Next Article in Journal
AG1®, a Novel Synbiotic, Maintains Gut Barrier Function following Inflammatory Challenge in a Caco-2/THP1-Blue™ Co-Culture Model
Next Article in Special Issue
Antimicrobial Polymer Films with Grape Seed and Skin Extracts for Food Packaging
Previous Article in Journal
Green Roof Substrate Microbes Compose a Core Community of Stress-Tolerant Taxa
Previous Article in Special Issue
Current and Ongoing Developments in Targeting Clostridioides difficile Infection and Recurrence
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Mechanisms of Bacterial Resistance to Antimicrobial Peptides in the Modern Era: An Updated Review

1
Laboratory of Applied Biotechnology (LBA3B), Azm Center for Research in Biotechnology and Its Applications, Department of Cell Culture, EDST, Lebanese University, Tripoli 1300, Lebanon
2
CNRS, Architecture et Réactivité de l’ARN, UPR 9002, Université de Strasbourg, 2 Allée Konrad Roentgen, F-67000 Strasbourg, France
3
College of Engineering and Technology, American University of the Middle East, Egaila 54200, Kuwait
4
CNRS, INP, Inst Neurophysiopathol, Aix-Marseille Université, 13385 Marseille, France
5
Department of Biology, Faculty of Sciences 3, Lebanese University, Campus Michel Slayman Ras Maska, Tripoli 1352, Lebanon
6
Faculty of Medicine and Medical Sciences, University of Balamand, Kalhat, P.O. Box 100, Tripoli, Lebanon
*
Authors to whom correspondence should be addressed.
Microorganisms 2024, 12(7), 1259; https://doi.org/10.3390/microorganisms12071259
Submission received: 8 May 2024 / Revised: 10 June 2024 / Accepted: 18 June 2024 / Published: 21 June 2024
(This article belongs to the Special Issue Latest Review Papers in Antimicrobial Agents and Resistance 2024)

Abstract

:
Antimicrobial resistance (AMR) poses a serious global health concern, resulting in a significant number of deaths annually due to infections that are resistant to treatment. Amidst this crisis, antimicrobial peptides (AMPs) have emerged as promising alternatives to conventional antibiotics (ATBs). These cationic peptides, naturally produced by all kingdoms of life, play a crucial role in the innate immune system of multicellular organisms and in bacterial interspecies competition by exhibiting broad-spectrum activity against bacteria, fungi, viruses, and parasites. AMPs target bacterial pathogens through multiple mechanisms, most importantly by disrupting their membranes, leading to cell lysis. However, bacterial resistance to host AMPs has emerged due to a slow co-evolutionary process between microorganisms and their hosts. Alarmingly, the development of resistance to last-resort AMPs in the treatment of MDR infections, such as colistin, is attributed to the misuse of this peptide and the high rate of horizontal genetic transfer of the corresponding resistance genes. AMP-resistant bacteria employ diverse mechanisms, including but not limited to proteolytic degradation, extracellular trapping and inactivation, active efflux, as well as complex modifications in bacterial cell wall and membrane structures. This review comprehensively examines all constitutive and inducible molecular resistance mechanisms to AMPs supported by experimental evidence described to date in bacterial pathogens. We also explore the specificity of these mechanisms toward structurally diverse AMPs to broaden and enhance their potential in developing and applying them as therapeutics for MDR bacteria. Additionally, we provide insights into the significance of AMP resistance within the context of host–pathogen interactions.

1. Introduction

AMR claims the lives of around 700,000 people annually, solidifying its status as one of the most significant global health threats. Looking ahead to 2050, estimates suggest that deaths attributed to this issue could soar to 10 million annually, marking a staggering increase of roughly 14.3 times compared to the current levels [1]. AMR is also expected to raise the occurrence of diseases, amplify the seriousness of illnesses, escalate disability rates, and accordingly, become the world’s primary cause of death [2,3].
Many factors contribute to the spread of AMR, including the poor medication quality (such as counterfeit production), inappropriate prescription practices, patient misuse of antibiotics, inadequate surveillance systems and tests for AMR, and overuse of antibiotics in the industrial, animal, and agricultural sectors [4]. Wars and conflicts also play a crucial role in contributing to this grave situation [5]. Last, but not least, during the coronavirus disease 2019 pandemic, there was a surge in the incorrect overuse of antibiotics [6].
In light of this, AMPs have emerged as a solution to address this challenging situation [7], demonstrating their effectiveness in combating resistant microorganisms. AMPs primarily damage bacteria through a non-receptor-mediated membrane mode, enabling them to impede microbes’ ability to repair their damaged membranes [8] (for a detailed review on AMPs, see [9]). AMPs are ubiquitous across all life domains [10]. Currently, more than 3940 peptides have been documented in the AMP Database [11]. Upon administration, AMPs selectively target bacterial membranes, primarily due to the prevalence of anionic lipids, such as cardiolipin (CL) and phosphatidylglycerol (PG). Puzzlingly, the presence of CL has been observed to protect against AMP-mediated membrane disruption in some cases [12], whereas in other scenarios, membrane disruption appears to be intensified by CL [13]. The subsequent Section 3.5 will delve into further detail regarding this aspect.
The presence of zwitterionic lipids, like phosphatidylcholine, in mammalian cells reduces the affinity of cationic AMPs to their membranes [14]. However, AMPs can still interact with eukaryotic cell membranes to some extent and interact with intracellular organelles including mitochondria, which may have implications for their therapeutic efficacy and potential side effects [15,16]. CL is a major functional component of the inner mitochondrial membrane in eukaryotes [17]. Despite some similarities in composition and electric charge between the cell membrane and outer mitochondrial membrane, it remains uncertain if AMPs interact with the outer mitochondrial membrane in a manner similar to that with the cell or bacterial membrane. Nonetheless, it is known that, once inside cells, AMPs alter the permeability of both the outer and inner mitochondrial membranes, thus triggering the mitochondrial pathway. Consequently, AMPs that induce mitochondrial apoptosis often simultaneously modulate the permeability of both mitochondrial membranes in a synergistic manner [18].
AMPs offer several advantages over traditional ATBs [11,14,19,20,21,22,23] (Figure 1), primarily due to their capability to target the challenging non-permeable double membrane of Gram-negative bacteria [24].
Bacteria have developed resistance to AMPs [25], although this occurrence is believed to be currently less widespread and progresses at a slower pace compared to what is observed with traditional antibiotics, apart from the non-host peptides polymyxins [22,26,27]. Indeed, Peschel and Sahl (2006) proposed that immune defense strategies using cationic AMPs as weapons and bacterial resistance mechanisms have co-evolved over millions of years. This co-evolution toward specificity and complexity enhances the fitness of bacteria [22]. However, hurdles in evolutionary adaptation to resist AMPs stem from various factors, such as functional compatibility, fitness costs, and the relatively lower occurrence of AMP resistance genes associated with mobile genetic elements [28]. Despite being in the gut microbiome, AMP resistance genes exhibit a diminished capacity for horizontal transfer comparing to antibiotic resistance genes [29]. In terms of fitness tradeoffs, studies have revealed that the increased expression of mcr-1 gene, a plasmid-borne enzyme altering lipid A to confer acquired colistin resistance, results in many cellular alterations and reduced cell viability and growth rate [30]. The molecular mechanisms underlying acquired resistance through gene mutation(s) have been elucidated experimentally in several bacterial pathogens by isolating AMP-resistant strains following serial passages in the presence of gradually increasing AMP concentrations, or by directly exposing them to supra-minimum inhibitory concentration (MICs) of clinically relevant AMPs through plating (for a detailed review, see [31]). Therefore, the long-standing belief that acquiring resistance to AMPs is extremely difficult, and not a major concern [32,33,34,35] is mistaken, as such resistance can indeed develop at high rates at least under in vitro conditions. This results in mutants that do not entail significant loss of fitness [31], with sometimes significant levels of resistance granting cross-resistance to human AMPs inherent to innate immunity or those in clinical use, each with different structures and modes of action. Resistance to AMPs can manifest as constitutive (passive and intrinsic), with mechanisms rooted in the inherent properties of an organism that confer resistance even in the absence of AMPs. These properties, such as the components of bacterial membranes and cell walls, can be essential for the bacterial survival. Some bacteria like the Gram-negative species Neisseria spp., Proteus spp., Providencia spp., Morganella morganii, Serratia spp., Edwardsiella tarda, and Burkholderia cepacia complex (BCC) exhibit intrinsic resistance to polymyxins in vitro. This resistance is mainly due to the constitutive expression of lipopolysaccharides (LPSs) substituted with positively charged 4-amino-4-deoxy-L-arabinose (L-Ara4N), which electrostatically repulses polymyxins [36,37,38]. Such substitutions were shown to be essential for survial of Burkholderia cenocepacia since only conditional mutants in this pathway are viable. Furthermore, the members of BCC show exceptionally high resistance to polymyxin B and AMPs, with some isolates having MIC exceeding 512 μg/mL. This resistance is partly attributed to the synthesis of hopanoids (eukaryotic sterol analogs) and isoprenoids, which stabilize the inner membrane permeability, thus preventing the self-promoted uptake of polymyxin B. Active efflux pump expression and other mechanisms, not yet fully understood, also contribute to high polymyxin B resistance in Burkholderia [36,38,39].
On the other hand, pathogenic bacteria can modify their transcriptome in response to conditions encountered within their hosts [40], for an optimal adaptation to the presence of AMPs. Inducible (adaptive) resistance to AMPs is often governed by sophisticated systems, such as two-component regulatory systems (TCSs) [41,42,43,44], which are widespread among both commensal and pathogenic bacteria [45]. For example, in Salmonella, inducible resistance to AMPs is governed by the PhoPQ TCS, activated directly by sublethal concentrations of AMPs or indirectly by the stress caused by these peptides [46], as well as other stressors potentially encountered inside phagosomes, such as low magnesium concentrations or an acidic pH [41]. Additionally, PhoPQ activates the adaptor pmrD protein, which in turn activates the PmrAB TCS. This system is also induced in the presence of high Fe2+ concentrations to mediate various LPS modifications associated with resistance to cationic AMPs, including polymyxins [37,41,47].
Significant advances in structural biochemistry and microbial genetics have shed light on the structures of several AMPs and the molecular and mechanistic bases of bacterial resistance to these peptides. For example, Groisman et al. (1992) demonstrated that a single gene, phoP, encoding a transcriptional regulator, could render the intracellular pathogen Salmonella Typhimurium resistant to AMPs by modifying its outer membrane [48].
In this review, we thoroughly investigate and discuss the molecular mechanisms experimentally proven to confer AMP resistance in bacteria, whether constitutively or induced by environmental signals including AMPs themselves. Furthermore, we analyze their correlation with the virulence phenotype and pathogenicity determinants observed through in vitro cell culture bioassays and in animal models of infection.

2. Antimicrobial Peptides

AMPs are typically small molecules comprising fewer than 100 amino acid residues [49] and play a vital role in innate immunity, serving as the first line of defense against invading microorganisms within various host organisms [50]. AMPs are synthesized in different locations depending on the organism producing them. For example, mammalian AMPs are synthetized in mucosal tissues, glandular cells, the skin and colostrum, endothelial cells, and certain immune cells, such as neutrophils, macrophages, and dendritic cells [51].
AMPs can be classified according to seven criteria, including targets, sources, characteristics, biosynthesis, structures, covalent bonding forms (universal classification), and ultimately, their predominantly antibacterial functions [11,52] (Figure 2).
Roughly 88% of AMPs are positively charged and display hydrophobic characteristics [11]. Consequently, AMPs interact with the negatively charged components of bacterial membranes, thereby disrupting them and ultimately killing microbes [53]. These AMPs have the capability to target multiple sites on both the membrane and within the bacterial cell [54,55]. Eradicating bacteria by AMPs can be achieved directly or indirectly. In the direct mode, AMPs target the membrane, either by binding to cytoplasmic membrane receptors [56] or by acting without direct binding [57]. In the latter case, the membrane is disrupted non-specifically, leading to damage through pore formation [58] or the permeation of cytoplasmic membranes without pore formation [59]. The pore formation mode is further categorized into two forms: the Barrel–Stave pore model and the toroidal pore model, while the non-pore formation mode is exemplified by the carpet model [60]. Additionally, AMPs can directly target the cell walls [58] or the intracellular components (DNA, RNA, and proteins) of bacteria [61]. Alternatively, the indirect mode involves modulating the immune system [62].
While the advantages of AMPs surpass those of classical ATBs, they also pose challenges, such as restricted bioactivity, potential biotoxicity, delivery, and the risk of hemolysis, particularly due to imprecise damage and failure to distinguish between foreign bodies and host cells [63,64]. For instance, defensins and cathelicidins are the two prominent mammalian families of cationic AMPs with amphipathic properties. Defensins are characterized by their cysteine-rich nature and the formation of β-sheet structures held together by disulfide bonds [65]. On the other hand, cathelicidins lack disulfide bonds and instead adopt amphipathic α-helical structures. In humans and mice, only one cathelicidin member exists, known as LL-37 and murine cathelicidin-related antimicrobial peptide (mCRAMP), respectively. The positive charge of these AMPs facilitates their selective interaction with negatively charged bacterial membranes [65,66]. LL-37 and defensins are produced in the bone marrow and epithelial cells [67] and have been observed to interact with host cell receptors even at concentrations significantly lower than the MIC of these peptides. This interaction contributes to various illnesses, including pulmonary disorders, autoimmune diseases, tumors, and cardiovascular and neurodegenerative diseases [68].
Additionally, AMPs are costly, poorly stable, and have a short half-life due to environmental factors, such as protease activity (mainly pancreatic enzymes), temperature, salts, extreme pH conditions, and reduced permeation through gastrointestinal membrane, which makes their oral administration much difficult [69,70,71,72,73]. Consequently, more than 50% of identified AMPs undergo modifications depending on their administration route and delivery system [74,75]. In contrast to other administration routes, such as intramuscular or subcutaneous routes, the stability requirement for peptides may be less stringent. Therefore, injection stands out as the preferred route of administration for most AMPs [76]. However, intravenous administration exposes peptides to the esterase and peptidase activity found in serum [77,78]. Moreover, optimizing the drug dosage and minimizing systemic exposure can reduce systemic side effects when administering a drug via the lungs. Inhaled peptide medications have shown a superior efficacy in terms of rapid onset [79]. The most prevalent and well-developed applications of AMPs involve their incorporation into nanoparticles, hydrogels, creams, gels, and ointments for topical use [80]. Further research is necessary to explore new and suitable administration routes.
Natural peptides serve as a template for creating synthetic peptides with enhanced stability, increased bioactivity, reduced toxicity, and additional antimicrobial activities [81,82]. These synthetic peptides can be generated through various modifications, including chemical alterations, targeted peptide mutagenesis, nanoengineering techniques, capping, motif usage, coupling with photosensitizers, or combination with ATBs [83,84]. However, approaches to enhancement should be pursued with caution and precision, as even minor alterations can compromise effectiveness.
For instance, lipidation involves attaching fatty acids to the active regions of the peptides [85]. The lipid addition confers numerous advantages to AMPs without altering their fundamental characteristics, such as increasing hydrophobicity, enhancing membrane interaction, improving penetrability, reducing degradation (thus increasing stability), and preserving cationic charges [86,87,88]. Nevertheless, this mechanism may also have significant drawbacks. For example, the elongated fatty acid chain increases hydrophobicity, leading to a greater toxicity against mammalian cells [89]. Additionally, it promotes self-assembly in water, resulting in a reduced interaction between peptides and the membranes of their targets [90].
Notwithstanding their potent antimicrobial effects in vitro, only some AMPs derived from bacteria, including polymyxin B, polymyxin E (commonly known as colistin), and bacitracin, are currently used to treat Gram-negative infections in humans, while others have shown promise in preclinical studies and clinical trials [91,92,93] (Table 1). Despite concerns regarding toxicity and the emergence of resistance [94,95,96], colistin has been reintroduced for use as monotherapy or in combination as a last resort for treating MDR and extensively drug-resistant (XDR) Gram-negative bacteria isolated from nosocomial infections. These infections involve pathogens such as P. aeruginosa, Acinetobacter baumannii, K. pneumoniae, and E. coli [97,98]. Because of poor pharmacokinetics, the intravenous administration of colistin is typically restricted to treating urinary tract infections or employed as inhalation therapy [99,100]. Efforts to optimize its use, including dosage adjustment and combination therapy with other antibiotics, are ongoing to balance its effectiveness with the risk of adverse effects.

3. Mechanisms of Resistance to AMPs

In this section, we explore the mechanisms of AMP resistance (Figure 3, Figure 4 and Figure 5), emphasizing their correlation with the bacterial morphological and functional structures encountered by AMPs from their exposure in the bacterial environment to their interaction with and traversal across bacterial envelopes until reaching the cytoplasmic membrane or other intracellular targets [66]. To reach their targets, AMPs must interact with and traverse the enveloping structures, which can vary in chemical and physical properties among bacteria. Understanding these interactions is crucial for elucidating the effectiveness of AMPs and devising strategies to overcome bacterial resistance.
Table 2 and Table 3 summarize the mechanisms of resistance to AMPs described for Gram-negative and Gram-positive bacteria and their association with virulence and pathogenicity.

