S.; Lee, K.-T. Probiotic Bacillus Affects Enterococcus faecalis Antibiotic Resistance Transfer by Interfering with Pheromone Signaling Cascades. Appl. Environ. Microbiol. 2021, 87, e00442-21. [Google Scholar] [CrossRef] [PubMed]
  77. Khalil, T.; Okla, M.K.; Al-Qahtani, W.H.; Ali, F.; Zahra, M.; Shakeela, Q.; Ahmed, S.; Akhtar, N.; AbdElgawad, H.; Asif, R.; et al. Tracing Probiotic Producing Bacterial Species from Gut of Buffalo (Bubalus bubalis), South-East-Asia. Braz. J. Biol. 2024, 84, e259094. [Google Scholar] [CrossRef] [PubMed]
  78. Wang, S.; Zhang, M.; Yu, L.; Tian, F.; Lu, W.; Wang, G.; Chen, W.; Wang, J.; Zhai, Q. Evaluation of the Potential Protective Effects of Lactobacillus Strains Against Helicobacter pylori Infection: A Randomized, Double-Blinded, Placebo-Controlled Trial. Can. J. Infect. Dis. Med. Microbiol. 2022, 2022, 6432750. [Google Scholar] [CrossRef]
  79. Chung The, H.; Nguyen Ngoc Minh, C.; Tran Thi Hong, C.; Nguyen Thi Nguyen, T.; Pike, L.J.; Zellmer, C.; Pham Duc, T.; Tran, T.-A.; Ha Thanh, T.; Van, M.P.; et al. Exploring the Genomic Diversity and Antimicrobial Susceptibility of Bifidobacterium pseudocatenulatum in a Vietnamese Population. Microbiol. Spectr. 2021, 9, e00526-21. [Google Scholar] [CrossRef]
  80. Wu, Y.; Wang, Y.; Hu, A.; Shu, X.; Huang, W.; Liu, J.; Wang, B.; Zhang, R.; Yue, M.; Yang, C. Lactobacillus plantarum-Derived Postbiotics Prevent Salmonella-Induced Neurological Dysfunctions by Modulating Gut–Brain Axis in Mice. Front. Nutr. 2022, 9, 946096. [Google Scholar] [CrossRef]
  81. Badr, H.; Nabil, N.M.; Tawakol, M.M. Effects of the Prebiotic Lactoferrin on Multidrug-Resistant Escherichia coli Infections in Broiler Chickens. Vet. World 2021, 14, 2197–2205. [Google Scholar] [CrossRef]
  82. Kaczmarczyk, M.; Szulińska, M.; Łoniewski, I.; Kręgielska-Narożna, M.; Skonieczna-Żydecka, K.; Kosciolek, T.; Bezshapkin, V.; Bogdański, P. Treatment with Multi-Species Probiotics Changes the Functions, Not the Composition of Gut Microbiota in Postmenopausal Women with Obesity: A Randomized, Double-Blind, Placebo-Controlled Study. Front. Cell. Infect. Microbiol. 2022, 12, 815798. [Google Scholar] [CrossRef]
  83. Kumar, S.; Kumar, B.; Chouraddi, R.; Bhatia, M.; Rashmi, H.M.; Vishnu Behare, P.; Tyagi, N. In Vitro Screening for Potential Probiotic Properties of Ligilactobacillus salivarius Isolated from Cattle Calves. Curr. Res. Biotechnol. 2022, 4, 275–289. [Google Scholar] [CrossRef]
  84. Pilakkavil Chirakkara, S.; Abraham, A. Bacillus Strains from the Ovaries of Swiss Albino Mice (Mus musculus): Deciphering of Probiotic Potential through an in Vitro Approach. Biomedicine 2022, 42, 1200–1208. [Google Scholar] [CrossRef]
  85. Rodríguez-Sorrento, A.; Castillejos, L.; López-Colom, P.; Cifuentes-Orjuela, G.; Rodríguez-Palmero, M.; Moreno-Muñoz, J.A.; Martín-Orúe, S.M. Effects of Bifidobacterium longum Subsp. Infantis CECT 7210 and Lactobacillus rhamnosus HN001, Combined or Not with Oligofructose-Enriched Inulin, on Weaned Pigs Orally Challenged with Salmonella typhimurium. Front. Microbiol. 2020, 11, 2012. [Google Scholar] [CrossRef]
  86. 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. 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. [Google Scholar] [CrossRef] [PubMed]
  87. Lockyer, S.; Stanner, S. Prebiotics—An Added Benefit of Some Fibre Types. Nutr. Bull. 2019, 44, 74–91. [Google Scholar] [CrossRef]
  88. Huang, F.-C.; Huang, S.-C. The Combined Beneficial Effects of Postbiotic Butyrate on Active Vitamin D3-Orchestrated Innate Immunity to Salmonella Colitis. Biomedicines 2021, 9, 1296. [Google Scholar] [CrossRef]
  89. Salemi, R.; Vivarelli, S.; Ricci, D.; Scillato, M.; Santagati, M.; Gattuso, G.; Falzone, L.; Libra, M. Lactobacillus rhamnosus GG Cell-Free Supernatant as a Novel Anti-Cancer Adjuvant. J. Transl. Med. 2023, 21, 195. [Google Scholar] [CrossRef]
  90. Huang, F.-C.; Lu, Y.-T.; Liao, Y.-H. Beneficial Effect of Probiotics on Pseudomonas aeruginosa–Infected Intestinal Epithelial Cells through Inflammatory IL-8 and Antimicrobial Peptide Human Beta-Defensin-2 Modulation. Innate Immun. 2020, 26, 592–600. [Google Scholar] [CrossRef]
  91. Gu, M.; Nguyen, H.T.; Cho, J.; Suh, J.; Cheng, J. Characterization of Leuconostoc mesenteroides MJM60376 as an Oral Probiotic and Its Antibiofilm Activity. Mol. Oral Microbiol. 2023, 38, 145–157. [Google Scholar] [CrossRef]
  92. Zhang, Q.; Zhao, W.; Zhao, Y.; Duan, S.; Liu, W.-H.; Zhang, C.; Sun, S.; Wang, T.; Wang, X.; Hung, W.-L.; et al. In Vitro Study of Bifidobacterium lactis BL-99 with Fructooligosaccharide Synbiotics Effected on the Intestinal Microbiota. Front. Nutr. 2022, 9, 890316. [Google Scholar] [CrossRef]