3.1. Extracellular Proteolytic Degradation

Many bacterial pathogens employ surface-secreted or surface-expressed proteases to recognize and cleave cationic AMPs, allowing them to evade their bactericidal effects [126] (Figure 5B). These enzymes can be divided according to their active sites in five classes: cysteine proteases, aspartic proteases, metalloproteases, serine proteases, and threonine proteases [127]. The simple linear form of the α-helical peptide LL-37 and other cationic AMPs renders them susceptible in vitro to various proteases produced by different pathogenic bacteria (Table 2 and Table 3) (Figure 5B). In particular, S. Typhimurium produces an endopeptidase in the outer membrane, named PgtE, which specifically degrades the linear cationic AMPs LL-37 and C18-G. The transcription or outer membrane localization of PgtE is regulated by an unknown post-transcriptional mechanism activated by PhoPQ [128]. PgtE exhibits structural characteristics consistent with those of the OmpT or VII proteases of E. coli [129], and Pla from Yersinia [130]. The ability of E. coli to produce PgtE and degrade protamine is associated with the rapid uptake of K+ ions by bacteria [131]. However, the PgtE protease does not enhance Salmonella resistance to AMPs that adopt amphipathic β-sheet conformations stabilized by intramolecular disulfide bridges (e.g., defensins and protegrins), as these AMPs are resistant to protease activity.
Table 2. Mechanisms of AMP resistance in Gram-negative bacteria and their correlation with virulence and pathogenicity.
Table 2. Mechanisms of AMP resistance in Gram-negative bacteria and their correlation with virulence and pathogenicity.
BacteriaMechanisms/Regulatory Pathways of ResistanceGene(s) InvolvedAMPsCorrelation with Pathogenicity/Virulence *Refs.
Acinetobacter
baumannii
Acylation of lipid A lpxLPolymyxin BUnknown[132]
LPS-full length-mediated protectionlpxA, lpxC, and lpxDLysozyme, LL-37, and lactoferrinBacteremia (mice)[133,134,135]
Deacylation of lipid A naxD (pmrB)Polymyxin BUnknown[136]
Hydroxylation of lipid A lpxO/pagQPolymyxin B, colistin, and HBD-3Bacteremia (Galleria mellonella)
Antiphagocytic
(invertebrates and mammalian cells)
[137]
Addition of PEtN to lipid Aept, mcr (plasmid-encoded),
(pmrAB and stkSR)
ColistinIncreased virulence in G. mellonella[138,139,140,141,142,143]
Uptake by/binding to porinsompA
ompW
Colistin (uptake)
LL-37, BMAP-28 (binding), and colistin
Virulence to the human airway epithelium,
adherence to cells, and biofilm formation
[144,145]
Active efflux (MFS and RND-types)emrB
adeABC
Colistin
Colistin heteroresistance
Overexpression increased virulence
in a pulmonary infection model
[146,147,148]
Manipulation of host AMP productionlpxOGaliomycin, gallerimycin,
and lysozyme
Bacteraemia (G. mellonella)[137]
Actinobacillus
pleuropneumoniae
Outer membrane permeabilityompW (soxS)Polymyxin BUnknown[149]
Active efflux (ABC family/K+ dependent)sapPR-39Respiratory tract infection (mice)[150]
Bordetella sp.O-antigen-mediated protectionwlbA and wlbL (bvgAS)Cecropin, magainin,
protamine, and melittin
Tracheal colonization (turkey)[151,152]
Acylation of lipid ApagP (bvgAS) and lpxL1C18GRespiratory tract infection (mice)
Infection of human macrophages
[153,154]
Brucella abortusDephosphorylation of lipid AlpxEPolymyxin BNot required[155]
Brucella melitensisActive efflux (ABC-type)yejABEFProtamine, melittin, polymyxins,
HBD-1, and HBD-2
Survival in macrophages (mice)[156]
Burkholderia
cenocepacia complex
Degradation by zinc metalloproteaseszmpB and zmpA (cepIR and cciIR)HBD-1 and LL-37, respectivelyChronic respiratory infection (mice)[157,158]
Inhibition by exopolysaccahrides
(mainly cepacian)
-CathelicidinsLung infections of cystic fibrosis (humans)[159]
Blockage of AMP uptake by the LPS-heptosylated core oligosaccharidewaaFPolymyxin B, melittin, and HNP-1Unknown[36,160]
Stabilization of the inner membrane lipidsispH (LytB; isoprenoid synthesis) and hpnJ (encodes hopanoid)Polymyxin BUnknown[36]
Protease-mediated protection
(unknown role/mechanism)
mucD (HtrA protease family)Polymyxin BUnknown[36]
Active efflux (MATE-type)norMPolymyxin BUnknown[161]
Addition of 4AraN to lipid AugdBCAL2946 (two ugd in B. caepacia)Polymyxin BUnknown (mutants of
4AraN synthesis not viable)
[160]
Alternative sigma factor regulon (37 °C)rpoEPolymyxin BPhagolysosomal fusion in macrophages[36]
Campylobacter jejuniLOS-heptosylated core-mediated protectionwaaFPolymyxin B, HNPs, LL-37, and BPIInvasion of INT407 cells in vitro[162,163]
LOS core-mediated protectiongalUPolymyxin B, colistin, magainin, cecropin, and bacitracinUnknown[164]
Active efflux (RND-type)cmePolymyxin BIntracellular survival and multiplication (Acanthamoeba polyphaga)[165,166]
Capnocytophaga
canimorsus
Dephosphorylation of lipid AlpxEPolymyxin BUnknown[167]
Enterobacter cloacae complexActive efflux (RND-type)acrAB-tolC (soxSR)
kexD/(crrC)
Polymyxin B and colistinSystemic infection
(intraperitoneal mouse model)
[168,169,170]
Addition of 4AraN to lipid Aarn (phoPQ and mgrB)Colistin heteroresistanceUnknown[171,172]
Addition of PEtN to lipid AmcrColistinUnknown[173]
Potential efflux mechanism
mediated by an inner membrane protein
dedA (ecl)Colistin heteroresistanceUnknown[169]
Erwinia chrysanthemiActive efflux (ABC family/K+-dependent)sapα-thionin and anakinUnknown[174]
Escherichia coliProtease-mediated degradation degPLactoferrinUrovirulence (humans)[175]
ompT and degPProtamine, C18G, and LL-37[131,176,177,178]
Core oligosaccharide-mediated protectionpmrDPolymyxin BUnknown[179]
Acylation of lipid ApagP and lpxM (pmrAB and mgrB)LL-37Unknown[180,181]
Addition of 4AraN to lipid Aarn (phoPQ, pmrAB, and mgrB)Polymyxin B and colistinUnknown[182,183,184]
Addition of PEtN to lipid AeptA, eptB, eptC, and mcr
(phoPQ, pmrAB, and mgrB)
Polymyxin B and colistinUnknown[184,185,186]
Decreased entry via porinsompFColicin and P6Unknown[187,188]
Peptidoglycan modificationamiA and amiC (cpxRA and nlpE)Protamine, magainin, and melittinUnknown[189]
Active efflux (ABC-type)macBBacitracin and colistinUnknown[190]
Active efflux (RND and MFS-types)acrAB and emrAB (cpxRA)ProtamineUnknown[191]
Transcriptional repression of
host’s AMP production (ETEC)
elt (heat-labile toxin-encoding gene)HBD-1 and LL-37Downregulation of kinase A, ERK MAP Kinase, and Cox-2 pathways
(intestinal epithelial cells)
[192]
Francisella novicidaDephosphorylation of lipid AlpxFPolymyxin BPulmonary and
subcutaneous infections (mice)
[193,194]
Francisella tularensisDeacylation of lipid AnaxDPolymyxin BIntracellular replication (mice)[195]
Haemophilus ducreyiActive efflux (ABC family/K+ dependent)sapALL-37Chancroid (human)[196]
Active efflux (RND-type)mtrLL-37 and β-defensinsUnknown[197]
Haemophilus influenzaeAcylation of lipid AlpxL (htrB) Polymyxin BColonization of human airway epithelial cells[198,199]
Active efflux (ABC family/K+-dependent)sapACanine β-defensin-1, HBD-1, -2, -3,
LL-37, HNP-1, and melittin
Otitis media (chinchillas)[200]
Helicobacter pyloriDephosphorylation of lipid AlpxE and lpxFPolymyxin B, LL-37, HBD-2, and P-113Gastrointestinal infection (mice)[201]
Addition of PEtN to lipid AeptAPolymyxin BUnknown[202]
O and N-acetylations of peptidoglycan patA and pgdA (synergistic) Lysozyme Stomach colonization (mice)[203]
Klebsiella pneumoniaeCapsule-mediated protectioncpsPolymyxin B and lactoferrinPulmonary infection (mice)[204]
O-antigen-mediated protectionwcaI, cpsB, wcaJ, and cpsGHistonesUnknown[205]
Acylation of lipid AlpxM
lpxL2 and pagP
(mgrB and crrab)
Polymyxin B, colistin, CP28, and C18G Polymyxins (lpxL2, pagP, crrab)
HNP-1, HBD-1, -2, and -3 (mgrB)
Antiphagocytic, limits the activation of inflammatory responses by macrophages, and survival (G. mellonella); pneumonia (mice) [143,206,207]
Hydroxylation of lipid AlpxO (phoPQ)PolymyxinsPulmonary infection (mice)[208]
Addition of 4AraN to lipid ApmrHFIJKLM
(phoPQ, pmrAB, and mgrB)
ColistinSame phenotypes as for lpxM
mutant in G. mellonella
[183,184,209,210]
Addition of PEtN to lipid AeptA, eptB, eptC, and mcr,
(phoPQ, pmrAB, and mgrB)
ColistinUnknown[184,209,210,211]
Activation of unknown systems
dedicated to ameliorating AMP cytotoxicity
ompAPolymyxin B and protaminePulmonary infection (murine)[212]
Active efflux (ABC family/K+-dependent)sapALL-37Systemic infection (mice)[213]
Active efflux (RND-type)acrRAB
H239_3064 (crrAB)
Polymyxin B, HNP-1,
HBD-1 and HBD-2, and colistin
Pneumonia (mice)[214,215]
Legionella pneumophilaAcylation of lipid Arcp
pagP (phoPQ)
LL-37
Polymyxin B and C18G
Pulmonary colonization and infection (mice)[216]
Neisseria gonorrhoeaeInhibition by lactoferrin-binding protein BlbpBLactoferricinUnknown[217,218]
Inhibition by type IV pili adhesins
(Same scenario in N. meningitides, involvement of host cell RhoA and Cdc42 signalling)
pilELL-37Adherence to human epithelial cells[219,220]
Active efflux (RND-type)mtrLL-37, PG-1, PC-8,
polymyxin B, and colistin
Genital tract infection (mice)[221,222,223,224]
Neisseria meningitidisInhibition by lactoferrin-binding protein BlbpB (nalP)LactoferricinUnknown[217,218,225]
Sequestration/shielding by capsulecpsHBD-1 and -2, HNP-1 and -2, LL-37, CRAMP, PG-1, and polymyxin BMeningitis (humans)[226,227]
Sequestering by blebs from
the OM and biofilm formation
-Cationic AMPsUnknown[228]
Addition of PEtN to lipid A (constitutive)
(Same scenario in N. gonorrhoeae)
eptA (formerly lptA) (misRS)
dsbA
Polymyxin BColonization, inflammation, and survival in neutrophils (EptA)[229,230]
Porin-mediated exportporBPolymyxin BUnknown[229]
Active efflux (RND-type)mtr (constitutive) Polymyxin B, PG, and LL-37Unknown[229]
Photorhabdus laumondiiActive efflux (RND-type)acrABPolymyxin B and colistinSlight effect on virulence (insects)[231,232]
Addition of 4AraNpbgPE
(phoPQ)
Polymyxin B, colistin,
cecropins A and C
Septicemia and virulence (insects)[233,234]
Porphyromonas
gingivalis
Dephosphorylation of lipid AlpxFPolymyxin BUnknown[233,234,235]
Outer membrane OmpA-like
porins (undefined mechanism)
pgm6 and pgm7HBD-1 and -3, and LL-37Unknown[236]
Inactivation by proteases RIA, RIB, and KgpprpR1 and kgpCecropin B, brevinin, cecropin A 1-7, melittin 2-9, and mastoparanUnknown[237]
Proteolytic degradation by
gingipains (serine proteases)
rgpABCecropin B Abscess formation (mice)
Prevotella sp.Proteolytic degradationUnknownCecropin B and brevininUnknown[237]
Proteus mirabilisDegradation by metalloproteasezapALL-37Urinary tract infection (mice)[238]
Addition of 4AraN to lipid ApmrABPolymyxin BUnknown[239]
Active efflux (ABC family/K+-dependent)sapProtegrinUnknown[239]
Pseudomonas
aeruginosa
Degradation by elastaselasLL-37Corneal infection (mice)[240]
Capsule-mediated protectioncpsPolymyxin BResistance to neutrophil-mediated killing[241]
Shedding of host proteoglycanslasALL-37 and human α-defensinsPulmonary infection (mice)[242,243]
Hydroxylation of lipid AlpxOPolymyxin BAcquisition of loss-of-function mutations during chronic CF lung infection (mice)[244,245]
Addition of 4AraN to lipid ApmrHFIJKLM (pmrAB, phoPQ,
parRS, colRS, and cpsRS)
Colistin and polymyxin BCystic fibrosis (humans)[246,247,248]
Addition of PEtN to lipid AeptA (only by ectopic expression in L-Ara4N-defective mutants) and mcrColistinUnknown[249,250,251]
Alteration of membrane
phospholipid composition
PA0920ProtamineUnknown[252,253]
Active efflux (RND family)mexAB, mexCD, and mexXYColistin and polymyxin BControversial and opposing roles
since some mutations increase virulence
[254,255,256]
Stimulation of host cathepsinsUnknownLL-37, HBD-1, and HBD-3Cystic fibrosis (humans)[257,258]
Pseudomonas
fluorescens
Alteration of cytoplasmic
membrane lipid composition
UnknownPolymyxin BUnknown[259]
Rhizobium etliDephosphorylation of lipid AlpxE and lpxFPolymyxin BUnknown[167]
Rhizobium
leguminosarum
Salmonella entericaEndopeptidase-mediated degradationpgtE (phoPQ)LL-37 and C18-GUnknown[128]
Acylation of lipid AlpxM (msbB)
pagP (phoPQ)
Polymyxin B
C18G and PG-1
Inflammation and septaecemia in mice (msbB), minor effects (pagP),
Required for full virulence (phoP)
[182,260]
Dephosphorylation of lipid ApagL and lpxR (phoPQ)Polymyxin BMinor effects in mice (pagL and lpxR)[261]
Hydroxylation of lipid AlpxO (phoPQ)LL-37Increased invasion of human epithelial cells and full virulence in animals (lpxO mutant)[262]
Addition of 4AraN to lipid ApmrHFIJKLM (phoPQ)Defensins and polymyxin BGastrointestinal infection (mice)[260,263]
Addition of PEtN to lipid A/core oligosaccharideeptA, eptBC (core), cptA (core) (pmrAB), mcrPolymyxin B and colistinUnknown[185,264]
Peptidoglycan modificationamiA and amiC (cpxRA and nlpE)Protamine, magainin, and melittinUnknown[189]
Active efflux (ABC transporter)macABC18GIntracellular survival (macrophages)[265,266]
Active efflux (ABC family/K+-dependent)sap
yejABEF
Protamine, melittin, and polymyxin BGastrointestinal infection (mice)[48,267,268]
Extracytoplasmic σE factorrpoEP2, polymyxin B, and murine
α-defensin cryptdin 4
Gastrointestinal infection (mice)[269,270]
Shigella dysenteriaeManipulation of host AMP productionUnknownLL-37Bacillary dysentery (human)[271]
Shigella flexneriO-antigen-mediated protectionUnknownHistonesUnknown[205,272]
Alteration of host AMP productionospF (mxiE)Rabbit α-defensin NP5Repression of NF-kB-responsive genes
(Caco-2 and HeLa cells, rabbits)
[273]
Ureaplasma parvumHost chromatin alterationsUnknownHBD-1, HNP-6, and LL-37Decreased histone H3K9 acetylation
(Human THP-1 monocytoid tumor cell line)
[274]
Vibrio choleraeAcylation of lipid AlpxLPolymyxin BUnknown[275]
Hydroxylation of lipid AlpxNPolymyxin BUnknown[275]
Sensing by OMP/Activation
of a DegS-dependent σE factor
ompU/degS/rpoEPolymyxin B and P2Bile resistance in vivo (ompU)[276,277]
Active efflux (RND-type)vexAB, vexCD, and vexIJKPolymyxin BSmall intestine colonization (mice)[278]
Transcriptional repression of
host’s AMP production
ctxA and ctxB
(cholera toxin-encoding genes)
LL-37Downregulation of kinase A,
ERK MAP Kinase, and Cox-2 pathways
(intestinal epithelial cells)
[192]
Vibrio vulnificusK+ uptake transporter systemtrkAProtamine and polymyxin BSepticemia (mice)[279]
Yersinia enterocoliticaAcylation of lipid AlpxP and htrBPolymyxin BGastrointestinal infection (mice)[280]
OMP-mediated protectionyadA (pYVe plasmid-encoded)Lysozyme and defensins
from human granulocytes
Adhesion, autoaggregation, resistance to complement-mediated killing[281]
Active efflux (MFS-type)rosABPolymyxin B, cecropin, and melittinUnknown[282]
Yersinia pestisDegradation by aspartate protease plaLL-37, rCRAMP, and rat β-defensin-1Plague (mice)[130]
Active efflux (RND-type)acrABPolymyxin BNot required[283]
Yersinia
pseudotuberculosis
Acylation of lipid ApagPPolymyxin B and cecropinUnknown[284]
* Roles in virulence and pathogenicity have been demonstrated in animal models, in vitro cell cultures and bioassays, and ex-vivo studies, either performed within the same study or independently. These roles may not conclusively be linked to antimicrobial peptide (AMP) resistance per se. The underlined genes are regulatory genes. Abbreviations: BMAP-28: bovine myeloid antimicrobial peptide (cathelicidin family), Blebs: vesicles containing DNA, LOS, and OMPs; BPI: bactericidal/permeability-increasing Protein; cBD-1: caveolin-1-binding domain; CRAMP: cathelicidin-related antimicrobial peptide C18G: synthetic α-helical peptide derived from human platelet factor I; ETEC: enterotoxigenic Escherichia coli; EPSs: extracellular polymeric substances; HBD: human β-defensin; HNP: human neutrophil peptide; LBP-B: lactoferrin-binding protein B; LPS: lipopolysaccharide; LOS: lipooligosaccharide; mCRAMP: murine cathelicidin; OM: outer membrane; OMP: outer membrane protein; P-113: AMP derived from the human salivary protein histatin 5; P2 and P6: bioactive peptide fragments; PC-8: linearized synthetic variant lacking both disulfide bonds; PG-1: protegrin-1 (porcine); PR-39: a proline-rich antibacterial peptide; RIA and RIB: enzymes arising by the differential processing of the prpR1 arginine-specific protease; rCRAMP: rat cathelicidin.
Table 3. Mechanisms of AMP resistance in Gram-positive bacteria and their correlation with virulence and pathogenicity.
Table 3. Mechanisms of AMP resistance in Gram-positive bacteria and their correlation with virulence and pathogenicity.
BacteriaMechanisms/Regulatory Pathways of ResistanceGene(s) InvolvedAMPsCorrelation with Pathogenicity/Virulence *Refs.
Bacillus anthracisDegradation by metalloproteasesclpXLL-37, α-defensins, and lysozymeLethal infection in CRAMP −/− mice[285,286]
Capsule-mediated protectioncapADefensins, gramicidin, polymyxin B, nisin, protegrin, and melittinDissemination of inhalation anthrax infection (guinea pig)[25,287,288]
D-alanylation of TAsdltPolymyxin B, colistin, nisin,
and magainin-2
Survival in macrophages and full virulence in a mouse model of inhalational infection[289]
Lysinylation of PGmprFProtamine, LL-37, and HNP-1Unknown[290]
Bacillus cereusD-alanylation of TAsdltProtamine, nisin, polymyxin B, colistin, lysozyme, and cecropin BSeptecaemia and virulence (insects)[291]
Proteolytic degradation by zinc metalloproteasesinhA1 and inhA2Cecropin and attacinEscape from host macrophages
Lethal infection by injection into insects
[292,293]
Bacillus subtilisD-alanylation of TAsdlt (spoO and abrB)NisinUnknown[294,295]
Alteration of cytoplasmic membrane lipid compositionsigXNisinUnknown[294]
Clostridium difficileD-alanylation of TAsdltNisin, polymyxin B, and galliderminUnknown[296]
Enterococcus faecalisDegradation by gelatinasegelELL-37 and HYL-20Peritonitis (mice)[240,297]
Degradation by serine proteasessprEHYL-20Peritonitis (mice)[298,299]
Shedding of host proteoglycans
and neutralization of AMPs
Undefined (probably gelE)Neutrophil-derived α-defensinsUnknown[300]
D-alanylation of TAsdltColistin, nisin, and polymyxin BUnknown[301]
Lysinylation of phospholipidsmprF1 and mprF2Defensins and daptomycinBacteremia (mice)[302,303,304]
Alteration of the localization of cardiolipin microdomainsliaRDaptomycin and telavancinUnknown[305]
O-acetylation of peptidoglycanEF_0783LysozymeSurvival in peritoneal macrophages (mice)[306]
Group A streptococcusDegradation by cysteine-proteinasesspeB/ideS (covRS also known as csrRS)LL-37SpeB highly expressed in vivo and colocalizes with LL-37 in human tissue samples[240,307]
Capsule (hyaluronic acid)-mediated repellinghasABC (covRS)LL-37Survival in neutrophil extracellular traps[308]
Secreted and surface-bound inhibitory proteinsemm1 (Fimbrial M1 proteins)LL-37Skin or systemic infection (mice)[309]
ska streptokinase LL-37 and other cationic AMPsSystemic dissemination and virulence (mice)[310,311]
sicLL-37 and defensinsSkin infection (mice)[312]
Shedding of host proteoglycans that bind cationic AMPslasA and speBLL-37 and defensinsSkin infection (mice)[300,313]
Cleavage by GRAB:SpeB complexspeB
grab
LL-37Skin infection (mice)[314]
Regulatory systems sensing and inducing AMP resistancecovRSLL-37In vivo induction by LL-37[315,316]
crgRmCRAMPCompetitive advantage (mice)[317]
D-alanylation of TAsdltLL-37, polymyxin B, and lysozymeResistance to neutrophil killing, adhesion, and invasion (pharyngeal epithelial cells)[318]
Active efflux (ABC-type)salYSalA and SalA1 lantibioticsIntramacrophage survival (zebrafish)[319]
Manipulation of host AMP productionUnknownDefensinsUnknown[320]
Group B streptococcusTCS regulatory pathwaysliaR and covRSPolymyxin B, colistin, nisin, and LL-37Sepsis and pneumonia (mice)[321,322]
ciaRmCRAMP and lyzozymeIntracellular survival within neutrophils,
murine macrophages, and human brain
microvascular endothelial cells
[322]
Competitive binding/inactivation by PBP1aponA (liar)LL-37, CRAMP, and defensinsAntiphacocytic
Pulmonary infection and sepsis (rats)
[323,324]
Sequestration by pilipilB (liar)LL-37 and mCRAMPInvasion and paracellular translocation mediating resistance to phagocytic killing and virulence (humans and animal models)[325,326]
D-alanylation of TAsdlt (dltSR)ColistinPulmonary or systemic infection (rats)[327,328]
Listeria
monocytogenes
N-deacetylation of peptidoglycanpgdALysozymeVirulence after oral and IV inoculations (mice)
Survival in macrophages, liver,
spleen, and intestinal lumen
[329]
O-acetylation of peptidoglycanoatLysozymeVirulence, survival in macrophages, and control of cytokine production (mice)[330]
Glycosylation of WTAsgttA, gltB, and rmlLL-37Intestinal epithelium colonization (mice)[331,332]
D-alanylation of TAsdlt (virR)Colistin, polymyxin B, and nisinBlood infection (mice)[333,334]
Lysinylation of PGmprF (virR)Gallidermin, HNP-1 and -3Survival in liver and spleen (mice)[334,335]
Active efflux (ABC-type)anrAB (virR and rpoN)Nisin, bacitracin, and galliderminUnknown[336]
Thermoregulated transcription factorprfAPotato defensin (thermo-dependent)Gastrointestinal infection (mice)[337]
Mycobacterium
tuberculosis
Lysinylation of PGlysXHNP-1 and lysozymeRespiratory infection (mice and pig)[338]
Mycobacterium marinumMycolic acid-mediated protectionkasBHNP-1 and lysozymeIntramacrophage survival[339]
Staphylococcus
aureus
Degradation by metalloprotease aureolysinaurHaloganan and LL-37Not required[340,341,342]
Degradation by V8
glutamylendopeptidase (serine protease)
sspALL-37Virulence and in vivo growth
(murine abscess models)
[341,343,344]
Inhibition by iron regulated surface determinant AisdALactoferrin and LL-37Initial stage of abscess formation
after IV infection (mice)
[345,346,347]
Inhibition by staphylokinase
(plasminogen activator protein)
Sak
(agr)
Lactoferricin, tritrpticin, and human
defensins
Establishment of skin infections
(humans and mice models)
[348,349,350,351]
O-acetylation of peptidoglycanoatHost lysozymeSeptic arthritis (mice)
Anti-inflammation
Inhibits the polarization of T-helper cells (mice)
[352,353,354,355,356,357]
WTA-mediated protectiontagOGroup IIA phospholipase A2 and
HBD-3
Induction and progression of endovascular infection (rabbit model of infective endocarditis), adherence to human epithelial cells biofilm formation, colony spreading, and virulence in mammals[358,359,360,361,362,363]
D-alanylation of TAsdlt (agr)HNP1–3, gallidermin, protegrins 3 and 5, tachyplesins-1 and 3, magainin-2, nisin, and tPMP-131Sepsis and septic arthritis (mice)[364,365,366,367,368,369]
Alteration of cytoplasmic membrane lipid compositionpgsA and cls2DaptomycinUnknown[370]
Lysinylation of PGmprF/lysSDefensinsSystemic infection (mice)[371,372,373]
Active efflux (ABC-type)pmtABCDDefensins and LL-37Skin infection (mice)[374]
Plasmid-mediated active efflux (MFS)qacAtPMPEndovascular infections (rabbit)[375]
TCS inducing AMP resistancegraRS and apsSX/apsRHBD3, nisin, indolicidin, and LL-37Kidney infections
(peritoneal infection murine model)
[367,368,376]
Active efflux (ABC-type)vraFGHBD3, nisin, indolicidin, and LL-37Hemolytic activity,
expansion of subcutaneous abscesses
[368,376,377]
Staphylococcus
epidermidis
Sequestration/inhibition by EPS: PIA also known as PNAGicaHBD-3, LL-37, and anionic dermcidinResistance to PMN killing (humans)[378]
AMP-inducible three component systemsapsXRS
(via dlt, mprF, and vraFG)
HBD3Resistance to PMN killing (humans)[369,379]
Staphylococcus
xylosus
D-alanylation of TAsdltGallidermin, magainin 2, and nisinUnknown[365]
Streptococcus iniaeShielding/ inhibition by polyanionic
surface capsule and cell wall structures
pgm (phosphoglucomutase)Moronecidin and mCRAMPMeningoencephalitis (hybrid striped bass model)[380]
O-acetylation of peptidoglycancpsY and oatA (metR/mtaR)LysozymeSurvival in neutrophils[381]
Streptococcus
mutans
Active efflux (ABC-type)bceABRS (formerly mbrABCD)Bacitracin and human α- or
β-defensins (also induce bce)
Unknown[382]
D-alanylation of TAs (planktonic cells and biofilm)dlt (ciaR) HBD-1, HBD-2, HBD-3, and LL37Regulation of cariogenic virulence[383,384]
Streptococcus
pneumoniae
Inhibition by pneumococcal surface protein ApspAApolactoferrinPneumococcal infection (mice)[385]
Neutralization by free anionic capsular polysaccharidecpsPolymyxin B and HNP-1Unknown[241]
D-alanylation of TAsdlt (ciaRH)Nisin and galliderminCompetitive advantage in murine
model of pneumococcal pneumonia
[386,387,388]
Active efflux (ABC-type)macAB homologBacitracin, LL-37, and nisinUnknown[389,390]
Active efflux (MFS-type)mefEDefensins, LL-37, and CRAMPUnknown[389,391]
N and O-acetylations of peptidoglycanpgdA and adrLysozymeColonization of the upper respiratory tract (mice)[392]
Streptococcus suisDegradation by cysteine proteaseapdSLL-37Meningitis and sepsis (humans)[393]
O-acetylation of peptidoglycanoatLysozymeUnknown[394]
* Roles in virulence and pathogenicity have been demonstrated in animal models, in vitro cell cultures and bioassays, and ex-vivo studies, either performed within the same study or independently. These roles may not conclusively be linked to antimicrobial peptide (AMP) resistance per se. The underlined genes are regulatory genes. Abbreviations: CRAMP: cathelicidin-related antimicrobial peptide; ETEC: enterotoxigenic Escherichia coli; EPSs: extracellular polymeric substances; GRAB: α2-macroglobulin/G-related α2-macroglobulin-binding protein; HBD: human β-defensin; HNP: human neutrophil peptide; HYL-20: peptide derived from the venom of the solitary bee Hylaeus signatus; mCRAMP: murine cathelicidin; PBP1a: penicillin-binding protein 1a; PG: phosphatidylglycerol; PIA: polysaccharide intercellular adhesin also known as Poly N-acetylglucosamine (PNAG); TA: teichoic acid; WTA: wall teichoic acid.

3.2. Extracellular Trapping and Inactivation

Given the limited efficacy of proteases against defensins and other AMPs, numerous bacterial pathogens have devised supplementary strategies to impede their function (Table 2 and Table 3). Many of these tactics are tailored to specific peptides, encompassing processes such as recognition and extracellular capture. The direct capture or neutralization of AMPs can occur through the action of secreted proteins or those affiliated with the bacterial surface, while indirect methods involve host cells releasing surface molecules capable of binding PAMs (Table 2 and Table 3). Consequently, when a substantial portion of AMPs released during the innate immune response hampers these molecules, inadequate quantities reach the bacterial cell membrane.

3.2.1. AMP-Inhibitory Proteins Associated with the Bacterial Cell Surface

The sequestering of AMPs by certain molecules prevent them from interacting with their targets (Figure 5C). For example, SGA generate fimbrial M proteins, which steadfastly adhere to the cell wall, enveloping themselves in mucoid layers. Lauth et al. (2004) revealed that virulent strains producing the M1 serotype of this protein display a notably heightened resistance to cathelicidin LL-37 compared to their less virulent counterparts (housing M proteins of alternative serotypes) [33]. This protein is capable of binding to the cathelicidin precursor (hCAP-18), inhibiting its transformation into a mature AMP, and it can also bind to LL-37, thereby evading its effect [395].
Streptococci and staphylococci produce a variety of surface proteins able to degrade cationic AMPs (Table 3). For example, S. aureus expresses the iron-regulated surface determinant A (IsdA), located on the peptidoglycan layer, which is responsible for resistance to lactoferrin and mCRAMP. This adhesin, highly expressed during iron deficiency, confers resistance to AMPs via its C-terminal domain, which binds to apolactoferrin and suppresses its protease activity. Notably, IsdA does not necessitate attachment to the bacterial surface to manifest its proteolytic activity [345,346].
Neisseria gonorrhoeae and N. meningitidis produce a membrane-bound lipoprotein known as lactoferrin-binding protein B (LbpB) (Table 2). This protein contains anionic domains capable of conferring resistance to lactoferricin B, derived from host lactoferrin, an iron-sequestering antimicrobial protein released by neutrophils [217,218]. Recent structural studies have revealed that the N-lobe of LbpB harbors the binding site for lactoferrin. While LbpB may not be essential for iron acquisition in vivo, it plays a crucial role in protecting against cationic AMPs when membrane-bound. The enzymatic release of N. meningitidis LbpB from the cell surface by the bacterial protease NalP is hypothesized to contribute to LbpB’s function as a “Cationic AMP sink” [225,396]. Studies have also shown that the anionic regions of LbpB from Moraxella catarrhalis protect bacteria against cationic AMPs [397]. Additionally, LbpB-mediated protection against cationic AMPs is highly specific, as it provides resistance to the mCRAMP and certain synthetic peptides (IDR-1002 and IDR-0018) but not others (Tritrp 1 or LL37, HH-2, or HHC10) [397]. Recently, Ostan et al. (2024) demonstrated that the addition of lactoferricin to the anionic domain of LbpB or full-length LbpB secreted from the cattle pathogen Moraxella bovis, which expresses the largest anionic domain of the LbpB homologs, results in the formation of phase-separated droplets of LbpB together with the AMP. The droplets displayed a low rate of diffusion, suggesting that cationic AMPs become trapped inside and are no longer able to kill bacteria. Authors suggested that pathogens, like M. bovis, leverage anionic intrinsically disordered domains for the broad recognition and neutralization of antimicrobials via the formation of biomolecular condensates [398] (Figure 5C).
Finally, N. gonorrhoeae expressing type IV pili demonstrates resistance against LL-37 and neutrophil-mediated killing. However, non-piliated strains exhibit heightened susceptibility, potentially due to elevated intracellular iron levels. Iron chelation effectively protects non-piliated mutants from hydrogen peroxide and LL-37-mediated killing, indicating a role for iron availability in modulating bacterial sensitivity [219]. Therefore, the above-mentioned observations emphasize a critical relationship between piliation status, iron homeostasis, and susceptibility to host-derived AMPs (Figure 5E).

3.2.2. AMP-Inhibitory Proteins Secreted by Bacteria

S. aureus secretes a plasminogen activator protein called staphylokinase (Sak), which has the ability to bind to certain AMPs and inhibit their activities. For example, Sak, especially the truncated Sak with its first ten residues, exhibits a lower affinity for melittin and mCRAMP, but shows a high affinity for lactoferricin and tritrpticin [350]. In addition, Sak can bind to human β-defensins and restrict their bactericidal activity [349]. Similarly, GAS hijacks host plasmin through the secreted plasminogen activator streptokinase (Ska) to evade the lethal effects of cathelicidin, facilitating their evasion from innate immunity [310]. Moreover, the virulent strains of GAS M1 secrete a protein known as SIC (Streptococcal Inhibitor of Complement), which is absent in the vast majority of non-invasive GAS strains. SIC is capable of binding to and inactivating LL-37 and human α-defensin [312] (Table 3) (Figure 5D).

3.2.3. AMP-Inhibitory Molecules Released by Host Tissues

Bacteria appear to trap and deactivate PAMs by exploiting the negatively charged proteoglycan molecules that decorate the surfaces of host epithelial cells. For instance, extracellular proteases, such as cysteine proteinases, gelatinases, elastases, and alkaline proteinases, secreted by bacteria such as S. pyogenes, E. faecalis, and Pseudomonas aeruginosa, respectively, degrade human cell surface proteoglycans, such as decorin, then releasing dermatan sulfate (Table 2 and Table 3). The released dermatan sulfate binds to and inactivates human α-defensins [300]. In addition, P. aeruginosa produces elastase LasA, which increases the release of syndecan-1, a cell-surface heparan sulphate proteoglycan, from host cells. Consequently, syndecan-1 binds to and inactivate cationic AMPs [242] (Table 2 and Table 3).

3.3. Electrostatic Shielding/Sequestration of AMPs by the Capsule

The production of capsule polysaccharides (CPSs) is involved in adjusting biofilm formation to cause persistent infections in the blood, respiratory tract, and gastrointestinal mucosa of mammals [399,400]. Additionally, it can either confer resistance or reduce susceptibility to several AMPs and evade phagocytosis [401]. While the chemical composition of CPSs varies widely among bacterial species, the majority of them are anionic. The interaction with the bacterial capsule is believed to trigger conformational changes that lead to the sequestration of AMPs, consequently inhibiting their capacity to reach their target on the bacterial membrane [402].
The role of CPSs as a physical barrier in resistance to AMPs has been extensively studied, as observed in K. pneumoniae, S. pyogenes TX72 [204], and N. meningitidis [226], whose CPS mutants display increased sensitivity to AMPs compared to wild-type (WT) strains (Table 2 and Table 3). Furthermore, the degree of K. pneumoniae resistance to polymyxin B and lactoferrin is proportional to the amount of CPSs produced as a result of cps gene induction by these two AMPs [204]. Purified CPSs from K. pneumoniae, S. pneumoniae, and P. aeruginosa effectively sequester cationic AMPs and neutralize their bactericidal activity. The release of CPSs into the milieu increases significantly when bacteria are exposed to polymyxin B and human neutrophil peptide-1 (HNP-1) [241]. Additionally, O’Brien et al. elucidated that non-capsulated Bacillus anthracis strains are markedly more susceptible to various AMPs, compared to their encapsulated counterparts [25] (Table 3) (Figure 5F).

3.4. AMP Resistance Mechanisms Associated with Bacterial Cell Wall Structures and Their Modifications

The bacterial cell wall, primarily composed of peptidoglycan, is an important target for several antimicrobial agents. It is found outside the bacterial cell membrane, serves as a protective barrier against environmental stresses, and reinforces cell shapes [403]. AMPs face the initial challenge of penetrating the bacterial cell envelopes, which varies between Gram-negative and Gram-positive bacteria. In Gram-negative bacteria, this barrier includes an outer membrane in addition to a thin peptidoglycan layer and the cytoplasmic membrane, with the outer membrane being composed of phospholipids, lipopolysaccharides (LPSs), and proteins. Conversely, Gram-positive bacteria lack an outer membrane but possess a thicker peptidoglycan layer to which teichoic acids (TAs) and lipoteichoic acids (LTAs) are anchored [404]. Hence, most bacterial resistance mechanisms operate by altering the chemical components of these surface structures to restrict the interaction with AMPs.

3.4.1. Involvement of LPS and Its Modifications in Resistance to AMPs

The outer membrane plays a crucial role in the Gram-negative bacterial natural defense by exhibiting reduced permeability, and regulating its permeability and integrity is pivotal [405,406]. The principal component of the OM is the negatively charged LPS, which attracts cationic AMPs, while outer membrane proteins (OMPs) also contribute to this resistance through their physiological roles, as discussed in Section 3.4.2. While bacterial plasma membranes share the same phospholipid composition, certain AMPs exhibit effectiveness against specific bacterial strains but not others [407]. This difference in activity highlights variations in the structure and composition of LPS among bacterial species, which consequently affect the barrier properties of the outer membrane permeability. The LPS consists of three components: lipid A, core oligosaccharide, and the O-polysaccharide or O antigen (Figure 3). It is highly polyanionic, primarily due to (i) the presence of two monophosphate groups substituting the lipid A at positions 1 and 4′, and (ii) the carboxylic groups of the 3-deoxy-D-manno-oct-2-ulosonic acid (Kdo), along with monophosphate (or pyrophosphate or diphosphate) groups substituting the heptose residue(s) at the inner core level. Additionally, the oligosaccharide core harbors numerous anionic groups, the distribution of which varies among Gram-negative bacterial species [405]. The O-antigen composition varies significantly and is unique to each bacterium [185,408]. The enzymes responsible for catalyzing the biosynthesis of phospholipids and LPS are well characterized in E. coli and are located either in the cytoplasm or on the inner cytoplasmic leaflet of the inner membrane [409,410]. Considering the crucial role of LPS in protecting bacteria from the entry of hydrophobic and hydrophilic molecules [411], enzymes involved in LPS biosynthesis, such as Lpx [206], and LPS modifications induced by PhoPQ and PmrAB, play a significant role in the resistance of many Gram-negative bacteria against AMPs [216,412] (Table 2). Additionally, specific mutations within these TCSs, as well as in ColRS, ParRS, and CprS TCSs of P. aeruginosa, lead to their constitutive activation, resulting in the subsequent overexpression of LPS-modifying genes in these bacteria [37].

Resistance to AMPs Due to O-Antigen and Core Oligosaccharide

The specific O chains define the surface properties of bacteria in species that do not produce capsules. It acts as a protective shield that prevents the entry of AMPs into the LPS bilayer (Table 2). These oligosaccharide chains are crucial for the bacteria that produce them to avoid recognition by the host’s immune defenses. For example, C. jejuni mimics the gangliosides (glycolipids) on the surface of its host’s eukaryotic cells [413]. O-antigens have been shown to provide protection to S. flexneri and K. pneumoniae by shielding them from histones, significant components of neutrophil extracellular traps and acting as AMPs [205]. In addition, a larger O-antigen provides increased protection against AMPs. Indeed, resistance to AMPs is significantly reduced in several bacteria that possess truncated or completely absent LPS antigens, commonly referred to as rough LPS mutants (LPS-R), compared to WT bacteria with smooth LPS (LPS-S) [239,414,415,416,417,418] (Table 2).
Such differences in susceptibility are most likely due to variations in the LPS content of divalent cations [419,420]. In fact, LPS chelates a multitude of divalent cations, including Mg2+ and Ca2+, owing to its negative charges [421,422,423]. The partial neutralization of these negative charges by the bridges formed with these cations diminishes repulsion between adjacent LPS molecules and stabilizes lateral interactions [424]. AMPs interacting with the outer layer of LPS competitively displace divalent cations, causing the destabilization of the LPS bilayer [425]. Interactions between neighboring O chains may thus contribute to the stability of this bilayer [426]. These intermolecular interactions through O chains (and likely the outer core) are markedly more significant in LPS-S than in LPS-R [427]. Yeh and Jacobs (1992) demonstrated that the presence of long O chains in LPS reduces the fluidity of the phospholipid bilayer more effectively than those with short or no O chains. This indicates that O chains influence the behavior of the lipid A domain [428].
The structure of LPS can influence the interaction of AMPs with the outer membrane. This has been elucidated in S. Typhimurium with magainin 2, showing that the sensitivity of Gram-negative bacteria to this peptide is associated with factors that facilitate the transport of AMPs across the outer membrane. These include the importance and location of LPS charge, LPS concentration in the outer membrane, its architecture, as well as the presence or absence of the lateral chain of the O-antigen [429]. Several studies have subsequently shown that resistance to AMPs depends on variations in the chemical composition of LPS, such as the importance of the charge of O antigen saccharide chains and the specific glycosylation of the latter (through gtr genes). Glycosylation leads to a modification of the conformation of the O antigen subunits by halving the thickness of the LPS layer, which can increase bacterial resistance to AMPs, particularly in S. flexneri and Bordetella sp. [151,272].
The enzymes involved in the biosynthesis of the outer and/or inner core oligosaccharide, the nature of the latter, and the density of the O-polysaccharide chain all contribute, either collectively or individually, to enhanced resistance to AMPs (Table 2). For example, the removal of this core can downregulate the expression of the pmrD gene responsible for E. coli resistance against polymyxins [179]. In C. jejuni, the galU gene is indeed responsible for synthesizing uridine diphosphate (UDP)-glucose, which is a precursor molecule required for the synthesis of LPS. When the galU gene is mutated or disrupted, the production of UDP-glucose is impaired, resulting in the truncation of the LPS core structure. This truncation makes the bacterium more susceptible to AMPs and other antimicrobial agents [164].