  93. Geng, S.; Zhang, T.; Gao, J.; Li, X.; Chitrakar, B.; Mao, K.; Sang, Y. In Vitro Screening of Synbiotics Composed of Lactobacillus Paracasei VL8 and Various Prebiotics and Mechanism to Inhibits the Growth of Salmonella typhimurium. LWT 2023, 180, 114666. [Google Scholar] [CrossRef]
  94. Piatek, J.; Krauss, H.; Ciechelska-Rybarczyk, A.; Bernatek, M.; Wojtyla-Buciora, P.; Sommermeyer, H. In-Vitro Growth Inhibition of Bacterial Pathogens by Probiotics and a Synbiotic: Product Composition Matters. Int. J. Environ. Res. Public Health 2020, 17, 3332. [Google Scholar] [CrossRef]
  95. Kaźmierczak-Siedlecka, K.; Daca, A.; Fic, M.; Van De Wetering, T.; Folwarski, M.; Makarewicz, W. Therapeutic Methods of Gut Microbiota Modification in Colorectal Cancer Management—Fecal Microbiota Transplantation, Prebiotics, Probiotics, and Synbiotics. Gut Microbes 2020, 11, 1518–1530. [Google Scholar] [CrossRef]
  96. Sommermeyer, H.; Pituch, H.M.; Wultanska, D.; Wojtyla-Buciora, P.; Piatek, J.; Bernatek, M. Inhibition of Quinolone- and Multi-Drug-Resistant Clostridioides Difficile Strains by Multi Strain Synbiotics—An Option for Diarrhea Management in Nursing Facilities. Int. J. Environ. Res. Public Health 2021, 18, 5871. [Google Scholar] [CrossRef]
  97. Gugjoo, M.B.; Sakeena, Q.; Wani, M.Y.; Abdel-Baset Ismail, A.; Ahmad, S.M.; Shah, R.A. Mesenchymal Stem Cells: A Promising Antimicrobial Therapy in Veterinary Medicine. Microb. Pathog. 2023, 182, 106234. [Google Scholar] [CrossRef]
  98. Silva-Carvalho, A.É.; Cardoso, M.H.; Alencar-Silva, T.; Bogéa, G.M.R.; Carvalho, J.L.; Franco, O.L.; Saldanha-Araujo, F. Dissecting the Relationship between Antimicrobial Peptides and Mesenchymal Stem Cells. Pharmacol. Ther. 2022, 233, 108021. [Google Scholar] [CrossRef]
  99. Gupta, N.; Mohan, C.D.; Shanmugam, M.K.; Jung, Y.Y.; Chinnathambi, A.; Alharbi, S.A.; Ashrafizadeh, M.; Mahale, M.; Bender, A.; Kumar, A.P.; et al. CXCR4 Expression Is Elevated in TNBC Patient Derived Samples and Z-Guggulsterone Abrogates Tumor Progression by Targeting CXCL12/CXCR4 Signaling Axis in Preclinical Breast Cancer Model. Environ. Res. 2023, 232, 116335. [Google Scholar] [CrossRef]
  100. Chinipardaz, Z.; Zhong, J.M.; Yang, S. Regulation of LL-37 in Bone and Periodontium Regeneration. Life 2022, 12, 1533. [Google Scholar] [CrossRef]
  101. McCulloch, T.R.; Wells, T.J.; Souza-Fonseca-Guimaraes, F. Towards Efficient Immunotherapy for Bacterial Infection. Trends Microbiol. 2022, 30, 158–169. [Google Scholar] [CrossRef]
  102. Ye, Z.; Li, G.; Lei, J. Influencing Immunity: Role of Extracellular Vesicles in Tumor Immune Checkpoint Dynamics. Exp. Mol. Med. 2024, 56, 2365–2381. [Google Scholar] [CrossRef]
  103. Pang, X.; Liu, X.; Cheng, Y.; Zhang, C.; Ren, E.; Liu, C.; Zhang, Y.; Zhu, J.; Chen, X.; Liu, G. Sono-immunotherapeutic Nanocapturer to Combat Multidrug-resistant Bacterial Infections. Adv. Mater. 2019, 31, 1902530. [Google Scholar] [CrossRef]
  104. Wang, H.; Wang, D.; Huangfu, H.; Lv, H.; Qin, Q.; Ren, S.; Zhang, Y.; Wang, L.; Zhou, Y. Branched AuAg Nanoparticles Coated by Metal–Phenolic Networks for Treating Bacteria-Induced Periodontitis via Photothermal Antibacterial and Immunotherapy. Mater. Des. 2022, 224, 111401. [Google Scholar] [CrossRef]
  105. Mitton, D.; Ackroyd, R. A Brief Overview of Photodynamic Therapy in Europe. Photodiagn. Photodyn. Ther. 2008, 5, 103–111. [Google Scholar] [CrossRef] [PubMed]
  106. Shih, Y.-H.; Yu, C.-C.; Chang, K.-C.; Tseng, Y.-H.; Li, P.-J.; Hsia, S.-M.; Chiu, K.-C.; Shieh, T.-M. Synergistic Effect of Combination of a Temoporfin-Based Photodynamic Therapy with Potassium Iodide or Antibacterial Agents on Oral Disease Pathogens In Vitro. Pharmaceuticals 2022, 15, 488. [Google Scholar] [CrossRef]
  107. Liu, X.; Liu, S.; Mai, B.; Su, X.; Guo, X.; Chang, Y.; Dong, W.; Wang, W.; Feng, X. Synergistic Gentamicin-Photodynamic Therapy against Resistant Bacteria in Burn Wound Infections. Photodiagn. Photodyn. Ther. 2022, 39, 103034. [Google Scholar] [CrossRef]
  108. Liu, S.; Mai, B.; Jia, M.; Lin, D.; Zhang, J.; Liu, Q.; Wang, P. Synergistic Antimicrobial Effects of Photodynamic Antimicrobial Chemotherapy and Gentamicin on Staphylococcus aureus and Multidrug-Resistant Staphylococcus aureus. Photodiagn. Photodyn. Ther. 2020, 30, 101703. [Google Scholar] [CrossRef]
  109. Yang, Y.; Chien, H.; Chang, P.; Chen, Y.; Jay, M.; Tsai, T.; Chen, C. Photodynamic Inactivation of Chlorin E6-loaded CTAB-liposomes against Candida albicans. Lasers Surg. Med. 2013, 45, 175–185. [Google Scholar] [CrossRef]
  110. Wainwright, M.; Maisch, T.; Nonell, S.; Plaetzer, K.; Almeida, A.; Tegos, G.P.; Hamblin, M.R. Photoantimicrobials—Are We Afraid of the Light? Lancet Infect. Dis. 2017, 17, e49–e55. [Google Scholar] [CrossRef]