Roles of the Acylation and Modifications of Lipid A Acyl Chains in Intrinsic and Induced Resistance to AMPs

The higher content of saturated fatty acids in LPS, coupled with divalent cations, induces rigidity in LPS [419,430,431]. Several experiments have conclusively shown that the portion of LPS anchored in the outer membrane, specifically the fatty acids, exhibits a remarkably low fluidity [432,433]. Fluidity, in this context, refers to the ability of molecules to move within the same layer. Unlike PG, the glycerophospholipids typically found in other biological membranes such as the cytoplasmic membrane, which contains only two chains of saturated fatty acids, LPS comprises between four and seven chains. However, unlike unsaturated fatty acids, saturated fatty acids reduce the fluidity of the lipid layer they compose due to their larger steric hindrance [434]. Consequently, the absence of unsaturated fatty acids renders the inner leaflet of the LPS (comprising fatty acid chains) much less fluid, thereby diminishing the permeability of the outer membrane. Furthermore, the number, distribution, length, and type of the acyl chain in the fatty acid of lipid A vary among various bacteria, such as Acinetobacter baumannii, B. cepacia, B. pseudomallei, C. jejuni, H. pylori, E. coli, N. gonorrhoeae, K. pneumoniae, S. enterica, P. aeruginosa, Vibrio cholerae, and Y. pestis, which can influence their membrane properties [435].
  • Acylation of Lipid A in LPS
Lipid A typically contains four primary chains of (R)-3-hydroxyacyl directly and covalently linked to the glucosamine disaccharide at positions 2, 3, 2′, and 3′. Some of these (R)-3-hydroxyacyl groups are modified by secondary acyl chains, forming acyloxyacyl fractions [435]. For example, in E. coli and S. Typhimurium, secondary acyl groups of laurate (12 carbons) and myristate (14 carbons) are attached to the 3-Kdo2-lipid A at positions 2′ and 3′ of the primary acyl chain, respectively (Figure 3). The incorporation of lauroyl chain is mediated by the LpxL acyltransferases (HtrB in Salmonella, Shigella, and H. influenzae), while the myristoyl chain is mediated by LpxM (formerly MsbB or WaaN) [435,436]. The activation of HtrB requires a high temperature in E. coli [437], whose genome also contains the lpxP gene (homologous to lpxL), induced by cold shock (12 °C). LpxP catalyzes the incorporation of palmitoleate (unsaturated fatty acid) instead of laurate, suggesting the bacterium needs to adjust the fluidity of its outer membrane [436,438].
Only a limited number of laboratory knockout mutations (through insertional inactivation) can be generated at the lipid A level without impacting bacterial viability. Among these mutations are those affecting the lpxA, lpxM, and htrB genes. For instance, in N. meningitidis, a viable lpxA (encoding UDP-GlcNAc acyltransferase required for the first step of lipid A biosynthesis) mutant that is entirely deficient in LPS experiences severe growth impairment. Additionally, there is an increase in phosphatidylethanolamine (PE) and PG species, predominantly composed of shorter chains, thus hindering in vivo analyses [439]. On the other hand, the inactivation of lpxM has been extensively studied in several Enterobacterales [440,441]. Myristoylation of lipooligosaccharide (LOS), catalyzed by the acyltransferase HtrB, is a crucial factor in conferring resistance to human β-defensin by the human pathogen H. influenzae. Mutant strains lacking HtrB are 45 times more sensitive than the WT [198]. However, mutations in the lpxM gene in several bacterial pathogens do not provide conclusive findings regarding the resulting phenotypes. For instance, lpxM mutants of S. Typhimurium and S. flexneri exhibit attenuated ability to induce inflammation in vivo [440,442]. Alterations in growth and membrane permeability observed in the lpxM mutants of S. Typhimurium lead to the selection of extragenic suppressor mutations in this bacterium [443]. The only instance of observed virulence attenuation in a murine infection model directly linked to the lpxM gene mutation was observed in a clinical strain of E. coli. The lpxM gene is not indispensable for E. coli growth [443,444,445], but its mutation leads to side effects such as capsule reduction [446]. Furthermore, the inactivation of the lpxP gene in E. coli does not affect bacterial growth but heightens bacterial sensitivity to certain antibiotics at low growth temperatures [447]. The role of covalent modifications catalyzed by LpxM in maintaining membrane integrity, as well as enhancing its permeability and resistance to cationic AMPs, has also been elucidated in K. pneumoniae, A. baumannii, and V. cholerae [206] (Table 2).
Resistance to polymyxins in E. coli and S. Typhimurium requires the LpxM-mediated myristoylation of their lipid A since the addition of L-Ara4N to the phosphate groups of lipid A (catalyzed by l- Ara4N transferase ArnT) depends on the presence of the secondary myristate (catalyzed by LpxM) chain at position 3′ of glucosamine [182]. Moreover, mutations of lpxP and htrB in Y. pestis result in increased susceptibility to polymyxin B at 21 °C and 37 °C, respectively. Additionally, the acylation of lipid A plays a role in the expression of virulence factors [280].
A. baumannii can develop resistance to colistin through the loss of its LPS (Table 2). The outer membranes of polymyxin-resistant strains, with lipid A modified with PEtN or deficient LPS, exhibit distinct atomic-level interactions with polymyxins [448]. In mouse models, LPS-deficient A. baumannii strains exhibit reduced virulence. The development of colistin-resistant LPS-deficient strains involves spontaneous mutations in genes such as lpxA, lpxC (encoding UDP-3-O-(R-3-hydroxymyristoyl)-N-acetylglucosamine deacetylase), and lpxD (encoding UDP-3-O-(3-hydroxymyristoyl) glucosamine N-acetyltransferase). Studies have shown that LPS-deficient A. baumannii weakly stimulate neutrophils, leading to lower levels of reactive oxygen species (ROS) and inflammatory cytokine production. Furthermore, LPS-deficient A. baumannii strains are more susceptible to antibacterial lysozyme and lactoferrin compared to their WT counterparts [133,135]. Jennifer H. Moffatt et al. (2013) also demonstrated that LPS-deficient A. baumannii are more susceptible to LL-37 and exhibit altered signaling through host Toll-like receptors [134].
Some Francisella species, like F. tularensis, responsible for tularemia, synthesize relatively short LPS molecules with a lipid A devoid of Kdo and other saccharide chains [449]. The lipid A of these bacteria is devoid of both a phosphate group and acyl chains at the 4′ and 3′ ends, respectively. The dephosphorylation of lipid A in Francisella is facilitated by the 4′- phosphatase LpxF, a constituent of the LPS biosynthesis pathway in this bacterium. The lpxF mutant of F. novicida displays a heightened sensitivity to cationic AMPs [193] (Table 2). This increased susceptibility damages the bacterial envelope, exposing membrane lipoproteins and bacterial DNA to Toll-like receptors TLR2 and TLR9, respectively [450]. Notably, the dephosphorylation of lipid A by LpxF is contingent upon the absence of a secondary myristate chain at position 3′ due to its steric hindrance chain near the 4′-phosphate group [193].
Similarly, H. pylori, Rhizobium leguminosarum, P. gingivalis, and Rhizobium etli tend to remove the anionic phosphate group from the lipid A to resist against AMPs by using the 4′-phosphatases LpxF [167]. On the other hand, some bacteria, such as Helicobacter pylori, has a tendency to remove the 1-phosphate group from lipid A using the 1-phosphatase LpxE encoded by lpxE gene, and subsequently add the phosphoethanolamine (PEtN) group, facilitated by the EptA enzyme [167,201,202]. Furthermore, the deletion mutants of lpxE gene in Capnocytophaga canimorsus and Brucella abortus exhibited increased polymyxin B sensitivity [155,167,245] (Table 2).
pagP, one of the genes induced by PhoPQ in Salmonella, encodes an unusual acetyltransferase named PagP, located in the outer membrane and involved in LPS biosynthesis. PagP catalyzes the transfer of palmitate from PE (donor) to the OH group of the 3-hydroxymyristate chain linked at position 2 to the lipid A of LPS (Figure 3). This increases the number of acyl groups per lipid A molecule to a total of seven chains (in addition to the six chains constitutively incorporated) [451]. Furthermore, the palmitate chain is longer than those previously incorporated during lipid A biosynthesis as it consists of 16 carbon atoms (C16:O). All these functions are likely to alter the fluidity of the outer membrane by increasing hydrophobic interactions and van der Waals bonds between the acyl chains of neighboring LPS molecules. This stabilization is suggested by the fact that palmitate incorporation catalyzed by PagP confers resistance to cationic AMPs in S. enterica [452] and attenuates the ability of LPS to trigger immune responses via the TLR4 Toll pathway [453]. PagP mutants of S. Typhimurium are sensitive to certain AMPs, such as C18G and protegrins, but are resistant to polymyxins. The increased acylation of lipid A alters the hydrophobic interaction between fatty acids and the AMP, thereby preventing or at least delaying its insertion into the bilayer [454]. The involvement of the pagP gene (or its homologues) in inducible resistance to AMPs has also been demonstrated in the pathogens Bordetella parapertussis, B. bronchiseptica (rcp gene for resistance to cationic antimicrobial peptides) [153], Legionella pneumophila, and Yersinia pseudotuberculosis [216,284,455]. The regulation of these genes is correlated with the bacterial lifestyle during infection. Some studies have highlighted that PagP might function as an apical sensory transducer, which can be activated by a breach in the outer membrane of the enterohemorrhagic strain E. coli O157:H7. Indeed, PagP is capable of sensing assaults that alter the permeability of the outer membrane, such as the action of EDTA, which chelates Mg2+ cations, or mutations affecting the presentation of LPS on the bacterial surface. Thus, the mutation of the lpxM (msbB) gene of E. coli O157:H7 triggers the palmitoylation of lipid A catalyzed by PagP, thereby restoring the permeability barrier role ensured by the outer membrane [180].
Recently, Sun et al. have revealed the role of Crrab TCS in upregulating the PagP-mediated palmitoylation of glycerophosphoglycerols and lipid A in K. pneumoniae, aiming to confer high-level polymyxin resistance and virulence. They demonstrated that PagP transfers palmitate from the glycerophospholipid of the OM inner layer to glycerophosphoglycerols, forming acyl-glycerophosphoglycerols, thereby boosting the ratio of acyl-glycerophosphoglycerols to glycerophosphoglycerols and enhancing outer membrane hydrophobicity [207].
  • Deacylation of Lipid A in LPS
pagL is a PhoPQ-governed gene encoding for PagL, an enzyme with lipase or deacylase activity, localized in the outer membrane of S. Typhimurium. It facilitates the deacylation or removal of the hydroxymyristate chain linked at position 3 to the glucosamine of lipid A in Salmonella [456] (Figure 3). Conversely, LpxR is another outer membrane lipase that cleaves the covalent bond between the acyloxyacyl chain linked at 3′ and the Kdo2-lipid A portion in a calcium-dependent manner [457]. PagL mutants of S. enterica are more sensitive to polymyxin B than the corresponding WT strain. Both LpxR and PagL lipases remain inactive in the outer membrane and are only activated upon overproduction, suggesting the presence of endogenous inhibitors, such as L-Ara4N [261]. The latency of PagL partly depends on the incorporation of PEtN at the Kdo2-lipid A level. The activation of PagL facilitates the deacylation of glucosamine at position 3, compensating for the decreased resistance to cationic AMPs caused by the absence of L-Ara4N at the lipid A of LPS [261]. Since this modification is only observed following PagL overproduction, it is possible to speculate that PagL is necessary for minor adjustments to the fluidity of the LPS layer after the addition of seven fatty acid chains to its lipid A.
Additionally, NaxD (N-acetylhexosamine deacetylase), a deacetylase from the YdjC superfamily, deacetylates N-acetylgalactosamine (linked to the lipid carrier undecaprenyl phosphate) to positively charged galactosamine, a necessary step for the subsequent addition of galactosamine to lipid A, contributing to polymyxin resistance in Francisella and A. baumannii [136,195] (Table 2). NaxD-mediated lipid A modification in A. baumannii is regulated by the sensor kinase PmrB [136]. Finally, spontaneous mutations and insertion sequences (IS) in the N-acetylglucosamine deacetylase encoding lpxC gene were shown to be associated with colistin resistance in colistin-resistant LPS-deficient strains of A. baumannii [133].
  • Hydroxylation of Lipid A in LPS
LpxO (or Pag Q) is an inner membrane Fe2+/alpha-ketoglutarate-dependent dioxygenase in Salmonella that catalyzes, in a PhoPQ-dependent manner, the hydroxylation (incorporation of the OH group) of the 2-hydroxymyristate chain of Kdo2-lipid A [458,459] (Figure 3). It compensates for the decrease in OH groups caused by PagP’s catalytic activity. The OH groups play a vital role as hydrogen bond donors, stabilizing interactions between neighboring LPS molecules. Unlike membrane phospholipids, the lipid A of LPS contains a number of hydroxylated fatty acids. Typically, the residues at positions 2 and 3 of glucosamine lack secondary myristate chains, resulting in two hydroxyl groups within the membrane that can act as hydrogen bond donors. The NH groups (2 groups at positions 2 and 2′) and the 4-OH group of reduced glucosamine are available to act as hydrogen bond donors. Additionally, the inner core oligosaccharide linked at position 6′ to the unreduced glucosamine provides numerous groups that could serve as both hydrogen bond donors and acceptors [405] (Figure 3). LpxO-dependent lipid A 2-hydroxylation has been shown to be essential for protecting A. baumannii against cationic AMPs, aiding its survival in human whole blood and the wax moth Galleria mellonella (Table 2). Additionally, LpxO shields Acinetobacter from G. mellonella AMPs, controlling their expression [137]. Similarly, in K. pneumoniae, LpxO-dependent modification dampens inflammatory responses and confers resistance to AMPs. Notably, a lpxO mutant exhibits attenuation in vivo, underscoring the significance of this lipid A alteration in Klebsiella infection. Colistin prompts the in vivo lipid A pattern, already expressed in colistin-resistant isolates, where LpxO-dependent lipid A modification is pivotal for resisting colistin [208]. Similarly, LpxO was suggested to confer resistance against AMPs and contribute to the pathogenicity of S. enterica, P. aeruginosa, and B. bronchiseptica [244,262,460]. Another enzyme called LpxN functions as a lipid A secondary hydroxy-acyltransferase in V. cholerae, responsible for the transfer of 3-hydroxylaurate to the lipid A domain significantly enhances polymyxin resistance [275] (Table 2).

Role of Cationic Polar Groups Added to Lipid A in Induced Resistance to AMPs

Cationic AMPs selectively bind to the negatively charged phosphate moieties and acidic groups within the lipid A and core polysaccharide by electrostatic interactions [419]. Many Gram-negative bacteria employ various strategies to reduce the interaction between cationic AMPs and their outer membranes by adding positively charged residues. The primary LPS modification involves the cationic substitution of phosphate groups with L-Ara4N, which neutralizes the net negative charge of lipid A, while the second most common modification is the addition of PEtN, reducing the net charge from −1.5 to −1 [136,405,461] (Figure 3). Among these modifications, the L-Ara4N substitution is deemed more effective due to its impact on charge alteration. For example, polymyxin-resistant Gram-negative species inherently possess LPS with one or more of these modifications. Indeed, S. Typhimurium emerges as a prominent example, highlighting how resistance to AMPs through alterations in LPS serves as a critical mechanism for survival. Although most of the genes required for these modifications are encoded in the chromosome, the recently identified plasmid-borne PEtN transferases (mcr-1 to mcr-10) pose a threat by potentially accelerating the spread of clinically significant colistin resistance [462].
  • Addition of Aminoarabinose to Lipid A in LPS
Gram-negative bacteria, including E. coli and S. enterica, reduce the negative charge of their surface by incorporating positively charged L-Ara4N at the 4′ phosphate end of their lipid A (Figure 3) (Table 2). This modification decreases the affinity of cationic AMPs, like polymyxin, to the outer membrane and reduces electrostatic repulsions between neighboring LPS molecules [463]. The addition of L-Ara4N occurs in a PhoPQ-dependent manner, which indirectly activates the PmrAB TCS via PmrD [183,464,465]. The biosynthesis and addition of L-Ara4N are catalyzed, respectively, by two genetic loci: the pmrE gene (also known as pagA or ugd) and the pmrHFIJKLM operon (also termed arnBCADTEF or pbgP) [466,467,468]. Non-polar mutagenesis studies have shown that all genes except pmrM are essential for the addition of L-Ara4N and resistance to cationic AMPs. A significant cause of colistin resistance in K. pneumoniae and E. coli involves mutations or the inactivation of the regulatory mgrB gene, which normally exerts a negative feedback on PhoPQ by inhibiting PhoQ kinase activity and/or enhancing its phosphatase activity. This results in the activation of PhoPQ, leading to the upregulation of all PhoP-mediated lipid A modifications, including the addition of L-Ara4N and PEtN to lipid A, as well as its deacylation [184,209]. Similarly, the PbgPE-mediated incorporation of 4-aminoarabinose into lipid A, in a PhoPQ-dependent manner, is essential for the resistance of the entomopathogen P. laumondii to polymyxins and insect-derived cecropins A and B, as well as for virulence in insects [233,234]. The substitution of lipid A moieties with L-Ara4N have also been demonstrated in vivo during cystic fibrosis caused by P. aeruginosa [247].
  • Addition of Phosphoethanolamine to Lipid A in LPS
The addition of PEtN to the LPS can counteract the effects of negatively charged residues. The PmrAB system controls the PEtN transferase EptA in Salmonella spp. and E. coli [185]. In E. coli, the PEtN addition to lipid A, Kdo, and heptose I of LPS is catalyzed by the EptA, EptB, and EptC enzymes, respectively. Another protein, termed CptA, responsible for the addition of PEtN to the core of Salmonella LPS, has been identified [264]. The incorporation of PEtN groups to lipid A and Kdo alters the net charge of LPS, providing defense against polymyxin B binding and penetration, while PEtN added to heptose I, alongside the reduction in negative charges, also prevents the insertion of the polymyxin B lipophilic tail into the outer membrane [186]. In A. baumannii, resistance to colistin is linked to the addition of PEtN to lipid A, facilitated by specific amino acid mutations in the PmrB protein. This results in the overexpression of pmrC, responsible for producing the PEtN transferase enzyme [138]. Recently, a novel TCS named StkSR was characterized in A. baumannii. Deleting stkR resulted in a notable upregulation of pmrA, pmrC, and pmrB expression, subsequently enhancing pmrC transcription and facilitating the replacement of lipid A with PEtN [139]. Another potential pathway that could result in the overproduction of PEtN in A. baumannii entails the integration of the ISAbaI insertion element upstream of an eptA isoform [140].
Further studies have demonstrated that constitutive mutations in the pmrA gene suppressed msbB mutant growth defects in Salmonella. In addition, the PEtN addition to lipid A confers polymyxin resistance in msbB-mutant strains, highlighting the necessity of aminoarabinose biosynthetic enzyme for the incorporation of PEtN and palmitate to the lipid A of these mutants [469].

3.4.2. Outer Membrane Proteins and Resistance to AMPs

The outer membrane proteins (OMPs) of Gram-negative bacteria comprise a diverse array of proteins, including, among others, porins and adhesins, which fulfill diverse roles, such as signaling, adhesion, catalyzing reactions, and the active and passive transport of solutes and nutrients into and out of the cell [470]. AMPs have been proposed to utilize OMPs as a gateway for entry into bacterial cells [471]. This notion was supported by the observation that the absence of the MtrE channel unexpectedly increased gonococcal survival after exposure to azurocidin, prompting speculation about whether MtrE might facilitate the passage of certain antimicrobials across the outer membrane [224]. Additionally, the deletion of tolC has been found to enhance the activity of the transcriptional regulators MarA, SoxS, and Rob, which regulate porin expression and could potentially modify membrane permeability [472]. The idea that MtrE could act as a portal for AMP entry gained support from a recent discovery demonstrating that TolC is capable of importing bacteriocins (MW 60 kDa) in Gram-negative bacteria [473].
On the other hand, Visser et al. (1996) were among the first to describe the role of Y. enterocolitica adhesin A (YadA), in bacterial resistance to cationic AMPs [281]. In V. cholerae, the outer membrane proteins OmpU and OmpT, whose expression depends on the bacterial growth phase, partially contribute to basal resistance to cationic AMPs [276]. However, this resistance is further enhanced in the presence of AMPs in the bacterial environment. OmpU appears to act as a membrane sensor that detects antimicrobial peptides and induces resistance pathways through the transcription factor σE and the activator DegS [277]. Furthermore, studies have demonstrated that OmpA confers resistance to polymyxin B and protamine in K. pneumoniae [212]. The only case where porins have been directly involved in resistance to AMPs is observed with E. coli, where blocking OmpF by the simultaneous addition of spermine or cefepime inhibits colicin action [187]. Interestingly, the presence of a gap formed by OmpF facilitates the insertion of AMPs into the LPS, enabling them to establish hydrogen bonds with the phosphate groups of the inner core oligosaccharides. OmpF was found to play a crucial role in the entry of AMPs such as P6, either by elaborating the binding site for LPS or by directly transporting AMPs across the outer membrane [188] (Table 2). Moreover, Zhang et al. discovered that the downregulation of OmpW mediated by SoxS, a transcriptional factor involved in oxidative stress, contributes to resistance against oxidative stress and may also play a role in resistance against AMPs in E. coli [474]. A reduction in the expression of OmpW was also observed in a colistin-resistant mutant of A. baumanni [475]. Finally, OmpA-like proteins were show to confer AMP resistance in many other Gram-negative pathogens (Table 2).