  111. Sperandio, F.F.; Huang, Y.-Y.; Hamblin, M.R. Antimicrobial Photodynamic Therapy to Kill Gram-Negative Bacteria. Recent Pat. Anti-Infect. Drug Discov. 2013, 8, 108–120. [Google Scholar] [CrossRef]
  112. Li, X.; Guo, H.; Tian, Q.; Zheng, G.; Hu, Y.; Fu, Y.; Tan, H. Effects of 5-Aminolevulinic Acid–Mediated Photodynamic Therapy on Antibiotic-Resistant Staphylococcal Biofilm: An in Vitro Study. J. Surg. Res. 2013, 184, 1013–1021. [Google Scholar] [CrossRef]
  113. de Moraes, M.; Vasconcelos, R.C.; Longo, J.P.F.; Muehlmann, L.A.; de Azevedo, R.B.; de Araújo Júnior, R.F.; Araujo, A.A.; de Lisboa Lopes Costa, A. Photodynamic Therapy Using. Chloro-Aluminum Phthalocyanine Decreases Inflammatory Response in an Experimental Rat. Periodontal Disease Model. J. Photochem. Photobiol. B Biol. 2017, 167, 208–215. [Google Scholar] [CrossRef]
  114. Kamruzzaman, M.; Iredell, J.R. CRISPR-Cas System in Antibiotic Resistance Plasmids in Klebsiella pneumoniae. Front. Microbiol. 2020, 10, 2934. [Google Scholar] [CrossRef]
  115. Tao, S.; Chen, H.; Li, N.; Fang, Y.; Xu, Y.; Liang, W. Association of CRISPR-Cas System with the Antibiotic Resistance and Virulence Genes in Nosocomial Isolates of Enterococcus. Infect. Drug Resist. 2022, 15, 6939–6949. [Google Scholar] [CrossRef] [PubMed]
  116. Greene, A.C. CRISPR-Based Antibacterials: Transforming Bacterial Defense into Offense. Trends Biotechnol. 2018, 36, 127–130. [Google Scholar] [CrossRef]
  117. Wu, Y.; Battalapalli, D.; Hakeem, M.J.; Selamneni, V.; Zhang, P.; Draz, M.S.; Ruan, Z. Engineered CRISPR-Cas Systems for the Detection and Control of Antibiotic-Resistant Infections. J. Nanobiotechnol. 2021, 19, 401. [Google Scholar] [CrossRef]
  118. Pursey, E.; Dimitriu, T.; Paganelli, F.L.; Westra, E.R.; Van Houte, S. CRISPR-Cas Is Associated with Fewer Antibiotic Resistance Genes in Bacterial Pathogens. Phil. Trans. R. Soc. B 2022, 377, 20200464. [Google Scholar] [CrossRef]
  119. Wang, Y.; Yang, J.; Sun, X.; Li, M.; Zhang, P.; Zhu, Z.; Jiao, H.; Guo, T.; Li, G. CRISPR-Cas in Acinetobacter baumannii Contributes to Antibiotic Susceptibility by Targeting Endogenous AbaI. Microbiol. Spectr. 2022, 10, e00829-22. [Google Scholar] [CrossRef]
  120. Gholizadeh, P.; Aghazadeh, M.; Ghotaslou, R.; Rezaee, M.A.; Pirzadeh, T.; Cui, L.; Watanabe, S.; Feizi, H.; Kadkhoda, H.; Kafil, H.S. Role of CRISPR-Cas System on Antibiotic Resistance Patterns of Enterococcus faecalis. Ann. Clin. Microbiol. Antimicrob. 2021, 20, 49. [Google Scholar] [CrossRef]
  121. Murugesan, A.C.; Varughese, H.S. Analysis of CRISPR–Cas System and Antimicrobial Resistance in Staphylococcus Coagulans Isolates. Lett. Appl. Microbiol. 2022, 75, 126–134. [Google Scholar] [CrossRef]
  122. Alduhaidhawi, A.H.M.; AlHuchaimi, S.N.; Al-Mayah, T.A.; Al-Ouqaili, M.T.; Alkafaas, S.S.; Muthupandian, S.; Saki, M. Prevalence of CRISPR-Cas Systems and Their Possible Association with Antibiotic Resistance in Enterococcus faecalis and Enterococcus faecium Collected from Hospital Wastewater. Infect. Drug Resist. 2022, 15, 1143–1154. [Google Scholar] [CrossRef]
  123. Sahu, R.; Singh, A.K.; Kumar, A.; Singh, K.; Kumar, P. Bacteriophages Concept and Applications: A Review on Phage Therapy. Curr. Pharm. Biotechnol. 2023, 24, 1245–1264. [Google Scholar]
  124. Ling, H.; Lou, X.; Luo, Q.; He, Z.; Sun, M.; Sun, J. Recent Advances in Bacteriophage-Based Therapeutics: Insight into the Post-Antibiotic Era. Acta Pharm. Sin. B 2022, 12, 4348–4364. [Google Scholar] [CrossRef]
  125. Dufour, N.; Delattre, R.; Ricard, J.-D.; Debarbieux, L. The Lysis of Pathogenic Escherichia coli by Bacteriophages Releases Less Endotoxin than by β-Lactams. Clin. Infect. Dis. 2017, 64, 1582–1588. [Google Scholar] [CrossRef] [PubMed]
  126. Oechslin, F.; Piccardi, P.; Mancini, S.; Gabard, J.; Moreillon, P.; Entenza, J.M.; Resch, G.; Que, Y.-A. Synergistic Interaction between Phage Therapy and Antibiotics Clears Pseudomonas aeruginosa Infection in Endocarditis and Reduces Virulence. J. Infect. Dis. 2017, 215, 703–712. [Google Scholar] [CrossRef] [PubMed]
  127. Leung, C.Y.J.; Weitz, J.S. Modeling the Synergistic Elimination of Bacteria by Phage and the Innate Immune System. J. Theor. Biol. 2017, 429, 241–252. [Google Scholar] [CrossRef]
  128. Durbas, I.; Machnik, G. Phage Therapy: An Old Concept with New Perspectives. J. Appl. Pharm. Sci. 2022, 12, 027–038. [Google Scholar] [CrossRef]
  129. Enault, F.; Briet, A.; Bouteille, L.; Roux, S.; Sullivan, M.B.; Petit, M.-A. Phages Rarely Encode Antibiotic Resistance Genes: A Cautionary Tale for Virome Analyses. ISME J. 2017, 11, 237–247. [Google Scholar] [CrossRef]
  130. Suhas, R. Structure, Function and Mechanistic Aspects of Scorpion Venom Peptides—A Boon for the Development of Novel Therapeutics. Eur. J. Med. Chem. Rep. 2022, 6, 100068. [Google Scholar] [CrossRef]
  131. Yacoub, T.; Rima, M.; Karam, M.; Sabatier, J.-M.; Fajloun, Z. Antimicrobials from Venomous Animals: An Overview. Molecules 2020, 25, 2402. [Google Scholar] [CrossRef]
  132. Ageitos, L.; Torres, M.D.T.; De La Fuente-Nunez, C. Biologically Active Peptides from Venoms: Applications in Antibiotic Resistance, Cancer, and Beyond. Int. J. Mol. Sci. 2022, 23, 15437. [Google Scholar] [CrossRef]
  133. Li, Z.; Hu, P.; Wu, W.; Wang, Y. Peptides with Therapeutic Potential in the Venom of the Scorpion Buthus martensii Karsch. Peptides 2019, 115, 43–50. [Google Scholar] [CrossRef]
  134. Zerouti, K.; Khemili, D.; Laraba-Djebari, F.; Hammoudi-Triki, D. Nontoxic Fraction of Scorpion Venom Reduces Bacterial Growth and Inflammatory Response in a Mouse Model of Infection. Toxin Rev. 2019, 40, 1–15. [Google Scholar] [CrossRef]
  135. Ahmadi, S.; Knerr, J.M.; Argemi, L.; Bordon, K.C.F.; Pucca, M.B.; Cerni, F.A.; Arantes, E.C.; Çalışkan, F.; Laustsen, A.H. Scorpion Venom: Detriments and Benefits. Biomedicines 2020, 8, 118. [Google Scholar] [CrossRef] [PubMed]
  136. Cesa-Luna, C.; Muñoz-Rojas, J.; Saab-Rincon, G.; Baez, A.; Morales-García, Y.E.; Juárez-González, V.R.; Quintero-Hernández, V. Structural Characterization of Scorpion Peptides and Their Bactericidal Activity against Clinical Isolates of Multidrug-Resistant Bacteria. PLoS ONE 2019, 14, e0222438. [Google Scholar] [CrossRef] [PubMed]
  137. Teerapo, K.; Roytrakul, S.; Sistayanarain, A.; Kunthalert, D. A Scorpion Venom Peptide Derivative BmKn‑22 with Potent Antibiofilm Activity against Pseudomonas aeruginosa. PLoS ONE 2019, 14, e0218479. [Google Scholar] [CrossRef] [PubMed]
  138. Shi, P.; Xie, S.; Yang, J.; Zhang, Y.; Han, S.; Su, S.; Yao, H. Pharmacological Effects and Mechanisms of Bee Venom and Its Main Components: Recent Progress and Perspective. Front. Pharmacol. 2022, 13, 1001553. [Google Scholar] [CrossRef]
  139. Tiwari, R.; Tiwari, G.; Lahiri, A.; Ramachandran, V.; Rai, A. Melittin: A Natural Peptide with Expanded Therapeutic Applications. Nat. Prod. J. 2022, 12, 13–29. [Google Scholar] [CrossRef]
  140. El-Seedi, H.; El-Wahed, A.A.; Yosri, N.; Musharraf, S.G.; Chen, L.; Moustafa, M.; Zou, X.; Al-Mousawi, S.; Guo, Z.; Khatib, A.; et al. Antimicrobial Properties of Apis mellifera’s Bee Venom. Toxins 2020, 12, 451. [Google Scholar] [CrossRef]
  141. Hegazi, A.G.; Abdou, A.M.H.; El-Moez, I.A.; Allah, F.M.A. Evaluation of the Antibacterial Activity of Bee Venom from Different Sources. World Appl. Sci. J. 2014, 30, 266–270. [Google Scholar]
  142. Fadl, A.E. Antibacterial and Antibiofilm Effects of Bee Venom from (Apis mellifera) on Multidrug-Resistant Bacteria (Mdrb). Al-Azhar J. Pharm. Sci. 2018, 58, 60–80. [Google Scholar] [CrossRef]
  143. Choi, J.H.; Jang, A.-Y.; Lin, S.; Lim, S.; Kim, D.; Park, K.; Han, S.M.; Yeo, J.-H.; Seo, H.S. Melittin, a Honeybee Venom-Derived Antimicrobial Peptide, May Target Methicillin-Resistant Staphylococcus aureus. Mol. Med. Rep. 2015, 12, 6483–6490. [Google Scholar] [CrossRef]
  144. Oguiura, N.; Corrêa, P.G.; Rosmino, I.L.; de Souza, A.O.; Pasqualoto, K.F.M. Antimicrobial Activity of Snake β-Defensins and Derived Peptides. Toxins 2021, 14, 1. [Google Scholar] [CrossRef]
  145. De Barros, E.; Gonçalves, R.M.; Cardoso, M.H.; Santos, N.C.; Franco, O.L.; Cândido, E.S. Snake Venom Cathelicidins as Natural Antimicrobial Peptides. Front. Pharmacol. 2019, 10, 489134. [Google Scholar] [CrossRef] [PubMed]
  146. Blaylock, R.S.M. Antibacterial Properties of KwaZulu Natal Snake Venoms. Toxicon 2000, 38, 1529–1534. [Google Scholar] [CrossRef] [PubMed]
  147. Dal Mas, C.; Rossato, L.; Shimizu, T.; Oliveira, E.B.D.; da Silva Júnior, P.I.; Meis, J.F.; Colombo, A.L.; Hayashi, M.A.F. Effects of the Natural Peptide Crotamine from a South American Rattlesnake on Candida auris, an Emergent Multidrug Antifungal Resistant Human Pathogen. Biomolecules 2019, 9, 205. [Google Scholar] [CrossRef] [PubMed]
  148. Bandeira, I.C.J.; Bandeira-Lima, D.; Mello, C.P.; Pereira, T.P.; de Menezes, R.R.P.P.B.; Sampaio, T.L.; Falcao, C.B.; Rádis-Baptista, G.; Martins, A.M.C. Antichagasic Effect of Crotalicidin, a Cathelicidin-like Vipericidin, Found in Crotalus durissus terrificus Rattlesnake’s Venom Gland. Parasitology 2017, 145, 1059–1064. [Google Scholar] [CrossRef]
  149. de Oliveira Junior, N.G.; e Silva Cardoso, M.H.; Franco, O.L. Snake Venoms: Attractive Antimicrobial Proteinaceous Compounds for Therapeutic Purposes. Cell. Mol. Life Sci. 2013, 70, 4645–4658. [Google Scholar] [CrossRef]
  150. Guo, C.; Liu, S.; Yao, Y.; Zhang, Q.; Sun, M.-Z. Past Decade Study of Snake Venom L-Amino Acid Oxidase. Toxicon 2012, 60, 302–311. [Google Scholar] [CrossRef]
  151. Oliveira, A.L.; Viegas, M.F.; da Silva, S.L.; Soares, A.M.; Ramos, M.J.; Fernandes, P.A. The Chemistry of Snake Venom and Its Medicinal Potential. Nat. Rev. Chem. 2022, 6, 451–469. [Google Scholar] [CrossRef]