3.4.3. Peptidoglycan Modifications and AMP Resistance

To reach their target structures inside bacterial cells, particularly in Gram-positive bacteria, AMPs must penetrate through the dense layer of peptidoglycan. The physicochemical properties and density of this layer are pivotal in determining bacterial sensitivity to antimicrobial agents [476].
In Salmonella and E. coli, the N-acetylmuramoyl-l-alanine amidases, encoded by amiA and amiC, contribute to increased resistance to protamine, magainin 2, and melittin. The expression of these genes is governed by the twin arginine translocation (Tat) system. Conversely, the CpxR/CpxA two-component system is regulated by the overexpression of the nlpE gene, which encodes an outer membrane lipoprotein involved in copper homeostasis and adhesion. This system upregulates amiA and amiC expression, thereby enhancing resistance [189].
The N-deacetylation of peptidoglycan by the peptidoglycan N-deacetylase (encoded by the pgdA gene) in L. monocytogenes is a crucial mechanism for evading the host innate immune response. Boneca et al. (2007) demonstrated that mutants of L. monocytogenes lacking PgdA are highly susceptible to lysozyme activity, rapidly eliminated by macrophages, and elicit a significant IFN-β response through the TLR2 and Nod1 pathways [329]. In S. aureus, the carbon at position 6 of the N-acetylmuramic acid is acetylated by the O-acetyltransferase Oat, which grants these bacteria resistance to the muramidase activity of lysozyme [352]. In group B streptococcus (GBS), a mutant of the ponA gene, encoding for the penicillin-binding protein 1a (PBP1a) (both transglycosylase and transpeptidase), exhibits high sensitivity to defensins and LL-37 [323]. Furthermore, a mutation in the pgm gene, which encodes a phosphoglucomutase essential for preserving peptidoglycan integrity in Streptococcus iniae, increases the bacterial susceptibility to AMPs [380]. N- and O-deacetylations of peptidoglycan have been linked to AMP resistance in many other bacterial pathogens (Table 1 and Table 2).

3.4.4. Modifications of Teichoic Acids and AMP Resistance

TAs constitute the second major component of the cell wall in Gram-positive bacteria, comprising more than 50% of the cell wall mass. TAs are polyanionic polymers that exist in many forms, such as wall teichoic acids (WTAs), which are covalently bound to the peptidoglycan and are presented on the outer surface of bacteria, and lipoteichoic acids (LTAs), which are anchored to glycolipids of the cytoplasmic membrane [477,478]. Mutants of S. aureus deficient in WTA exhibits up to a 100-fold increase in resistance to degradation and killing by gIIA phospholipase A2 (PLA2) and human β-defensin 3 (HBD-3) compared to the WT (Table 3). However, these mutants maintain sensitivity similar to the WT to other cationic AMPs, including Magainin 2 amide, HNP1-3, LL-37, and lactoferrin [358].
Specific enzymes called WTA glycosyltransferases mediate the glycosylation of WTA in L. monocytogenes. These enzymes catalyze the transfer of sugar molecules, such as rhamnose and N-acetylglucosamine in serovar 1/2a, and galactose and glucose in serovar 4b, into the WTAs, thereby modifying their structure and leading to resistance to AMPs [331,332]. On the other hand, many studies have revealed that the D-alanylation of TAs reduces the overall negative charge of the bacterial surface and confers resistance to AMPs (Figure 4) [479] (Table 3). This substitution is also crucial for stimulating TLR2 [480,481], and the absence of such a substituent leads to a significant decrease in bacterial inflammatory properties [482]. The incorporation of positively charged D-alanine residues into LTAs is catalyzed by products of the dlt operon [479]. This operon has been characterized in at least thirteen species of Gram-positive bacteria where it confers resistance to various AMPs (Table 3). The dlt operon generally includes the genes dltA, dltB, dltC, and dltD, all of which are essential for the D-alanyl esterification of TAs (Figure 4). The D-alannylation of TAs is completed in two steps: first by the D-Alanine-D-alanyl carrier protein ligase (Dcl), encoded by the dltA gene, and then by the D-alanyl carrier protein (Dcp), encoded by the dltC gene. The dltB and dltC genes are also indispensable in this D-alanylation process. Depending on the bacterial species, the content of D-alanyl ester in WTAs and LTAs depends, among other factors, on the pH of the environment, temperature, and concentration of NaCl and KCl salts [479].
The inactivation of the dltA, dltB, dltC, or dltD genes in the non-pathogenic bacterium B. subtilis results in mutants with WTAs and LTAs devoid of D-alanyl ester residues [295]. The knock-out mutations of the dlt operon in many Gram-positive species have been shown to significantly reduce resistance to both natural and synthetic cationic AMPs (Table 3). Furthermore, many dlt mutants are severely impaired in their virulence in animal models and their ability to adhere to and invade several eukaryotic cell lines in vitro, probably due to the highly electronegative surface of the bacteria and/or altered adhesin-binding activities [301,333] (Table 3). These features also suggest that D-alanine esterification of TAs contributes in various ways to the ability of some bacteria to circumvent mucosal and systemic antimicrobial defenses to produce systemic infections. Although the pathogen S. pneumoniae R6 contains phosphorylcholine esters instead of D-alanyl esters in its TAs [483], the dlt operon has been identified in its genome and confers resistance to cationic AMPs [484]. The inactivation of dltA in these pneumococci increases bacterial sensitivity to antibacterial molecules (chromatin and protein granules) from neutrophil extracellular traps (NETs) [485]. The reduction in D-alanyl esters in teichoic acids within Lactobacillus reuteri was also shown to decrease resistance to nisin and to impair the colonization of the mouse gastrointestinal tract [486].
The control of dlt gene expression varies among different Gram-positive bacterial species because they colonize different niches and adapt to different growth and stress conditions. For example, the dlt operon of B. subtilis is part of the σx regulon, and its expression also depends on the global regulators SpoOA and AbrB [294,295]. Additionally, Poyart et al. (2001) showed that the activation of the DltSR TCS depends on the amount of D-alanine available in the cytoplasm, providing resistance to cationic peptides by increasing cell wall density [327,328]. In S. aureus, dlt operon transcription is repressed in response to high concentrations of Na+ as well as moderate concentrations of Mg2+ and Ca2+ in the extracellular environment, while the dltD gene is derepressed (58-fold) by the agr locus (accessory gene regulator) encoding the Agr quorum sensing regulator [366]. The GraRS TCS has also been shown to control D-alanylation in S. aureus, conferring resistance to host AMPs [367]. The regulation of the dlt operon in S. epidermidis and S. aureus is also under the control of a three-component regulation system, ApsSX/ApsR (Aps, antimicrobial peptide sensor), homologous to GraRS. The sensor AspX is highly specific in recognizing and binding cationic AMPs, which subsequently induces the expression of other genes involved in AMP resistance, including mprF and the ABC-type VraFG transporters [368,369] (Figure 4). The regulation of the dlt operon in L. monocytogenes is controlled by the regulator VirR belonging to the pleiotropic VirRS regulation system [334]. Finally, S. pneumoniae upregulates the dlt locus through the CiaRH sensoring system, which senses stresses caused by AMPs, resulting in the D-alanylation of TAs and increased inflammatory responses [387].

3.5. Role of the Cytoplasmic Membrane Phospholipids and Their Modifications in Resistance to AMPs

The primary lipids found in the cytoplasmic membrane of bacteria consist of anionic PG, CL (also known as diphosphatidylglycerol, synthesized by cardiolipin synthase), and the neutral or zwitterionic PE. For example, the membrane lipid composition of E. coli comprises predominantly 80% PE and only 15% PG, in contrast to S. aureus, which lacks PE and contains 58% PG and 42% CL [487]. These variations in phospholipid content may account for the ineffectiveness of some cationic antimicrobials against bacteria. In particular, the alteration of CL levels within membranes has been proposed as a mechanism through which pathogens develop resistance to cationic AMPs [488]. Molecular dynamics simulations of the interaction between the short AMP aurein 1.2 and an anionic CL-containing lipid bilayer showed that the structural properties of CL, including rigidity, exposed charged phosphate groups, and the promotion of negative membrane curvature, oppose aurein 1.2’s membrane destabilization mechanism, potentially contributing to bacterial resistance against positive membrane curvature-dependent AMPs [12]. In fact, negative curvature-inducing lipids, like CL and phosphatidylserine, block the lytic activity of short AMPs, like aurein 1.2, magainin 2, polybia-MP1, LL-37, and ΔM2. This hindrance requires higher peptide-to-lipid ratios for peptide-induced transmembrane pore formation compared to PG [12]. Furthermore, a minor amount of phosphatidic acid (anionic), lysyl-phosphatidylglycerol (LPG), which bears a positive charge, and glycolipids are also present in bacterial membranes [489]. Other mechanisms within the cytoplasmic membrane, such as alterations in lipid charge and membrane energetics, also contribute to bacterial resistance to AMPs.
The first observations regarding modifications of bacterial cytoplasmic membrane phospholipids in response to environmental cues were made in P. fluorescens. It has been shown that P. fluorescens can adapt to cationic AMPs by altering its cytoplasmic membrane. Transitioning P. fluorescens from a phosphate-rich to a phosphate-deficient environment led to a significant reduction in the levels of PE, PG, and CL in the cytoplasmic membrane. This shift in membrane composition, particularly in anionic phospholipids, was followed by the emergence of a cationic lipid component containing ornithine. The resistance of P. fluorescens to polymyxin B was found to be proportional to the amount of this lipid found in the cytoplasmic membrane of these bacteria [259]. Cao and Helmann (2004) identified the pss-ybfM-psd operon as part of the σX regulon in B. subtilis. This operon is involved in PE biosynthesis, which, when incorporated into the bacterial cytoplasmic membrane, decreases its overall net negative charge, leading to a reduced affinity for cationic AMPs [294]. Other Gram-positive bacteria, such as certain Staphylococcus species, consistently express membrane phospholipids with reduced negative charge [490]. The analysis of the constitutive phospholipid composition in several of these species reveals predominantly polar lipid profiles comprising PG and CL. Among the staphylococci tested, S. aureus stands out for its lipid composition enriched in unsaturated menaquinones with eight isoprene units, and LPG, a derivative of PG that exhibits considerably lower electronegativity [491]. In staphylococci, resistance to daptomycin by the alteration of membrane phospholipid composition is postulated to change the fluidity of the membrane, thus interfering with daptomycin binding and subsequent oligomerization. Indeed, by analyzing the action of a membrane active AMP on giant unilamellar vesicles (GUVs), increasing LPG levels did not correlate with reduced peptide binding by electrostatic repulsion, but with the inhibition of intravesicular dye leakage post-binding, indicating a protective effect on membrane integrity [492]. Considering the protective role of phospholipids and fatty acid types in regulating membrane fluidity, it is plausible that CL contributes to preventing daptomycin translocation upon insertion into the membrane [493]. Mutations that modify enzyme function could potentially influence daptomycin resistance by shifting the PG to CL ratio in the cell membrane. Supporting this notion, the genomic analysis of 33 daptomycin-resistant Staphylococcus strains revealed associations between mutations in genes such as pgsA (involved in PG synthesis) and cls2 (cardiolipin synthase) and daptomycin resistance [370] (Figure 5G) (Table 3). Moreover, Bayer et al. (2000) demonstrated that S. aureus strains resistant to tPMP-1 (thrombin-induced platelet microbicidal protein 1) in vitro exhibit a notable increase in unsaturated membrane lipids compared to genetically similar strains that are more susceptible to this cationic AMP [494].
Several Gram-positive bacteria mitigate the negative charge of PG by replacing it with a cationic residue, lysine, forming LPG [371] (Figure 5G) (Table 3). The enzyme responsible for this incorporation is the product of the mprF (multiple peptide resistance factor) gene, known as the LPG synthase MprF. MprF is a bifunctional enzyme, consisting of a carboxy-terminal cytoplasmic tail responsible for the lysinylation of PG using lysyl-tRNA as a donor and an amino-terminal domain that facilitates “flippase” activity, transporting LPG from the inner to the outer membrane. Additionally, a central domain appears to assist in both lysinylation and flippase activities [495].
S. aureus responds to daptomycin by increasing its overall cell-surface charge, likely to repel daptomycin insertion and maintain membrane integrity, a process facilitated by mutations in mprF. These mutations, resulting in amino acid changes clustering in the central bifunctional region that overall confer a “gain-of-function” to the enzyme [496], lead to a gain-of-function phenotype, enhancing the synthesis of positively charged LPG. Studies have also demonstrate that expressing mprF with daptomycin-resistant mutations (but not WT mprF) restores daptomycin resistance in mprF mutant strains of S. aureus [497]. Additionally, inhibiting MprF protein synthesis by antisense RNA (directed against mprF transcripts) reverses daptomycin resistance in vitro in strains with gain-of-function mutations [498].
On the other hand, the exposure of S. aureus to subinhibitory levels of magainin 2 and gramicidin D prompts the bacterium to develop resistance against these peptides in a MprF-dependent manner [372]. The inactivation of the mprF gene in S. aureus results in the depletion of membrane LPG, heightening the bacterial susceptibility to defensins and PLA2 by accumulating significantly more surface-bound cationic AMPs than WT bacteria [371,499,500]. Similarly, human neutrophils containing a high quantity of α-defensins deactivate ΔmprF mutants of S. aureus much more rapidly and effectively than WT bacteria [371] (Table 3).
Homologues of MprF have also been identified and functionally characterized in B. anthracis, L. monocytogenes, and Enterococcus sp. [290,335] (Table 3). The lysylation of PGs in L. monocytogenes depends on the VirR response regulator, which is critical for the virulence of this bacterium [334]. Furthermore, the genomes of E. faecalis and E. faecium encode two paralogs, MprF1 and MprF2, where MprF2 seems to play a primary role in PG aminoacylation in E. faecalis. The absence of L-PG from E. faecalis membrane was shown to increase bacterial susceptibility to cationic AMPs [302,303,501] (Table 3). These findings propose a model where the targeted action of AMPs correlates with their effectiveness in killing bacteria and interfering with the assembly of virulence factors. Moreover, Reyes et al. (2015) demonstrated that the deletion of the gene encoding the response regulator LiaR, a component of the LiaFSR system controlling cell envelope homeostasis, from daptomycin-resistant E. faecalis, altered the CL microdomain localization in the cell membrane. This deletion led to the hypersusceptibility to daptomycin and telavancin as well as to various AMPs of diverse origins and mechanisms of action. The observed changes in susceptibility to these AMPs were associated with reduced virulence in a Caenorhabditis elegans model [305]. Recent in vitro evolution experiments targeting daptomycin resistance unveiled numerous novel mutations associated with resistance, notably including mutations in the protease-encoding ftsH gene, which were found to be enriched exclusively in a ΔmprF1 ΔmprF2 background (slowed evolution to daptomycin resistance). Moreover, it was shown that FtsH indirectly modulates the levels of the chaperone operon repressor (HrcA), consequently affecting the pace of daptomycin resistance evolution [304]. Interestingly, Klein et al. (2008) demonstrated that membrane PGs from both the outer and inner membranes of P. aeruginosa, when cultured in acidic conditions, undergo substitution with alanine to form alanyl-phosphatidylglycerols (APGs). This substitution is catalyzed by an APG synthase encoded by the PA0920 locus identified in the genome of P. aeruginosa, which utilizes alanine tRNAs as alanine donors. The addition of alanine to PG reduces membrane permeability and neutralizes the negative charge of PG, consequently reducing its affinity for protamine [252,253]. Recently, an orthologue of S. aureus MprF, known as LysX, was identified in Mycobacterium tuberculosis, suggesting its potential membership in the aminoacyl-phosphatidylglycerol synthase family. LysX plays a critical role in Lys-PG formation, conferring resistance against cationic AMPs and a low pH. The deletion of lysX leads to disruptions in membrane potential, growth impairment, and defects in intracellular replication [500]. Additionally, another LysX orthologue, named LysX2, is exclusive to pathogenic mycobacteria and has been demonstrated to decrease the surface negative charge under acidic conditions, enhance bacterial cell viability at lethal pH levels, and hinder biofilm formation [502].