  152. Wang, Y.-J.; Wang, L.; Yang, H.; Xiao, H.; Farooq, A.; Liu, Z.; Hu, M.; Shi, X. The Spider Venom Peptide Lycosin-II Has Potent Antimicrobial Activity against Clinically Isolated Bacteria. Toxins 2016, 8, 119. [Google Scholar] [CrossRef]
  153. Abreu, T.F.; Sumitomo, B.N.; Nishiyama, M.Y.; de Oliveira, U.C.; Souza, G.H.M.F.; Kitano, E.S.; Zelanis, A.; Serrano, S.M.T.; Junqueira-de-Azevedo, I.L.M.; da Silva, P.I.; et al. Peptidomics of Acanthoscurria gomesiana Spider Venom Reveals New Toxins with Potential Antimicrobial Activity. J. Proteom. 2017, 151, 232–242. [Google Scholar] [CrossRef]
  154. Wu, T.; Wang, M.; Wu, W.; Luo, Q.; Jiang, L.; Tao, H.; Deng, M. Spider Venom Peptides as Potential Drug Candidates Due to Their Anticancer and Antinociceptive Activities. J. Venom. Anim. Toxins Incl. Trop. Dis. 2019, 25, e146318. [Google Scholar] [CrossRef]
  155. Torres-Larios, A.; Gurrola, G.B.; Zamudio, F.Z.; Possani, L.D. Hadrurin, a New Antimicrobial Peptide from the Venom of the Scorpion Hadrurus aztecus. Eur. J. Biochem. 2000, 267, 5023–5031. [Google Scholar] [CrossRef] [PubMed]
  156. Corzo, G.; Escoubas, P.; Villegas, E.; Barnham, K.J.; He, W.; Norton, R.S.; Nakajima, T. Characterization of Unique Amphipathic Antimicrobial Peptides from Venom of the Scorpion Pandinus Imperator. Biochem. J. 2001, 359, 35–45. [Google Scholar] [CrossRef] [PubMed]
  157. Rodríguez, A.; Villegas, E.; Montoya-Rosales, A.; Rivas-Santiago, B.; Corzo, G. Characterization of Antibacterial and Hemolytic Activity of Synthetic Pandinin 2 Variants and Their Inhibition against Mycobacterium tuberculosis. PLoS ONE 2014, 9, e101742. [Google Scholar] [CrossRef]
  158. Zeng, X.-C.; Wang, S.; Nie, Y.; Zhang, L.; Luo, X. Characterization of BmKbpp, a Multifunctional Peptide from the Chinese Scorpion Mesobuthus martensii Karsch: Gaining Insight into a New Mechanism for the Functional Diversification of Scorpion Venom Peptides. Peptides 2012, 33, 44–51. [Google Scholar] [CrossRef]
  159. Ramírez-Carreto, S.; Jiménez-Vargas, J.M.; Rivas-Santiago, B.; Corzo, G.; Possani, L.D.; Becerril, B.; Ortiz, E. Peptides from the Scorpion Vaejovis punctatus with Broad Antimicrobial Activity. Peptides 2015, 73, 51–59. [Google Scholar] [CrossRef]
  160. Guo, X.; Ma, C.; Du, Q.; Wei, R.; Wang, L.; Zhou, M.; Chen, T.; Shaw, C. Two Peptides, TsAP-1 and TsAP-2, from the Venom of the Brazilian Yellow Scorpion, Tityus serrulatus: Evaluation of Their Antimicrobial and Anticancer Activities. Biochimie 2013, 95, 1784–1794. [Google Scholar] [CrossRef]
  161. Cao, L.; Li, Z.; Zhang, R.; Wu, Y.; Li, W.; Cao, Z. StCT2, a New Antibacterial Peptide Characterized from the Venom of the Scorpion Scorpiops tibetanus. Peptides 2012, 36, 213–220. [Google Scholar] [CrossRef]
  162. Ratajczak, M.; Kaminska, D.; Matuszewska, E.; Hołderna-Kedzia, E.; Rogacki, J.; Matysiak, J. Promising Antimicrobial Properties of Bioactive Compounds from Different Honeybee Products. Molecules 2021, 26, 4007. [Google Scholar] [CrossRef]
  163. Pluta, P.; Sokol, R. Changes in the Expression of Antimicrobial Peptide Genes in Honey Bees (Apis mellifera) under the Influence of Various Pathogens. Ann. Parasitol. 2020, 66, 457–465. [Google Scholar]
  164. Falcao, C.B.; de La Torre, B.; Pérez-Peinado, C.; Barron, A.E.; Andreu, D.; Rádis-Baptista, G. Vipericidins: A Novel Family of Cathelicidin-Related Peptides from the Venom Gland of South American Pit Vipers. Amino Acids 2014, 46, 2561–2571. [Google Scholar] [CrossRef]
  165. Blower, R.J.; Popov, S.G.; van Hoek, M.L. Cathelicidin Peptide Rescues G. mellonella Infected with B. Anthracis. Virulence 2018, 9, 287–293. [Google Scholar] [CrossRef] [PubMed]
  166. Creane, S.E.; Carlile, S.R.; Downey, D.; Weldon, S.; Dalton, J.P.; Taggart, C.C. The Impact of Lung Proteases on Snake-Derived Antimicrobial Peptides. Biomolecules 2021, 11, 1106. [Google Scholar] [CrossRef] [PubMed]
  167. Kuhn-Nentwig, L. Complex Precursor Structures of Cytolytic Cupiennins Identified in Spider Venom Gland Transcriptomes. Sci. Rep. 2021, 11, 4009. [Google Scholar] [CrossRef] [PubMed]
  168. Salimo, Z.M.; Barros, A.L.; Adrião, A.A.; Rodrigues, A.M.; Sartim, M.A.; de Oliveira, I.S.; Pucca, M.B.; Baia-da-Silva, D.C.; Monteiro, W.M.; de Melo, G.C. Toxins from Animal Venoms as a Potential Source of Antimalarials: A Comprehensive Review. Toxins 2023, 15, 375. [Google Scholar] [CrossRef]
  169. Rodriguez, C.; Carrasco, J.; Bruner-Montero, G.; Júnior, O.R.P.; Gutiérrez, M.; Díaz-Ferguson, E. Components and Biological Activities of Venom from Lionfishes (Scorpaenidae: Pterois). Mar. Drugs 2025, 23, 55. [Google Scholar] [CrossRef]
  170. Shin, M.K.; Hwang, I.-W.; Kim, Y.; Kim, S.T.; Jang, W.; Lee, S.; Bang, W.Y.; Bae, C.-H.; Sung, J.-S. Antibacterial and Anti-Inflammatory Effects of Novel Peptide Toxin from the Spider Pardosa astrigera. Antibiotics 2020, 9, 422. [Google Scholar] [CrossRef]
  171. Han, Y.; Zhang, M.; Lai, R.; Zhang, Z. Chemical Modifications to Increase the Therapeutic Potential of Antimicrobial Peptides. Peptides 2021, 146, 170666. [Google Scholar] [CrossRef]
  172. Carratalá, J.V.; Serna, N.; Villaverde, A.; Vázquez, E.; Ferrer-Miralles, N. Nanostructured Antimicrobial Peptides: The Last Push towards Clinics. Biotechnol. Adv. 2020, 44, 107603. [Google Scholar] [CrossRef]
  173. Bellotto, O.; Semeraro, S.; Bandiera, A.; Tramer, F.; Pavan, N.; Marchesan, S. Polymer Conjugates of Antimicrobial Peptides (AMPs) with D-Amino Acids (D-Aa): State of the Art and Future Opportunities. Pharmaceutics 2022, 14, 446. [Google Scholar] [CrossRef]
  174. Tyagi, R.; Srivastava, M.; Jain, P.; Pandey, R.P.; Asthana, S.; Kumar, D.; Raj, V.S. Development of Potential Proteasome Inhibitors against Mycobacterium tuberculosis. J. Biomol. Struct. Dyn. 2022, 40, 2189–2203. [Google Scholar] [CrossRef]
  175. Nazir, F.; Tabish, T.A.; Tariq, F.; Iftikhar, S.; Wasim, R.; Shahnaz, G. Stimuli-Sensitive Drug Delivery Systems for Site-Specific Antibiotic Release. Drug Discov. Today 2022, 27, 1698–1705. [Google Scholar] [CrossRef] [PubMed]
  176. Kassinger, S.J.; van Hoek, M.L. Biofilm Architecture: An Emerging Synthetic Biology Target. Synth. Syst. Biotechnol. 2020, 5, 1–10. [Google Scholar] [CrossRef] [PubMed]
  177. Himanshu; Mukherjee, R.; Vidic, J.; Leal, E.; da Costa, A.C.; Prudencio, C.R.; Raj, V.S.; Chang, C.-M.; Pandey, R.P. Nanobiotics and the One Health Approach: Boosting the Fight Against Antimicrobial Resistance at the Nanoscale. Biomolecules 2023, 13, 1182. [Google Scholar] [CrossRef]
  178. Elfadil, D.; Elkhatib, W.F.; El-Sayyad, G.S. Promising Advances in Nanobiotic-Based Formulations for Drug Specific Targeting Against Multidrug-Resistant Microbes and Biofilm-Associated Infections. Microb. Pathog. 2022, 170, 105721. [Google Scholar] [CrossRef]
  179. Mohamed, M.A.; Nasr, M.; Elkhatib, W.F.; Eltayeb, W.N.; Elshamy, A.A.; El-Sayyad, G.S. Nanobiotic Formulations as Promising Advances for Combating MRSA Resistance: Susceptibilities and Post-Antibiotic Effects of Clindamycin, Doxycycline, and Linezolid. RSC Adv. 2021, 11, 39696–39706. [Google Scholar] [CrossRef] [PubMed]
  180. Milewska, S.; Niemirowicz-Laskowska, K.; Siemiaszko, G.; Nowicki, P.; Wilczewska, A.Z.; Car, H. Current Trends and Challenges in Pharmacoeconomic Aspects of Nanocarriers as Drug Delivery Systems for Cancer Treatment. Int. J. Nanomed. 2021, 16, 6593–6644. [Google Scholar] [CrossRef]
  181. Chakraborty, N.; Jha, D.; Roy, I.; Kumar, P.; Gaurav, S.S.; Marimuthu, K.; Ng, O.-T.; Lakshminarayanan, R.; Verma, N.K.; Gautam, H.K. Nanobiotics against Antimicrobial Resistance: Harnessing the Power of Nanoscale Materials and Technologies. J. Nanobiotechnol. 2022, 20, 375. [Google Scholar] [CrossRef]
  182. Makabenta, J.M.V.; Nabawy, A.; Li, C.-H.; Schmidt-Malan, S.; Patel, R.; Rotello, V.M. Nanomaterial-Based Therapeutics for Antibiotic-Resistant Bacterial Infections. Nat. Rev. Microbiol. 2021, 19, 23–36. [Google Scholar] [CrossRef]
  183. Geersing, A.; de Vries, R.H.; Jansen, G.; Rots, M.G.; Roelfes, G. Folic Acid Conjugates of a Bleomycin Mimic for Selective Targeting of Folate Receptor Positive Cancer Cells. Bioorganic Med. Chem. Lett. 2019, 29, 1922–1927. [Google Scholar] [CrossRef]
  184. Tripathi, N.; Goshisht, M.K. Recent Advances and Mechanistic Insights into Antibacterial Activity, Antibiofilm Activity, and Cytotoxicity of Silver Nanoparticles. ACS Appl. Bio Mater. 2022, 5, 1391–1463. [Google Scholar] [CrossRef]
  185. Wagi, S.; Ahmed, A. Bacterial Nanobiotic Potential. Green Process. Synth. 2020, 9, 203–211. [Google Scholar] [CrossRef]
  186. Brown, A.N.; Smith, K.; Samuels, T.A.; Lu, J.; Obare, S.O.; Scott, M.E. Nanoparticles Functionalized with Ampicillin Destroy Multiple-Antibiotic-Resistant Isolates of Pseudomonas aeruginosa and Enterobacter aerogenes and Methicillin-Resistant Staphylococcus aureus. Appl. Environ. Microbiol. 2012, 78, 2768–2774. [Google Scholar] [CrossRef] [PubMed]
  187. Okkeh, M.; Bloise, N.; Restivo, E.; De Vita, L.; Pallavicini, P.; Visai, L. Gold Nanoparticles: Can. They Be the Next Magic. Bullet for Multidrug-Resistant Bacteria? Nanomaterials 2021, 11, 312. [Google Scholar] [CrossRef]
  188. Cheng, X.; Pei, X.; Xie, W.; Chen, J.; Li, Y.; Wang, J.; Gao, H.; Wan, Q. pH-Triggered Size-Tunable Silver Nanoparticles: Targeted Aggregation for Effective Bacterial Infection Therapy. Small 2022, 18, 2200915. [Google Scholar] [CrossRef]