3.6. Active Efflux and Transport of AMPs

When the previously outlined mechanisms fail and AMPs amass on bacterial membranes, they might undergo conformational changes upon interacting with phospholipids. Upon reaching a certain threshold, they could breach the periplasm or cytoplasm via transient pore formation [503]. Even in such scenarios, bacteria retain a last resort mechanism to eliminate AMPs, orchestrated by efflux pumps and transporters. This includes the ability to either extrude AMPs that have crossed the outer membrane intact to prevent their intracellular accumulation [156,223,266,268,283,504,505,506,507,508], or transport them intracellularly for degradation by cytoplasmic proteases, such as those facilitated by the Sap (Sensitive to Antimicrobial Peptides) transporter (for a detailed review, see [471]). Efflux pumps are active transporters able to recognize and expel a wide array of structurally diverse compounds, depending on their substrate specificity conferred by the presence of one or multiple drug-binding sites [509]. These pumps not only play a role in resistance against toxic agents, but also in regulating physiological functions and can contribute to the formation of biofilm and virulence [510]. Efflux pumps are classified into seven widely acknowledged categories: (i) ATP-binding cassette (ABC) superfamily; (ii) major facilitator system (MFS) superfamily; (iii) multidrug and toxic-compound extrusion (MATE) family; (iv) drug/metabolite transporter superfamily (DMT), which encompasses the small multidrug resistance (SMR) family; (v) resistance–nodulation–division (RND) superfamily; (vi) proteobacterial antimicrobial compound efflux (PACE) family; and (vii) the p-Aminobenzoyl-glutamate transporter (AbgT) family [511]. Many studies highlighted the implication of ABC-type transporters and RND-type efflux pumps in bacterial resistance to AMPs, while only few others reported the contribution of MFS family members to this phenotype [510] (Table 2 and Table 3).
Bacterial ABC transporters are categorized into exporters and importers [512]. Among them, those responsible in transporting AMPs display diverse structural configurations, such as the tripartite structures found in Gram-negative bacteria, as seen in the MacAB–TolC efflux pump [513]. Peptide substrates recognized by MacB and its homologues typically resemble small disulfide-bonded peptides, reminiscent of certain AMPs like mammalian defensins. The expression of MacB-like proteins has been shown to impact the survival of Salmonella and S. pyogenes within macrophages [265,319] (Table 2 and Table 3). Notably, the transcription of macAB is upregulated during macrophage infection and following exposure to AMPs, a response supported by PhoPQ. The constitutive expression of macAB enhances the survival of Salmonella in the presence of the AMP C18G [266]. Moreover, the multifunctional inner membrane protein complex Sap transporter serves the purpose of importing various AMPs into the cytoplasm of various Gram-negative bacterial pathogens (Table 2). It consists of SapA, which functions as a periplasmic solute-binding protein, along with SapB and SapC, acting as transmembrane proteins to create a pore in the inner membrane. Additionally, SapD and SapF serve as ATPase subunits, while SapA is an integral membrane protein likely associated with SapC [267]. SapA binds directly to AMP and transports it from the periplasm to the SapBCDF complex for further transfer into the bacterial cytoplasm, where the AMPs undergo degradation and their amino acids are recycled [196,197,200,213,267,514,515,516].
In Gram-positive bacteria, particularly S. aureus, the GraSR TCS controls the expression of the adjacent ABC transporter, vraFG, which, while not conferring resistance independently, plays a vital role in sensing cationic AMPs and activating the GraR-dependent transcription of dlt and mprF genes [376,517] (Table 3).
The efflux pumps of the RND superfamily, renowned for their polyselective nature, have emerged as clinically significant determinants, conferring MDR in numerous bacterial pathogens [509]. RND pumps, powered by the proton motive force (PMF) and reliant on the pH gradient across the inner membrane, form a tripartite structure with periplasmic adaptor proteins and outer membrane protein channels, such as TolC, to facilitate the export of substrates from the cell [509,511,518]. The most studied RND systems involved in AMP resistance include the AcrAB–TolC of Enterobacterales species, Mex of P. aeruginosa, Vex of Vibrio, and Ade of Acinetobacter (Table 2). Moreover, there is increasing evidence supporting the capability of the MtrCDE pump to recognize and extrude AMPs in various bacteria, including N. gonorrhoeae, N. meningitidis, and H. ducreyi [197,221,222,223,224,229] (Table 2). MtrD appears to be the only RND pump protein for which mutagenesis studies have identified key residues crucial for the binding of an AMP, polymyxin B [222,519], suggesting that the latter is an actively extruded substrate of this pump.
Despite its broad specificity, the involvement of AcrAB–TolC in the AMP resistance of E. coli has been a contentious subject, as evidenced by various studies. For instance, a study performed in 2009 concluded that LL-37, polymyxin B, and various defensins were not substrates of AcrAB–TolC [520]. However, a subsequent study in 2010 suggested otherwise, indicating increased susceptibility to these AMPs in acrAB-deficient mutants [521]. Warner and Levy (2010) revealed that the discrepancies in AMP susceptibility among WT and the acrAB and tolC mutants noted by Rieg and Warner were attributed to variations in microbiological media [521], with differing ion concentrations potentially affecting AMP stability and binding to negatively charged surfaces [522,523]. Moreover, the increased susceptibility of tolC-deficient mutants implied the participation of other efflux pumps utilizing TolC as a mediator for AMP efflux. Additionally, the CpxR/CpxA TCS has been demonstrated to upregulate the multidrug resistance cascade, enhancing E. coli resistance to a model AMP [191]. Despite these findings, there remains a gap in the literature regarding specific studies identifying AMP-binding sites within the AcrAB efflux pump in Gram-negative bacteria, unlike the situation with the MtrCDE pump. The majority of research on AcrAB centers on its interactions with antibiotics and biocides rather than AMPs [511,518,524,525]. Considering the evidence demonstrating that AcrAB–TolC can expel large molecules, such as macrolide antibiotics (with a molecular weight smaller than 1000 Da) [526,527], it is plausible that cyclic AMPs like colistin and polymyxin B, sharing a similar size (approximately MW 1200 Da), could also be substrates of this pump. However, this hypothesis does not apply to larger AMPs such as LL-37 (4493 Da) or defensins (3000–5000 Da) [471]. Numerous studies utilizing isogenic acrAB mutants (lacking AcrAB proteins), loss of efflux function mutations (possessing non-functional AcrAB proteins), and efflux pump inhibitors have shown that AcrAB contributes to polymyxin resistance in many Gram-negative pathogens (Table 2). Yet, it remains unclear whether this contribution arises from a direct efflux mechanism of the AMP itself, or indirectly through the transport of other substrates affecting the bacterial surface charge, or alterations in membrane permeability beyond simply the loss of efflux, which may also affect susceptibility [521,528]. This uncertainty persists due to the physiological multifunctionality of AcrAB and its dependence on stress-induced regulators, like CpxR, Mar, Ram, Rob, and Sox [509,518]. For instance, an acrB knockout mutant of K. pneumoniae displayed a significantly increased susceptibility to AMPs [214], as well as a reduced ability to cause pneumonia in a mouse model, compared to the WT strain [214] (Table 2). In highly colistin-resistant clinical strains of K. pneumoniae, a newly discovered unusual RND pump (H239_3064), which lacks a known periplasmic adaptive protein but shares 49% amino acid identity with the parental AcrB, has been associated with colistin resistance. This pump is highly induced by the CrrAB TCS, which harbors missense amino acid substitutions in the CrrB histidine kinase, leading to a reduced susceptibility to colistin. The deletion of H239_3064 in the crrB background resulted in an 8-fold decrease in colistin MIC [215] (Table 2).
In addition to Klebsiella, other members of the ESKAPE pathogens, such as Enterobacter asburiae and Enterobacter cloacae, partly depend on the AcrAB-TolC pump to exhibit polymyxin heteroresistance [168]. An AcrB ortholog, KexD, similar to H239_3064 of K. pneumoniae, was discovered in various Enterobacter species, including E. bugandensis, and linked to colistin resistance [529] (Table 2). KexD interacts with the predicted small transmembrane protein CrrC via its membrane domain. Comparable interactions were simulated for AcrB and AcrD efflux pumps. Recently, García-Romero et al. (2024) proposed a model in which drug efflux, enhanced by CrrC interactions with major efflux pumps, along with lipid A modifications, regulated by PhoPQ and CrrAB, confer high-level resistance and heteroresistance to polymyxin B in E. bugandensis. They demonstrated that KexD and CrrC play a significant role in polymyxin B resistance and heteroresistance, with their genes being highly overexpressed in response to this AMP. Through co-immunoprecipitation experiments and proteomic analysis, they suggested that TolC and AcrA interact with KexD, indicating that KexD likely functions with AcrA and TolC. This conclusion is supported by the increased polymyxin B susceptibility of the acrAB and tolC mutants. Additionally, the increased susceptibility to polymyxin B of crrCkexD and crrC mutants does not depend on L-Ara4N modifications. Moreover, the stronger effect of crrCkexD on reducing polymyxin B resistance suggests that both proteins have a synergistic effect [530]. Colistin heteroresistance in a clinical isolate of A. baumannii was also associated with the overexpression of another RND component, AdeB, highlighting the involvement of the AdeABC system when exposed to colistin [147] (Table 2).
In P. aeruginosa, although MexA shares a high similarity with AcrA, and MexB with AcrB (71% and 89%, respectively), Pamp et al. (2008) reported that the adaptation of P. aeruginosa to colistin in vitro depends on the MexAB-OprM efflux pump only in bacteria growing as biofilms and has not been observed in the same bacteria maintained in a planktonic state. Specifically, the induction of the MexAB-OprM pump was observed only within the metabolically active bacterial subpopulation located on the surface of biofilms [254]. However, other efflux pumps in P. aeruginosa, such as MexCD–OprJ and MexXY–OprM, were found ineffective in expelling polymyxin B [255]. In C. jejuni, a mutation in the inner membrane component-encoding cmeE gene of the CmeDEF efflux pump led to increased sensitivity to polymyxin B. This sensitivity was observed consistently in the cmeF/cmeB double mutants, which lack the inner membrane component of the primary efflux pump CmeABC [165] (Table 2). Furthermore, a single mutant affecting acrA and a double mutant affecting both acrA and the membrane fusion protein-encoding mdtA gene of the MdtABC RND efflux pump in the insect pathogen Photorhabdus laumondii exhibited at least an 8- to 16-fold decrease in colistin and polymyxin B MICs, while resistance to cecropins A and C was not affected compared to the parental strain. The authors suggested that AcrAB indirectly contributes to AMP tolerance at supra-physiological concentrations, likely by maintaining membrane integrity rather than through direct efflux, which is probably irrelevant for Photorhabdus ability to resist AMP-induced killing in the insect hemolymph [231,232] (Table 2).
The efflux of polymyxin B, cecropin P1, and melittinin in Yersinia enterocolitica seems to occur through the RosA MFS efflux pump coupled to the RosB potassium antiporter, involving TolC. The RosA/RosB regulon appears to be inducible by cationic AMPs and temperature (37 °C), enabling bacterial adaptation to an acidic environment rich in cationic AMPs, such as the phagolysosome [282]. Furthermore, the MFS-type emrB knockout mutation in A. baumannii resulted in an increased susceptibility to colistin compared to the WT strain [146]. Also, the EmrAB–TolC system confers resistance to protamine in E. coli, as demonstrated by the notable difference in survival rates between the emrB (20%) and tolC (0%) mutants when exposed to protamine [191]. The EmrAB/TolC and AcrAB/TolC efflux systems simultaneously contribute to E. coli protamine resistance in a CpxR/CpxA-dependent manner through the activation of mar operon transcription. Moreover, CpxR/CpxA induces the transcription of the aroK gene, enhancing the production of aromatic metabolites that release the MarR repressor from the marO site, thereby increasing marA expression and subsequently triggering the efflux pump-mediated multidrug resistance cascade [191].
MFS efflux systems have also been associated with cationic AMP resistance in Gram-positive bacteria, such as staphylococci [531]. The pSK1plasmid confers the resistance of staphylococci to numerous organic cations through an MFS pump encoded by a plasmid gene. Kupferwasser et al. (1999) showed that transferring the qacA gene from S. aureus to another more sensitive parental strain increases its resistance to tPMP-1 (but not to α-defensins and protamine). Additionally, the mere presence of the qacA gene product seems sufficient to confer this resistance, independently of the efflux mechanism itself, especially as tPMPs do not appear to be the major substrate of this pump [375]. This possibility suggests that the QacA protein could modify the composition of the cytoplasmic membrane to avoid disruption by tPMP-1.
Another system relevant to AMP resistance is the ubiquitous TrkG/H–TrkA potassium uptake system, functioning through K+ uptake coupled with H+ symport [532]. Notably, a trkA isogenic mutant of the highly virulent marine bacterium Vibrio vulnificus displayed increased susceptibility to protamine and polymyxin B compared to the WT. Furthermore, infection trials revealed reduced virulence in both normal and iron-treated mice following the intraperitoneal or subcutaneous administration of this mutant [279] (Figure 5A).

3.7. Alteration of Membrane Energetics and Resistance to AMPs

The activities of several types of AMPs are affected by the transmembrane potential [533]. Some microbial pathogens utilize the regulation of their energy status by altering membrane energetics as a strategy to evade the action of AMPs. For instance, type II defensins show significantly lower activity against organisms with an inherently reduced transmembrane potential Δψ (due to respiration deficiency), or those capable of adapting to such conditions. Therefore, S. aureus strains with constitutively reduced Δψ demonstrate decreased sensitivity to specific AMPs [534]. Furthermore, cytochrome c oxidase modulates the transmembrane potential in C. jejuni, affecting membrane permeability to polymyxin B [164]. Disulfide reductase has also been implicated in AMP resistance in S. enterica [535].

3.8. Cellular Differentiations and Resistance to AMPs

Some bacteria exhibit susceptibility to AMPs in one morpho-physiological state while demonstrating resistance in another. This difference in susceptibility is attributed to variations in behavior between the planktonic and aggregated forms of bacteria. Notably, the development of small colony variants (SCVs) and the formation of biofilms play significant roles in this resistance mechanism [254,536]. These phenomena, observed in various pathogenic bacteria, directly or indirectly contribute to their ability to resist the antimicrobial effects of AMPs, posing challenges for effective treatment strategies.

3.8.1. Small Colony Variants and Niche-Specific Resistance to AMPs

Small colony variants (SCVs) are a subset of bacteria characterized by slow growth and distinct phenotypic and pathogenic features under environmental stress. They exhibit unique colony morphology and biochemical traits, making their identification challenging. Clinically, SCVs demonstrate reduced susceptibility to antimicrobial agents compared to typical bacterial strains and may contribute to latent or recurrent infections [536]. Numerous studies have delved into the resistance of S. aureus SCVs against AMPs. The SCV phenotype in S. aureus can be triggered within the intracellular microenvironment at a frequency of 10−3, significantly higher than the spontaneous rate of less than 10−7, aiding in the evasion of the host immune response [537]. There is evidence indicating that reductions in cellular energetics detrimentally affect the effectiveness of AMPs relying on Δψ for their mechanism of action or target affinity. Proctor et al. (1998) illustrated how S. aureus reversibly adopts this advantageous strategy to survive within the microenvironment of vascular endothelial cells by forming SCV colonies [537,538,539]. These variants typically exhibit atypical morphology, electron transport defects, and diminished envelope affinity for cationic antimicrobial agents. Mutants of S. aureus strains with reduced Δψ display significantly higher resistance to AMPs compared to their parental counterparts [538,539]. Furthermore, clinical S. aureus strains sensitive to tPMP1 exhibit slower multiplication rates in cardiac tissues and splenic abscesses, whereas no discrepancy in multiplication is observed at the renal level. This discrepancy is attributed to the relatively diminished protective role of tPMP1 in environments with high osmotic forces, such as the kidneys [35]. A similar mechanism is observed in P. aeruginosa, frequently infecting tissues where salt transport dysfunctions elevate the local ion content, particularly in the respiratory tract of cystic fibrosis patients, where airway fluids contain abundant salts and AMPs [35]. Such observations underscore the risk of certain pathogens exploiting specific tissues or physiological microenvironments to evade the action of AMPs.
Additionally, Samuelsen et al. (2005) demonstrated that S. aureus SCVs exhibit acquired resistance to lactoferricin B, primarily due to metabolic variations. Hemin, menadione, or thymidine SCV hemB auxotrophic mutants of S. aureus have been found to be more susceptible to AMPs compared to bacteria having one of these components [540]. Furthermore, Glaser et al. (2014) have shown that clinically derived S. aureus SCVs as well as a hemB auxotroph S. aureus SCVs are less susceptible to various human skin-derived AMPs, such as human β-defensin and LL-37 [541].

3.8.2. Biofilm Formation and Resistance to AMPs

Some AMPs, such as the human peptide LL-37 and a novel synthetic cationic peptide named 1037, comprising just nine amino acids, can hinder biofilm formation even at concentrations lower than their MIC. The 1037 peptide effectively hindered biofilm formation by more than 50%, notably against the Gram-negative pathogens P. aeruginosa and B. cenocepacia, as well as the Gram-positive bacterium L. monocytogenes [542]. However, bacterial implantation into biofilms triggers changes in gene expression, resulting, among other effects, in a greater resistance to antimicrobial agents compared to individual bacteria. Mechanisms behind biofilm-mediated AMR resistance involve multiple factors simultaneously, such as shielding by the polysaccharidic extracellular polymeric substance (EPS), nutrient deficiency and altered environment, the formation of persister cells, extracellular DNA release, stress response, and enhanced efflux pump activity [543,544,545,546]. The biofilm environment also increases the mutation rate and exchange of genetic material, enabling the transfer of resistance traits on mobile genetic elements [547,548].
P. aeruginosa produces significant amounts of a highly anionic exopolysaccharide known as alginate. Friedrich et al. (1999) purified alginate and demonstrated its interference with the activity of cationic AMPs [549] (Table 2). Alginic acid entraps bacterial cells, contributing to the formation of biofilms involved in the development and persistence of several chronic bacterial infections in animals and humans, particularly in the lungs during cystic fibrosis [550,551]. The fact that the activity of several cationic AMPs is inhibited in the presence of mono- and divalent cations could enhance the protective role of alginate [35]. However, S. epidermidis produces an atypical anionic extracellular polysaccharide called PIA (polysaccharide intercellular adhesin), which repels AMPs secreted onto the skin (except for the anionic dermcidin) in the presence of low saline concentrations (Table 3). However, it enables resistance to dermcidin only in salt-rich microenvironments, such as the skin [378]. A study performed by Begun et al. (2007) examined the role of PIA in protecting against C. elegans immune system and the production of lethal infections [552]. Induced resistance factors in biofilms include those resulting from induction by the antimicrobial agent itself [553]. The slow growth (quiescence or dormancy) of the bacteria inside biofilms, particularly in the deeper layers, likely stems from nutrient and oxygen deficiency, leading to an extremely reduced metabolic state. This reduced metabolic activity renders the bacteria less receptive to antimicrobial agents compared to metabolically active ones [554]. Efflux pumps are highly active in biofilms, essential for their formation and/or maintenance, and responsible for their multiple antibiotic tolerance/resistance [555,556]. Pamp et al. (2008) reported that the adaptation of P. aeruginosa biofilms to colistin in vitro occurs through the active expulsion of this peptide via the MexAB-OprM efflux pump. This mechanism is exclusively induced within biofilms and has not been observed in the same bacteria maintained in the planktonic state. Specifically, the induction of the MexAB-OprM pump has been observed only in the metabolically active bacterial subpopulation located on the biofilm surface [254]. Conversely, Folkesson et al. (2008) demonstrated that biofilm formation in E. coli confers a high tolerance to colistin on cells within the biofilm structure. However, this protection is conditional, depending on the structural and architectural organization of the biofilm, as well as the induction of specific tolerance mechanisms (not by colistin itself) involving the BasS/BasR regulatory system (Salmonella pmrAB) controlling the expression of the ybf operon (Salmonella pmrHFIJKLM) [557]. Furthermore, in the extracellular polysaccharide matrix of P. aeruginosa biofilms, much like in biofilms formed by other Gram-positive and Gram-negative bacteria, there is extracellular DNA (eDNA) among other components, likely released by bacteria or dead immune cells [558]. This eDNA (negatively charged) plays a role in shielding cationic AMPs [559]. When present at physiological concentrations, eDNA causes bacterial lysis by chelating the cations that stabilize bacterial LPS and cell membranes. Therefore, the genomic DNA released in sub-inhibitory quantities by lysed bacteria in the biofilm creates a cation-deficient environment. This deficiency leads to the induction of resistance to cationic AMPs (colistin and polymyxin B) via the activation of the PhoPQ and PmrAB TCSs [560].
Many studies have reported associations between biofilms and swarming motility [263,561]. Notably, the strong resistance to colistin and polymyxin B of swarming bacteria of S. Typhimurium has been linked to the induction under “swarming”-promoting conditions of the pmrHFIJKLM operon involved in polymyxin B resistance [562].

3.9. Manipulation of Host Cell AMP Production

In the absence of infections, the epithelia of higher animals and leukocytes produce small quantities of AMPs. However, this production is strongly induced in response to bacterial structures such as LPS and teichoic acids recognized by Toll-like receptors and may involve the NF-κB-dependent signaling pathway [563,564,565]. Indeed, it has been demonstrated that the LPSs of several bacteria, such as E. coli, S. Typhimurium, and P. aeruginosa, are capable of bypassing the induction of genes encoding AMPs [566]. Moreover, S. pyogenes is a very weak inducer of AMP production [320]. This observation suggests that S. pyogenes avoids recognition receptors or actively represses the induction pathway of β-defensin-2. An interesting study of the human gastrointestinal pathogen Shigella dysenteriae has revealed that the expression of LL-37 and β-defensin-1 is transcriptionally repressed in the early stages of infection. This plasmid-encoded repression was first observed in biopsies from patients with bacillary dysentery and then confirmed by experiments on cell cultures of monocytes and epithelial cells (Table 2). On the other hand, Arbibe et al. (2007) showed that virulent strains of S. flexneri repress the transcription of several genes encoding cationic AMPs of innate immunity (in particular, β-defensin and HBD-3, which are specifically active against S. flexneri) following the in vitro infection of polarized human intestinal cells [273]. This repression depends on the regulator MxiE of S. flexneri known to activate more than 10 genes encoding effectors of the type III secretion system (TTSS), including OpsF and OspG, which target signaling pathways via NF-κB and MAPK [273] (Table 2). The inhibition of cationic AMP production via MxiE has been confirmed in vivo at the transcriptional and translational levels in a murine model of human intestinal xenotransplantation [567]. Moreover, Krishnenduno et al. (2008) showed that the enteric pathogens V. cholerae and enterotoxigenic E. coli (ETEC) repress the expression of LL-37 and HβD-1 in intestinal epithelial cells while acting on signaling pathways known in eukaryotes. Indeed, cholera toxin (CT) and heat-labile toxin (LT), which are the two major virulence proteins of V. cholerae and ETEC, respectively, are primarily responsible for these repressions both in vitro and in vivo. CT represses the expression of AMPs at the transcriptional level by activating several intracellular signaling pathways involving protein kinase A (PKA), ERK MAPK, and Cox-2 [192] (Table 2).
Contrary to the inhibition of AMP production, another bacterial survival strategy may involve the stimulation of host mechanisms that counter-regulate the response by AMPs. Such a process has been demonstrated in the lungs of patients with cystic fibrosis, where the causative pathogen P. aeruginosa stimulates the accumulation of host cysteine proteases (cathepsins B, L, and S) in the respiratory tract fluids. These proteases degrade and inactivate human β-defensins 2 and 3, thereby promoting chronic infection and bacterial colonization characteristic of the disease [257].
Ureaplasma spp. are commonly found in the urogenital tract of adults and are associated with serious infections. Infections with U. parvum induce modifications in histone H3K9, particularly a decrease in histone acetylation. This alteration is linked to a downregulation of AMP gene expression, enabling bacteria to evade the host immune response, thereby leading to chronic infection [274] (Table 2). A similar mechanism involving histone modification, including the phosphorylation of H3 and deacetylation of H4, is observed in L. monocytogenes. This modification is induced by the Listeria toxin listeriolysin O in a pore-forming-independent manner. Comparable mechanisms are employed by Clostridium perfringens and S. pneumoniae, which utilize perfringolysin and pneumolysin, respectively [568]. Moreover, Chlamydia inhibits NF-κB activation by blocking the degradation of the NF-κB retention factor, IκBα, and preventing the nuclear translocation of NF-κB, ultimately repressing NF-κB transcription [569].

4. Conclusive Remarks and Future Directions

AMPs offer great potential in combating bacterial infections, yet they encounter significant challenges due to the easy development of resistance mechanisms, even to last-resort treatment options against MDR strains such as colistin. It is noteworthy that both Gram-negative and Gram-positive bacteria, whether pathogenic or commensals, have adopted various alterations or remodeling of their envelope structures to counteract AMP-mediated killing. As illustrated in this review, these cell surface modifications are evident across a wide range of bacterial species, including members of Enterobacterales, P. aeruginosa, Neisseria sp., Bacillus spp., Lactobacillus spp., Streptococcus spp., Staphylococcus spp., L. monocytogenes, and even Mycobacterium spp. These examples, along with other non-associated cell wall modifications, underscore the convergence of resistance mechanisms to AMPs by the common selective pressure exerted by these peptides. Therefore, modifications of bacterial cell surfaces appear to be the most conserved and effective resistance mechanism to AMPs. This is supported by numerous experimental observations indicating that spontaneous mutations and those introduced in structural and regulatory genes involved in these processes lead to the most in vitro AMP-hypersensitive phenotypes, along with the severe attenuation or abolition of virulence in various animal models. Understanding the intricate molecular pathways employed by bacteria to evade AMPs is paramount in developing effective strategies to combat resistance. Interdisciplinary approaches, such as genomics, metabolomics, structural biology, and bioinformatics, offer promising avenues to unravel the complexity of AMP resistance mechanisms.
From the limited research conducted to examine the speed of AMP resistance emergence and how resistance mechanisms would affect the fitness and virulence of mutants, it is well established that acquiring such resistance through mutation is not challenging for bacteria, with multiple facilitating mechanisms characterized to date. Some of these mechanisms have a minimal impact on bacterial fitness, thanks to compensatory mutations that can mitigate the associated fitness cost. These resistance mutations often confer cross-resistance to human AMPs, posing concerns about the therapeutic use of AMPs potentially selecting for mutants showing broad cross-resistance to diverse AMPs. This suggests that the routine clinical use of AMPs could select for resistant strains capable of evading the immune system. Researchers have emphasized the need for careful study and monitoring as AMPs are increasingly used in clinical settings. Therefore, thorough investigation and monitoring are essential to understand the implications of AMP resistance and to ensure their effective clinical use. Moving forward, novel therapeutic interventions should not only target bacterial resistance mechanisms but also strive to minimize the likelihood of resistance development through innovative peptide design and delivery strategies. Additionally, exploring alternative approaches, such as combination therapies with conventional antibiotics or AMP adjuvants, could enhance antimicrobial efficacy while mitigating the risk of resistance emergence.
Finally, fostering collaborative efforts across scientific disciplines is crucial to staying ahead of evolving resistance mechanisms and ensuring the continued efficacy of AMPs in combating bacterial infections. By embracing innovative future perspectives and leveraging synergistic approaches, we can enhance our arsenal against AMP resistance, ultimately advancing the fight against infectious diseases.