  189. Mamun, M.M.; Sorinolu, A.J.; Munir, M.; Vejerano, E.P. Nanoantibiotics: Functions and Properties at the Nanoscale to Combat Antibiotic Resistance. Front. Chem. 2021, 9, 687660. [Google Scholar] [CrossRef]
  190. Chung, H.J.; Castro, C.M.; Im, H.; Lee, H.; Weissleder, R. A Magneto-DNA Nanoparticle System for Rapid Detection and Phenotyping of Bacteria. Nat. Nanotechnol. 2013, 8, 369–375. [Google Scholar] [CrossRef]
  191. Kulshrestha, S.; Khan, S.; Hasan, S.; Khan, M.E.; Misba, L.; Khan, A.U. Calcium Fluoride Nanoparticles Induced Suppression of Streptococcus Mutans Biofilm: An In Vitro and In Vivo Approach. Appl. Microbiol. Biotechnol. 2016, 100, 1901–1914. [Google Scholar] [CrossRef]
  192. Wang, Y. Liposome as a Delivery System for the Treatment of Biofilm-mediated Infections. J. Appl. Microbiol. 2021, 131, 2626–2639. [Google Scholar] [CrossRef]
  193. Cano, A.; Ettcheto, M.; Espina, M.; López-Machado, A.; Cajal, Y.; Rabanal, F.; Sánchez-López, E.; Camins, A.; García, M.L.; Souto, E.B. State-of-the-Art Polymeric Nanoparticles as Promising Therapeutic Tools against Human Bacterial Infections. J. Nanobiotechnol. 2020, 18, 156. [Google Scholar] [CrossRef]
  194. Forier, K.; Raemdonck, K.; De Smedt, S.C.; Demeester, J.; Coenye, T.; Braeckmans, K. Lipid and Polymer Nanoparticles for Drug Delivery to Bacterial Biofilms. J. Control. Release 2014, 190, 607–623. [Google Scholar] [CrossRef]
  195. Rajendiran, K.; Zhao, Z.; Pei, D.-S.; Fu, A. Antimicrobial Activity and Mechanism of Functionalized Quantum Dots. Polymers 2019, 11, 1670. [Google Scholar] [CrossRef] [PubMed]
  196. Courtney, C.M.; Goodman, S.M.; McDaniel, J.A.; Madinger, N.E.; Chatterjee, A.; Nagpal, P. Photoexcited Quantum Dots for Killing Multidrug-Resistant Bacteria. Nature Mater. 2016, 15, 529–534. [Google Scholar] [CrossRef] [PubMed]
  197. Kalhapure, R.S.; Suleman, N.; Mocktar, C.; Seedat, N.; Govender, T. Nanoengineered Drug Delivery Systems for Enhancing Antibiotic Therapy. J. Pharm. Sci. 2015, 104, 872–905. [Google Scholar] [CrossRef] [PubMed]
  198. Gupta, B.; Malviya, R.; Srivastava, S.; Ahmad, I.; Rab, S.O.; Uniyal, P. Construction, Features and Regulatory Aspects of Organ-Chip for Drug Delivery Applications: Advances and Prospective. Curr. Pharm. Des. 2024, 30, 1952–1965. [Google Scholar] [CrossRef]
  199. Lazar, V.; Oprea, E.; Ditu, L. Resistance, Tolerance, Virulence and Bacterial Pathogen Fitness—Current State and Envisioned Solutions for the near Future. Pathogens 2023, 12, 746. [Google Scholar] [CrossRef]
  200. Kulchar, R.J.; Singh, R.; Ding, S.; Alexander, E.; Leong, K.W.; Daniell, H. Delivery of Biologics: Topical Administration. Biomaterials 2023, 302, 122312. [Google Scholar] [CrossRef]
  201. Aslam, B.; Wang, W.; Arshad, M.I.; Khurshid, M.; Muzammil, S.; Rasool, M.H.; Nisar, M.A.; Alvi, R.F.; Aslam, M.A.; Qamar, M.U.; et al. Antibiotic Resistance: A Rundown of a Global Crisis. Infect. Drug Resist. 2018, 11, 1645–1658. [Google Scholar] [CrossRef]
Figure 1. Mechanisms of antibiotic resistance in bacteria include decreased uptake (Antibiotic A, red), enzyme degradation (Antibiotic B, purple), resistance mutation (Antibiotic C, orange), and efflux pumps (Antibiotic D, blue-green).
Figure 1. Mechanisms of antibiotic resistance in bacteria include decreased uptake (Antibiotic A, red), enzyme degradation (Antibiotic B, purple), resistance mutation (Antibiotic C, orange), and efflux pumps (Antibiotic D, blue-green).
Pharmaceuticals 18 00402 g001
Figure 2. Schematic representation of the major multidrug efflux pump families. RND (tripartite system, proton-gradient-dependent, exports large molecules), ABC (ATP-dependent, found in all domains of life), MATE (sodium/proton-gradient-dependent, single-component system), MFS (proton-gradient-dependent, single-component, ubiquitous), SMR (proton-gradient-dependent, smallest transporter, functions as a dimer), and PACE (proton-gradient-dependent, specific to Proteobacteria, resists biocides and dyes). Each transporter is depicted with distinct shapes and colors, indicating its energy source and substrate transport mechanism. A generic antibiotic is symbolized as a pill.
Figure 2. Schematic representation of the major multidrug efflux pump families. RND (tripartite system, proton-gradient-dependent, exports large molecules), ABC (ATP-dependent, found in all domains of life), MATE (sodium/proton-gradient-dependent, single-component system), MFS (proton-gradient-dependent, single-component, ubiquitous), SMR (proton-gradient-dependent, smallest transporter, functions as a dimer), and PACE (proton-gradient-dependent, specific to Proteobacteria, resists biocides and dyes). Each transporter is depicted with distinct shapes and colors, indicating its energy source and substrate transport mechanism. A generic antibiotic is symbolized as a pill.
Pharmaceuticals 18 00402 g002
Figure 3. Mechanism diagram of the quorum quenching inhibition of the quorum sensing of pathogenic bacteria.
Figure 3. Mechanism diagram of the quorum quenching inhibition of the quorum sensing of pathogenic bacteria.
Pharmaceuticals 18 00402 g003
Figure 4. Definition of probiotics, prebiotics, synbiotics, and postbiotics.
Figure 4. Definition of probiotics, prebiotics, synbiotics, and postbiotics.
Pharmaceuticals 18 00402 g004
Figure 5. MSCs produce antimicrobial peptides (AMPs), including hepcidin, LL-37, and β-defensin, and their applications are shown.
Figure 5. MSCs produce antimicrobial peptides (AMPs), including hepcidin, LL-37, and β-defensin, and their applications are shown.
Pharmaceuticals 18 00402 g005
Figure 6. Schematic illustration of antimicrobial photodynamic therapy.
Figure 6. Schematic illustration of antimicrobial photodynamic therapy.
Pharmaceuticals 18 00402 g006
Figure 7. CRISPR-Cas systems as antimicrobial agents. CRISPR-Cas targets bacterial resistance genes by integrating foreign DNA sequences into the CRISPR array. Processed crRNA guides Cas9 to cleave matching DNA in plasmids or chromosomes, disrupting resistance genes. This leads to re-sensitization to antibiotics or bacterial cell death, offering a targeted strategy against antibiotic resistance.
Figure 7. CRISPR-Cas systems as antimicrobial agents. CRISPR-Cas targets bacterial resistance genes by integrating foreign DNA sequences into the CRISPR array. Processed crRNA guides Cas9 to cleave matching DNA in plasmids or chromosomes, disrupting resistance genes. This leads to re-sensitization to antibiotics or bacterial cell death, offering a targeted strategy against antibiotic resistance.
Pharmaceuticals 18 00402 g007
Table 1. Comprehensive analysis of some bacterial antibiotic resistance mechanisms across bacterial species.
Table 1. Comprehensive analysis of some bacterial antibiotic resistance mechanisms across bacterial species.
Type of ResistanceBacterial SpeciesSpecific MechanismAntibiotic ClassResistance StrategyClinical ImplicationsReferences
Intrinsic Pseudomonas aeruginosaEfflux Pump OverexpressionCarbapenemsReduced Antibiotic AccumulationHigh Treatment Failure Rates[25]
Acquired Staphylococcus aureusmecA Gene Horizontalβ-Lactam AntibioticsTransfer Penicillin-Binding Protein ModificationMRSA Infections[26]
Adaptive Acinetobacter baumanniiBiofilm FormationMultiple AntibioticsPhenotypic HeterogeneityPersistent Infections[27]
Intrinsic Klebsiella pneumoniaeβ-Lactamase ProductionCephalosporinsEnzymatic Antibiotic DegradationExtended-Spectrum Resistance[28]
Acquired Enterococcus faeciumVancomycin ResistanceGene (vanA) GlycopeptideAntibiotics Target Site ModificationVRE Nosocomial Infections[29]
Adaptive Multiple Bacterial SpeciesMetabolic DormancyBroad-SpectrumReduced Metabolic ActivityAntibiotic Tolerance[30]
Intrinsic Multiple Bacterial SpeciesABC Transporter RegulationBroad-Spectrum Antibiotics Complex Efflux Mechanism TherapeuticTargeting Challenges[31,32]
Table 5. Role of nanobiotics in defeating the challenge of antimicrobial resistance.
Table 5. Role of nanobiotics in defeating the challenge of antimicrobial resistance.
AdvantageDescriptionRef.
Reduced Toxicity and Enhanced Stability Encapsulating antibiotics in nanoparticles can reduce their overall toxicity and enhance their stability in vivo, preventing premature degradation.[173]
Targeted Delivery to Sites of Infection Nanoparticles can be designed to target specific sites of infection, either passively or actively, allowing for higher antibiotic concentrations at the infected site while minimizing systemic exposure and adverse effects.[174]
Stimuli-Sensitive Drug Release Nanoparticles can be engineered to release antibiotics in response to specific stimuli (e.g., pH, enzymes, reactive oxygen species) present in the infected tissues, enabling targeted and controlled drug release.[175]
Directed towards Biofilm Microenvironments Nanoparticles can be tailored to target and disrupt biofilms, which are a significant contributor to antimicrobial resistance, by exploiting the unique microenvironment of biofilms.[176]
Combined Physical Therapy Nanoparticles can be combined with other physical therapies, such as photothermal therapy (PTT) and antibacterial photodynamic therapy (aPDT), to enhance their antimicrobial efficacy through synergistic mechanisms.[177]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Elshobary, M.E.; Badawy, N.K.; Ashraf, Y.; Zatioun, A.A.; Masriya, H.H.; Ammar, M.M.; Mohamed, N.A.; Mourad, S.; Assy, A.M. Combating Antibiotic Resistance: Mechanisms, Multidrug-Resistant Pathogens, and Novel Therapeutic Approaches: An Updated Review. Pharmaceuticals 2025, 18, 402. https://doi.org/10.3390/ph18030402

AMA Style

Elshobary ME, Badawy NK, Ashraf Y, Zatioun AA, Masriya HH, Ammar MM, Mohamed NA, Mourad S, Assy AM. Combating Antibiotic Resistance: Mechanisms, Multidrug-Resistant Pathogens, and Novel Therapeutic Approaches: An Updated Review. Pharmaceuticals. 2025; 18(3):402. https://doi.org/10.3390/ph18030402

Chicago/Turabian Style

Elshobary, Mostafa E., Nadia K. Badawy, Yara Ashraf, Asmaa A. Zatioun, Hagar H. Masriya, Mohamed M. Ammar, Nourhan A. Mohamed, Sohaila Mourad, and Abdelrahman M. Assy. 2025. "Combating Antibiotic Resistance: Mechanisms, Multidrug-Resistant Pathogens, and Novel Therapeutic Approaches: An Updated Review" Pharmaceuticals 18, no. 3: 402. https://doi.org/10.3390/ph18030402

APA Style

Elshobary, M. E., Badawy, N. K., Ashraf, Y., Zatioun, A. A., Masriya, H. H., Ammar, M. M., Mohamed, N. A., Mourad, S., & Assy, A. M. (2025). Combating Antibiotic Resistance: Mechanisms, Multidrug-Resistant Pathogens, and Novel Therapeutic Approaches: An Updated Review. Pharmaceuticals, 18(3), 402. https://doi.org/10.3390/ph18030402

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