Author Contributions

Conceptualization: Z.F. and Z.A.K.; validation: Z.F., J.-C.P., H.D., J.-M.S. and Z.A.K.; visualization: H.D., J.-M.S., J.-C.P. and Z.A.K.; investigation: Z.A.K. and Z.F.; writing—original draft preparation: L.T. and Z.A.K.; writing—review and editing: L.T., Z.F. and Z.A.K.; supervision: Z.F., J.-C.P. and Z.A.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. World Health Organization. New Report Calls for Urgent Action to Avert Antimicrobial Resistance Crisis. Available online: https://www.who.int/news/item/29-04-2019-new-report-calls-for-urgent-action-to-avert-antimicrobial-resistance-crisis (accessed on 6 April 2024).
  2. World Health Organization. Antimicrobial Resistance. Available online: https://www.who.int/news-room/fact-sheets/detail/antimicrobial-resistance (accessed on 7 April 2024).
  3. O’Neill, J. Review on Antimicrobial Resistance. Tackling Drug-Resistant Infections Globally. 2016. Available online: https://amr-review.org/sites/default/files/160525_Final%20paper_with%20cover.pdf (accessed on 6 April 2024).
  4. Ayukekbong, J.A.; Ntemgwa, M.; Atabe, A.N. The Threat of Antimicrobial Resistance in Developing Countries: Causes and Control Strategies. Antimicrob. Resist. Infect. Control 2017, 6, 47. [Google Scholar] [CrossRef] [PubMed]
  5. Pallett, S.J.C.; Boyd, S.E.; O’Shea, M.K.; Martin, J.; Jenkins, D.R.; Hutley, E.J. The Contribution of Human Conflict to the Development of Antimicrobial Resistance. Commun. Med. 2023, 3, 153. [Google Scholar] [CrossRef] [PubMed]
  6. Garg, S.K. Antibiotic Misuse during COVID-19 Pandemic: A Recipe for Disaster. Indian J. Crit. Care Med. Peer-Rev. Off. Publ. Indian Soc. Crit. Care Med. 2021, 25, 617–619. [Google Scholar] [CrossRef] [PubMed]
  7. Yu, G.; Baeder, D.Y.; Regoes, R.R.; Rolff, J. Predicting Drug Resistance Evolution: Insights from Antimicrobial Peptides and Antibiotics. Proc. R. Soc. B Biol. Sci. 2018, 285, 20172687. [Google Scholar] [CrossRef]
  8. Ciumac, D.; Gong, H.; Hu, X.; Lu, J.R. Membrane Targeting Cationic Antimicrobial Peptides. J. Colloid Interface Sci. 2019, 537, 163–185. [Google Scholar] [CrossRef]
  9. Zhang, Q.-Y.; Yan, Z.-B.; Meng, Y.-M.; Hong, X.-Y.; Shao, G.; Ma, J.-J.; Cheng, X.-R.; Liu, J.; Kang, J.; Fu, C.-Y. Antimicrobial Peptides: Mechanism of Action, Activity and Clinical Potential. Mil. Med. Res. 2021, 8, 48. [Google Scholar] [CrossRef]
  10. Mitra, J.B.; Sharma, V.K.; Kumar, M.; Mukherjee, A. Antimicrobial Peptides: Vestiges of Past or Modern Therapeutics? Mini-Rev. Med. Chem. 2020, 20, 183–195. [Google Scholar] [CrossRef]
  11. APD3. The Antimicrobial Peptide Database. 2023. Available online: https://aps.unmc.edu (accessed on 7 May 2024).
  12. Poger, D.; Pöyry, S.; Mark, A.E. Could Cardiolipin Protect Membranes against the Action of Certain Antimicrobial Peptides? Aurein 1.2, a Case Study. ACS Omega 2018, 3, 16453–16464. [Google Scholar] [CrossRef]
  13. Harrison, P.L.; Heath, G.R.; Johnson, B.R.G.; Abdel-Rahman, M.A.; Strong, P.N.; Evans, S.D.; Miller, K. Phospholipid Dependent Mechanism of Smp24, an α-Helical Antimicrobial Peptide from Scorpion Venom. Biochim. Biophys. Acta BBA—Biomembr. 2016, 1858, 2737–2744. [Google Scholar] [CrossRef]
  14. von Deuster, C.I.E.; Knecht, V. Antimicrobial Selectivity Based on Zwitterionic Lipids and Underlying Balance of Interactions. Biochim. Biophys. Acta BBA—Biomembr. 2012, 1818, 2192–2201. [Google Scholar] [CrossRef]
  15. Maturana, P.; Martinez, M.; Noguera, M.E.; Santos, N.C.; Disalvo, E.A.; Semorile, L.; Maffia, P.C.; Hollmann, A. Lipid Selectivity in Novel Antimicrobial Peptides: Implication on Antimicrobial and Hemolytic Activity. Colloids Surf. B Biointerfaces 2017, 153, 152–159. [Google Scholar] [CrossRef] [PubMed]
  16. Vignoli Muniz, G.S.; De la Torre, L.I.; Duarte, E.L.; Lorenzón, E.N.; Cilli, E.M.; Balan, A.; Lamy, M.T. Interaction of Synthetic Antimicrobial Peptides of the Hylin A1 Family with Models of Eukaryotic Structures: Zwitterionic Membranes and DNA. Biochem. Biophys. Rep. 2020, 24, 100827. [Google Scholar] [CrossRef] [PubMed]
  17. Schlame, M. Thematic Review Series: Glycerolipids. Cardiolipin Synthesis for the Assembly of Bacterial and Mitochondrial Membranes. J. Lipid Res. 2008, 49, 1607–1620. [Google Scholar] [CrossRef] [PubMed]
  18. Wang, H.; Zhang, C.; Li, M.; Liu, C.; Wang, J.; Ou, X.; Han, Y. Antimicrobial Peptides Mediate Apoptosis by Changing Mitochondrial Membrane Permeability. Int. J. Mol. Sci. 2022, 23, 12732. [Google Scholar] [CrossRef] [PubMed]
  19. Magana, M.; Pushpanathan, M.; Santos, A.L.; Leanse, L.; Fernandez, M.; Ioannidis, A.; Giulianotti, M.A.; Apidianakis, Y.; Bradfute, S.; Ferguson, A.L.; et al. The Value of Antimicrobial Peptides in the Age of Resistance. Lancet Infect. Dis. 2020, 20, e216–e230. [Google Scholar] [CrossRef] [PubMed]
  20. CDC. Are You Using Antibiotics Wisely? Available online: https://www.cdc.gov/antibiotic-use/do-and-dont.html (accessed on 8 April 2024).
  21. Hassan, S.N.; Ahmad, F. The Relevance of Antibiotic Supplements in Mammalian Cell Cultures: Towards a Paradigm Shift. Gulhane Med. J. 2020, 62, 224–230. [Google Scholar] [CrossRef]
  22. Peschel, A.; Sahl, H.-G. The Co-Evolution of Host Cationic Antimicrobial Peptides and Microbial Resistance. Nat. Rev. Microbiol. 2006, 4, 529–536. [Google Scholar] [CrossRef]
  23. Okdeh, N.; Mahfouz, G.; Harb, J.; Sabatier, J.-M.; Roufayel, R.; Gazo Hanna, E.; Kovacic, H.; Fajloun, Z. Protective Role and Functional Engineering of Neuropeptides in Depression and Anxiety: An Overview. Bioengineering 2023, 10, 258. [Google Scholar] [CrossRef]
  24. Fjell, C.D.; Hiss, J.A.; Hancock, R.E.W.; Schneider, G. Designing Antimicrobial Peptides: Form Follows Function. Nat. Rev. Drug Discov. 2012, 11, 37–51. [Google Scholar] [CrossRef]
  25. O’Brien, D.K.; Ribot, W.J.; Chabot, D.J.; Scorpio, A.; Tobery, S.A.; Jelacic, T.M.; Wu, Z.; Friedlander, A.M. The Capsule of Bacillus Anthracis Protects It from the Bactericidal Activity of Human Defensins and Other Cationic Antimicrobial Peptides. PLoS Pathog. 2022, 18, e1010851. [Google Scholar] [CrossRef]
  26. Rima, M.; Rima, M.; Fajloun, Z.; Sabatier, J.-M.; Bechinger, B.; Naas, T. Antimicrobial Peptides: A Potent Alternative to Antibiotics. Antibiotics 2021, 10, 1095. [Google Scholar] [CrossRef] [PubMed]
  27. Spohn, R.; Daruka, L.; Lázár, V.; Martins, A.; Vidovics, F.; Grézal, G.; Méhi, O.; Kintses, B.; Számel, M.; Jangir, P.K.; et al. Integrated Evolutionary Analysis Reveals Antimicrobial Peptides with Limited Resistance. Nat. Commun. 2019, 10, 4538. [Google Scholar] [CrossRef]
  28. Jangir, P.K.; Ogunlana, L.; MacLean, R.C. Evolutionary Constraints on the Acquisition of Antimicrobial Peptide Resistance in Bacterial Pathogens. Trends Microbiol. 2021, 29, 1058–1061. [Google Scholar] [CrossRef] [PubMed]
  29. Kintses, B.; Méhi, O.; Ari, E.; Számel, M.; Györkei, Á.; Jangir, P.K.; Nagy, I.; Pál, F.; Fekete, G.; Tengölics, R.; et al. Phylogenetic Barriers to Horizontal Transfer of Antimicrobial Peptide Resistance Genes in the Human Gut Microbiota. Nat. Microbiol. 2019, 4, 447–458. [Google Scholar] [CrossRef] [PubMed]
  30. Yang, Q.; Li, M.; Spiller, O.B.; Andrey, D.O.; Hinchliffe, P.; Li, H.; MacLean, C.; Niumsup, P.; Powell, L.; Pritchard, M.; et al. Balancing Mcr-1 Expression and Bacterial Survival Is a Delicate Equilibrium between Essential Cellular Defence Mechanisms. Nat. Commun. 2017, 8, 2054. [Google Scholar] [CrossRef] [PubMed]
  31. Andersson, D.I.; Hughes, D.; Kubicek-Sutherland, J.Z. Mechanisms and Consequences of Bacterial Resistance to Antimicrobial Peptides. Drug Resist. Updat. 2016, 26, 43–57. [Google Scholar] [CrossRef] [PubMed]
  32. Batoni, G.; Maisetta, G.; Lisa Brancatisano, F.; Esin, S.; Campa, M. Use of Antimicrobial Peptides Against Microbial Biofilms: Advantages and Limits. Curr. Med. Chem. 2011, 18, 256–279. [Google Scholar] [CrossRef] [PubMed]
  33. Nizet, V. Antimicrobial Peptide Resistance Mechanisms of Human Bacterial Pathogens. Curr. Issues Mol. Biol. 2006, 8, 11–26. [Google Scholar] [CrossRef]
  34. Ghosh, C.; Haldar, J. Membrane-Active Small Molecules: Designs Inspired by Antimicrobial Peptides. ChemMedChem 2015, 10, 1606–1624. [Google Scholar] [CrossRef]
  35. Yeaman, M.R.; Yount, N.Y. Mechanisms of Antimicrobial Peptide Action and Resistance. Pharmacol. Rev. 2003, 55, 27–55. [Google Scholar] [CrossRef]
  36. Loutet, S.A.; Valvano, M.A. Extreme Antimicrobial Peptide and Polymyxin B Resistance in the Genus Burkholderia. Front. Microbiol. 2011, 2, 159. [Google Scholar] [CrossRef]
  37. Olaitan, A.O.; Morand, S.; Rolain, J.-M. Mechanisms of Polymyxin Resistance: Acquired and Intrinsic Resistance in Bacteria. Front. Microbiol. 2014, 5, 643. [Google Scholar] [CrossRef]
  38. Ghimire, J.; Guha, S.; Nelson, B.J.; Morici, L.A.; Wimley, W.C. The Remarkable Innate Resistance of Burkholderia Bacteria to Cationic Antimicrobial Peptides: Insights into the Mechanism of AMP Resistance. J. Membr. Biol. 2022, 255, 503–511. [Google Scholar] [CrossRef]
  39. Malott, R.J.; Steen-Kinnaird, B.R.; Lee, T.D.; Speert, D.P. Identification of Hopanoid Biosynthesis Genes Involved in Polymyxin Resistance in Burkholderia Multivorans. Antimicrob. Agents Chemother. 2012, 56, 464–471. [Google Scholar] [CrossRef]
  40. La, M.-V.; Raoult, D.; Renesto, P. Regulation of Whole Bacterial Pathogen Transcription within Infected Hosts. FEMS Microbiol. Rev. 2008, 32, 440–460. [Google Scholar] [CrossRef]
  41. Groisman, E.A. The Pleiotropic Two-Component Regulatory System PhoP-PhoQ. J. Bacteriol. 2001, 183, 1835–1842. [Google Scholar] [CrossRef]
  42. Erickson, K.D.; Detweiler, C.S. The Rcs Phosphorelay System Is Specific to Enteric Pathogens/Commensals and Activates ydeI, a Gene Important for Persistent Salmonella Infection of Mice. Mol. Microbiol. 2006, 62, 883–894. [Google Scholar] [CrossRef]
  43. Fernández, L.; Jenssen, H.; Bains, M.; Wiegand, I.; Gooderham, W.J.; Hancock, R.E.W. The Two-Component System CprRS Senses Cationic Peptides and Triggers Adaptive Resistance in Pseudomonas Aeruginosa Independently of ParRS. Antimicrob. Agents Chemother. 2012, 56, 6212–6222. [Google Scholar] [CrossRef]
  44. Santos, D.E.S.; Pol-Fachin, L.; Lins, R.D.; Soares, T.A. Polymyxin Binding to the Bacterial Outer Membrane Reveals Cation Displacement and Increasing Membrane Curvature in Susceptible but Not in Resistant Lipopolysaccharide Chemotypes. J. Chem. Inf. Model. 2017, 57, 2181–2193. [Google Scholar] [CrossRef]
  45. Jacob-Dubuisson, F.; Mechaly, A.; Betton, J.-M.; Antoine, R. Structural Insights into the Signalling Mechanisms of Two-Component Systems. Nat. Rev. Microbiol. 2018, 16, 585–593. [Google Scholar] [CrossRef]
  46. Bader, M.W.; Sanowar, S.; Daley, M.E.; Schneider, A.R.; Cho, U.; Xu, W.; Klevit, R.E.; Le Moual, H.; Miller, S.I. Recognition of Antimicrobial Peptides by a Bacterial Sensor Kinase. Cell 2005, 122, 461–472. [Google Scholar] [CrossRef]
  47. Vaara, M.; Vaara, T.; Jensen, M.; Helander, I.; Nurminen, M.; Rietschel, E.T.; Mäkelä, P.h. Characterization of the Lipopolysaccharide from the Polymyxin-Resistant pmrA Mutants of Salmonella Typhimurium. FEBS Lett. 1981, 129, 145–149. [Google Scholar] [CrossRef]
  48. Groisman, E.A.; Parra-Lopez, C.; Salcedo, M.; Lipps, C.J.; Heffron, F. Resistance to Host Antimicrobial Peptides Is Necessary for Salmonella Virulence. Proc. Natl. Acad. Sci. USA 1992, 89, 11939–11943. [Google Scholar] [CrossRef]
  49. Sun, J.; Xia, Y.; Li, D.; Du, Q.; Liang, D. Relationship between Peptide Structure and Antimicrobial Activity as Studied by de Novo Designed Peptides. Biochim. Biophys. Acta BBA—Biomembr. 2014, 1838, 2985–2993. [Google Scholar] [CrossRef]
  50. Lyu, Z.; Yang, P.; Lei, J.; Zhao, J. Biological Function of Antimicrobial Peptides on Suppressing Pathogens and Improving Host Immunity. Antibiotics 2023, 12, 1037. [Google Scholar] [CrossRef]
  51. Iksanova, A.M.; Arzumanian, V.G.; Konanykhina, S.Y.; Samoylikov, P.V. Antimicrobial Peptides and Proteins in Human Biological Fluids. Microbiol. Indep. Res. J. MIR J. 2022, 9, 37–55. [Google Scholar] [CrossRef]
  52. Wang, G. Unifying the Classification of Antimicrobial Peptides in the Antimicrobial Peptide Database. Methods Enzymol. 2022, 663, 1–18. [Google Scholar] [CrossRef]
  53. Saint Jean, K.D.; Henderson, K.D.; Chrom, C.L.; Abiuso, L.E.; Renn, L.M.; Caputo, G.A. Effects of Hydrophobic Amino Acid Substitutions on Antimicrobial Peptide Behavior. Probiotics Antimicrob. Proteins 2018, 10, 408–419. [Google Scholar] [CrossRef]
  54. Ambroggio, E.E.; Separovic, F.; Bowie, J.H.; Fidelio, G.D.; Bagatolli, L.A. Direct Visualization of Membrane Leakage Induced by the Antibiotic Peptides: Maculatin, Citropin, and Aurein. Biophys. J. 2005, 89, 1874–1881. [Google Scholar] [CrossRef]
  55. Sneideris, T.; Erkamp, N.A.; Ausserwöger, H.; Saar, K.L.; Welsh, T.J.; Qian, D.; Katsuya-Gaviria, K.; Johncock, M.L.L.Y.; Krainer, G.; Borodavka, A.; et al. Targeting Nucleic Acid Phase Transitions as a Mechanism of Action for Antimicrobial Peptides. Nat. Commun. 2023, 14, 7170. [Google Scholar] [CrossRef]
  56. Runti, G.; Lopez Ruiz, M. del C.; Stoilova, T.; Hussain, R.; Jennions, M.; Choudhury, H.G.; Benincasa, M.; Gennaro, R.; Beis, K.; Scocchi, M. Functional Characterization of SbmA, a Bacterial Inner Membrane Transporter Required for Importing the Antimicrobial Peptide Bac7(1-35). J. Bacteriol. 2013, 195, 5343–5351. [Google Scholar] [CrossRef]
  57. Akbari, R.; Hakemi Vala, M.; Hashemi, A.; Aghazadeh, H.; Sabatier, J.-M.; Pooshang Bagheri, K. Action Mechanism of Melittin-Derived Antimicrobial Peptides, MDP1 and MDP2, de Novo Designed against Multidrug Resistant Bacteria. Amino Acids 2018, 50, 1231–1243. [Google Scholar] [CrossRef]
  58. Li, Q.; Montalban-Lopez, M.; Kuipers, O.P. Increasing the Antimicrobial Activity of Nisin-Based Lantibiotics against Gram-Negative Pathogens. Appl. Environ. Microbiol. 2018, 84, e00052-18. [Google Scholar] [CrossRef]
  59. Majewska, M.; Zamlynny, V.; Pieta, I.S.; Nowakowski, R.; Pieta, P. Interaction of LL-37 Human Cathelicidin Peptide with a Model Microbial-like Lipid Membrane. Bioelectrochemistry 2021, 141, 107842. [Google Scholar] [CrossRef]
  60. Kumar, P.; Kizhakkedathu, J.N.; Straus, S.K. Antimicrobial Peptides: Diversity, Mechanism of Action and Strategies to Improve the Activity and Biocompatibility In Vivo. Biomolecules 2018, 8, 4. [Google Scholar] [CrossRef]
  61. Wang, Q.; Xu, Y.; Hu, J. Intracellular Mechanism of Antimicrobial Peptide HJH-3 against Salmonella pullorum. RSC Adv. 2022, 12, 14485–14491. [Google Scholar] [CrossRef]
  62. Ruan, Y.; Shen, T.; Wang, Y.; Hou, M.; Li, J.; Sun, T. Antimicrobial Peptide LL-37 Attenuates LTA Induced Inflammatory Effect in Macrophages. Int. Immunopharmacol. 2013, 15, 575–580. [Google Scholar] [CrossRef]
  63. Starr, C.G.; He, J.; Wimley, W.C. Host Cell Interactions Are a Significant Barrier to the Clinical Utility of Peptide Antibiotics. ACS Chem. Biol. 2016, 11, 3391–3399. [Google Scholar] [CrossRef]
  64. Inui Kishi, R.N.; Stach-Machado, D.; Singulani, J. de L.; dos Santos, C.T.; Fusco-Almeida, A.M.; Cilli, E.M.; Freitas-Astúa, J.; Picchi, S.C.; Machado, M.A. Evaluation of Cytotoxicity Features of Antimicrobial Peptides with Potential to Control Bacterial Diseases of Citrus. PLoS ONE 2018, 13, e0203451. [Google Scholar] [CrossRef]
  65. Fazly Bazzaz, B.S.; Seyedi, S.; Hoseini Goki, N.; Khameneh, B. Human Antimicrobial Peptides: Spectrum, Mode of Action and Resistance Mechanisms. Int. J. Pept. Res. Ther. 2021, 27, 801–816. [Google Scholar] [CrossRef]
  66. Brogden, K.A. Antimicrobial Peptides: Pore Formers or Metabolic Inhibitors in Bacteria? Nat. Rev. Microbiol. 2005, 3, 238–250. [Google Scholar] [CrossRef]
  67. Yang, D.; Chen, Q.; Schmidt, A.P.; Anderson, G.M.; Wang, J.; Oppenheim, J.J.; Chertov, O. Ll-37, the Neutrophil Granule–And Epithelial Cell–Derived Cathelicidin, Utilizes Formyl Peptide Receptor–Like 1 (Fprl1) as a Receptor to Chemoattract Human Peripheral Blood Neutrophils, Monocytes, and T Cells. J. Exp. Med. 2000, 192, 1069–1074. [Google Scholar]
  68. Sowers, A.; Wang, G.; Xing, M.; Li, B. Advances in Antimicrobial Peptide Discovery via Machine Learning and Delivery via Nanotechnology. Microorganisms 2023, 11, 1129. [Google Scholar] [CrossRef]
  69. Heymich, M.-L.; Srirangan, S.; Pischetsrieder, M. Stability and Activity of the Antimicrobial Peptide Leg1 in Solution and on Meat and Its Optimized Generation from Chickpea Storage Protein. Foods 2021, 10, 1192. [Google Scholar] [CrossRef]
  70. Chu, H.-L.; Yu, H.-Y.; Yip, B.-S.; Chih, Y.-H.; Liang, C.-W.; Cheng, H.-T.; Cheng, J.-W. Boosting Salt Resistance of Short Antimicrobial Peptides. Antimicrob. Agents Chemother. 2013, 57, 4050–4052. [Google Scholar] [CrossRef]
  71. Nguyen, L.T.; Chau, J.K.; Perry, N.A.; de Boer, L.; Zaat, S.A.J.; Vogel, H.J. Serum Stabilities of Short Tryptophan- and Arginine-Rich Antimicrobial Peptide Analogs. PLoS ONE 2010, 5, e12684. [Google Scholar] [CrossRef]
  72. Chen, C.H.; Lu, T.K. Development and Challenges of Antimicrobial Peptides for Therapeutic Applications. Antibiotics 2020, 9, 24. [Google Scholar] [CrossRef]
  73. Schiffter, H.A. The Delivery of Drugs—Peptides and Proteins. In Comprehensive Biotechnology; Elsevier, 2011; pp. 587–604. ISBN 978-0-08-088504-9. Available online: https://linkinghub.elsevier.com/retrieve/pii/B978008088504900338X (accessed on 7 June 2024).
  74. Gao, Y.; Fang, H.; Fang, L.; Liu, D.; Liu, J.; Su, M.; Fang, Z.; Ren, W.; Jiao, H. The Modification and Design of Antimicrobial Peptide. Curr. Pharm. Des. 2018, 24, 904–910. [Google Scholar] [CrossRef]
  75. Allen, S.D.; Bobbala, S.; Karabin, N.B.; Scott, E.A. On the Advancement of Polymeric Bicontinuous Nanospheres toward Biomedical Applications. Nanoscale Horiz. 2019, 4, 258–272. [Google Scholar] [CrossRef]
  76. Di, L. Strategic Approaches to Optimizing Peptide ADME Properties. AAPS J. 2015, 17, 134–143. [Google Scholar] [CrossRef]
  77. Vlieghe, P.; Lisowski, V.; Martinez, J.; Khrestchatisky, M. Synthetic Therapeutic Peptides: Science and Market. Drug Discov. Today 2010, 15, 40–56. [Google Scholar] [CrossRef]
  78. Fosgerau, K.; Hoffmann, T. Peptide Therapeutics: Current Status and Future Directions. Drug Discov. Today 2015, 20, 122–128. [Google Scholar] [CrossRef]
  79. Larijani, B.; Ozra, T.; Akbar, S.; Eghbal, T.; Mohammad, P.; Mohammad-Hassan, B.; Shahin, A.; Mahmood, M.; Fathemeh, B. Comparison of Desmopressin (Ddavp) Tablet and Intranasal Spray in the Treatment of Central Diabetes Insipidus. DARU J. Pharm. Sci. 2005, 13, 155–159. [Google Scholar]
  80. Mahlapuu, M.; Håkansson, J.; Ringstad, L.; Björn, C. Antimicrobial Peptides: An Emerging Category of Therapeutic Agents. Front. Cell. Infect. Microbiol. 2016, 6, 194. [Google Scholar] [CrossRef]
  81. Hansen, I.K.Ø.; Lövdahl, T.; Simonovic, D.; Hansen, K.Ø.; Andersen, A.J.C.; Devold, H.; Richard, C.S.M.; Andersen, J.H.; Strøm, M.B.; Haug, T. Antimicrobial Activity of Small Synthetic Peptides Based on the Marine Peptide Turgencin A: Prediction of Antimicrobial Peptide Sequences in a Natural Peptide and Strategy for Optimization of Potency. Int. J. Mol. Sci. 2020, 21, 5460. [Google Scholar] [CrossRef]
  82. Ito, T.; Matsunaga, N.; Kurashima, M.; Demizu, Y.; Misawa, T. Enhancing Chemical Stability through Structural Modification of Antimicrobial Peptides with Non-Proteinogenic Amino Acids. Antibiotics 2023, 12, 1326. [Google Scholar] [CrossRef]
  83. Huan, Y.; Kong, Q.; Mou, H.; Yi, H. Antimicrobial Peptides: Classification, Design, Application and Research Progress in Multiple Fields. Front. Microbiol. 2020, 11, 582779. [Google Scholar] [CrossRef]
  84. Reinhardt, A.; Neundorf, I. Design and Application of Antimicrobial Peptide Conjugates. Int. J. Mol. Sci. 2016, 17, 701. [Google Scholar] [CrossRef]
  85. Kamysz, E.; Sikorska, E.; Jaśkiewicz, M.; Bauer, M.; Neubauer, D.; Bartoszewska, S.; Barańska-Rybak, W.; Kamysz, W. Lipidated Analogs of the LL-37-Derived Peptide Fragment KR12—Structural Analysis, Surface-Active Properties and Antimicrobial Activity. Int. J. Mol. Sci. 2020, 21, 887. [Google Scholar] [CrossRef]
  86. Wenzel, M.; Schriek, P.; Prochnow, P.; Albada, H.B.; Metzler-Nolte, N.; Bandow, J.E. Influence of Lipidation on the Mode of Action of a Small RW-Rich Antimicrobial Peptide. Biochim. Biophys. Acta BBA—Biomembr. 2016, 1858, 1004–1011. [Google Scholar] [CrossRef]
  87. Lai, Z.; Chen, H.; Yuan, X.; Tian, J.; Dong, N.; Feng, X.; Shan, A. Designing Double-Site Lipidated Peptide Amphiphiles as Potent Antimicrobial Biomaterials to Combat Multidrug-Resistant Bacteria. Front. Microbiol. 2022, 13, 1074359. [Google Scholar] [CrossRef]
  88. Makowska, M.; Kosikowska-Adamus, P.; Zdrowowicz, M.; Wyrzykowski, D.; Prahl, A.; Sikorska, E. Lipidation of Naturally Occurring α-Helical Antimicrobial Peptides as a Promising Strategy for Drug Design. Int. J. Mol. Sci. 2023, 24, 3951. [Google Scholar] [CrossRef]
  89. Lin, B.; Hung, A.; Singleton, W.; Darmawan, K.K.; Moses, R.; Yao, B.; Wu, H.; Barlow, A.; Sani, M.-A.; Sloan, A.J.; et al. The Effect of Tailing Lipidation on the Bioactivity of Antimicrobial Peptides and Their Aggregation Tendency. Aggregate 2023, 4, e329. [Google Scholar] [CrossRef]
  90. Húmpola, M.V.; Rey, M.C.; Carballeira, N.M.; Simonetta, A.C.; Tonarelli, G.G. Biological and Structural Effects of the Conjugation of an Antimicrobial Decapeptide with Saturated, Unsaturated, Methoxylated and Branched Fatty Acids: Antimicrobial Lipopeptides Containing Unusual Fatty Acids. J. Pept. Sci. 2017, 23, 45–55. [Google Scholar] [CrossRef]
  91. Mahlapuu, M.; Björn, C.; Ekblom, J. Antimicrobial Peptides as Therapeutic Agents: Opportunities and Challenges. Crit. Rev. Biotechnol. 2020, 40, 978–992. [Google Scholar] [CrossRef]
  92. Moretta, A.; Salvia, R.; Scieuzo, C.; Di Somma, A.; Vogel, H.; Pucci, P.; Sgambato, A.; Wolff, M.; Falabella, P. A Bioinformatic Study of Antimicrobial Peptides Identified in the Black Soldier Fly (BSF) Hermetia Illucens (Diptera: Stratiomyidae). Sci. Rep. 2020, 10, 16875. [Google Scholar] [CrossRef]
  93. Jiang, Y.; Chen, Y.; Song, Z.; Tan, Z.; Cheng, J. Recent Advances in Design of Antimicrobial Peptides and Polypeptides toward Clinical Translation. Adv. Drug Deliv. Rev. 2021, 170, 261–280. [Google Scholar] [CrossRef]
  94. Mousavi, S.M.; Babakhani, S.; Moradi, L.; Karami, S.; Shahbandeh, M.; Mirshekar, M.; Mohebi, S.; Moghadam, M.T. Bacteriophage as a Novel Therapeutic Weapon for Killing Colistin-Resistant Multi-Drug-Resistant and Extensively Drug-Resistant Gram-Negative Bacteria. Curr. Microbiol. 2021, 78, 4023–4036. [Google Scholar] [CrossRef]
  95. Binsker, U.; Käsbohrer, A.; Hammerl, J.A. Global Colistin Use: A Review of the Emergence of Resistant Enterobacterales and the Impact on Their Genetic Basis. FEMS Microbiol. Rev. 2021, 46, fuab049. [Google Scholar] [CrossRef]
  96. Olaitan, A.O.; Diene, S.M.; Kempf, M.; Berrazeg, M.; Bakour, S.; Gupta, S.K.; Thongmalayvong, B.; Akkhavong, K.; Somphavong, S.; Paboriboune, P.; et al. Worldwide Emergence of Colistin Resistance in Klebsiella Pneumoniae from Healthy Humans and Patients in Lao PDR, Thailand, Israel, Nigeria and France Owing to Inactivation of the PhoP/PhoQ Regulator mgrB: An Epidemiological and Molecular Study. Int. J. Antimicrob. Agents 2014, 44, 500–507. [Google Scholar] [CrossRef]
  97. Poulikakos, P.; Tansarli, G.S.; Falagas, M.E. Combination Antibiotic Treatment versus Monotherapy for Multidrug-Resistant, Extensively Drug-Resistant, and Pandrug-Resistant Acinetobacter Infections: A Systematic Review. Eur. J. Clin. Microbiol. Infect. Dis. 2014, 33, 1675–1685. [Google Scholar] [CrossRef]
  98. Wang, T.; Liu, H.; Huang, H.; Weng, Y.; Wang, X. Colistin Monotherapy or Combination for the Treatment of Bloodstream Infection Caused by Klebsiella Pneumoniae: A Systematic Review and Meta-Analysis. BMC Infect. Dis. 2024, 24, 161. [Google Scholar] [CrossRef]
  99. Kalin, G.; Alp, E.; Demiraslan, H.; Doganay, M.; Coskun, R.; Gündogan, K. Use of High-Dose IV and Aerosolized Colistin for the Treatment of Multidrug-Resistant Acinetobacter Baumannii Ventilator-Associated Pneumonia: Do We Really Need This Treatment? J. Infect. Chemother. 2012, 18, 872–877. [Google Scholar] [CrossRef]
  100. Tsuji, B.T.; Pogue, J.M.; Zavascki, A.P.; Paul, M.; Daikos, G.L.; Forrest, A.; Giacobbe, D.R.; Viscoli, C.; Giamarellou, H.; Karaiskos, I.; et al. International Consensus Guidelines for the Optimal Use of the Polymyxins: Endorsed by the American College of Clinical Pharmacy (ACCP), European Society of Clinical Microbiology and Infectious Diseases (ESCMID), Infectious Diseases Society of America (IDSA), International Society for Anti-Infective Pharmacology (ISAP), Society of Critical Care Medicine (SCCM), and Society of Infectious Diseases Pharmacists (SIDP). Pharmacotherapy 2019, 39, 10–39. [Google Scholar] [CrossRef]
  101. Falagas, M.E.; Kyriakidou, M.; Voulgaris, G.L.; Vokos, F.; Politi, S.; Kechagias, K.S. Clinical Use of Intravenous Polymyxin B for the Treatment of Patients with Multidrug-Resistant Gram-Negative Bacterial Infections: An Evaluation of the Current Evidence. J. Glob. Antimicrob. Resist. 2021, 24, 342–359. [Google Scholar] [CrossRef]
  102. Rigatto, M.H.; Falci, D.R.; Zavascki, A.P. Clinical Use of Polymyxin B. Adv. Exp. Med. Biol. 2019, 1145, 197–218. [Google Scholar] [CrossRef]
  103. Rychlíčková, J.; Kubíčková, V.; Suk, P.; Urbánek, K. Challenges of Colistin Use in ICU and Therapeutic Drug Monitoring: A Literature Review. Antibiot. Basel Switz. 2023, 12, 437. [Google Scholar] [CrossRef]
  104. El-Sayed Ahmed, M.A.E.-G.; Zhong, L.-L.; Shen, C.; Yang, Y.; Doi, Y.; Tian, G.-B. Colistin and Its Role in the Era of Antibiotic Resistance: An Extended Review (2000-2019). Emerg. Microbes Infect. 2020, 9, 868–885. [Google Scholar] [CrossRef]
  105. Taglialegna, A. Reviving Colistin. Nat. Rev. Microbiol. 2023, 21, 411. [Google Scholar] [CrossRef]
  106. Shin, J.M.; Gwak, J.W.; Kamarajan, P.; Fenno, J.C.; Rickard, A.H.; Kapila, Y.L. Biomedical Applications of Nisin. J. Appl. Microbiol. 2016, 120, 1449–1465. [Google Scholar] [CrossRef]
  107. Soltani, S.; Hammami, R.; Cotter, P.D.; Rebuffat, S.; Said, L.B.; Gaudreau, H.; Bédard, F.; Biron, E.; Drider, D.; Fliss, I. Bacteriocins as a New Generation of Antimicrobials: Toxicity Aspects and Regulations. FEMS Microbiol. Rev. 2021, 45, fuaa039. [Google Scholar] [CrossRef]
  108. Moretta, A.; Scieuzo, C.; Petrone, A.M.; Salvia, R.; Manniello, M.D.; Franco, A.; Lucchetti, D.; Vassallo, A.; Vogel, H.; Sgambato, A.; et al. Antimicrobial Peptides: A New Hope in Biomedical and Pharmaceutical Fields. Front. Cell. Infect. Microbiol. 2021, 11, 668632. [Google Scholar] [CrossRef]
  109. Taylor, S.D.; Palmer, M. The Action Mechanism of Daptomycin. Bioorg. Med. Chem. 2016, 24, 6253–6268. [Google Scholar] [CrossRef]
  110. Wang, S.Z.; Hu, J.T.; Zhang, C.; Zhou, W.; Chen, X.F.; Jiang, L.Y.; Tang, Z.H. The Safety and Efficacy of Daptomycin versus Other Antibiotics for Skin and Soft-Tissue Infections: A Meta-Analysis of Randomised Controlled Trials. BMJ Open 2014, 4, e004744. [Google Scholar] [CrossRef]
  111. David, J.M.; Rajasekaran, A.K. Gramicidin A: A New Mission for an Old Antibiotic. J. Kidney Cancer VHL 2015, 2, 15–24. [Google Scholar] [CrossRef]
  112. Yang, X.; Yousef, A.E. Antimicrobial Peptides Produced by Brevibacillus Spp.: Structure, Classification and Bioactivity: A Mini Review. World J. Microbiol. Biotechnol. 2018, 34, 57. [Google Scholar] [CrossRef]
  113. Douglas, E.J.A.; Wulandari, S.W.; Lovell, S.D.; Laabei, M. Novel Antimicrobial Strategies to Treat Multi-Drug Resistant Staphylococcus Aureus Infections. Microb. Biotechnol. 2023, 16, 1456–1474. [Google Scholar] [CrossRef]
  114. Isaksson, J.; Brandsdal, B.O.; Engqvist, M.; Flaten, G.E.; Svendsen, J.S.M.; Stensen, W. A Synthetic Antimicrobial Peptidomimetic (LTX 109): Stereochemical Impact on Membrane Disruption. J. Med. Chem. 2011, 54, 5786–5795. [Google Scholar] [CrossRef]
  115. Díez-Aguilar, M.; Hernández-García, M.; Morosini, M.-I.; Fluit, A.; Tunney, M.M.; Huertas, N.; Del Campo, R.; Obrecht, D.; Bernardini, F.; Ekkelenkamp, M.; et al. Murepavadin Antimicrobial Activity against and Resistance Development in Cystic Fibrosis Pseudomonas Aeruginosa Isolates. J. Antimicrob. Chemother. 2021, 76, 984–992. [Google Scholar] [CrossRef]
  116. Ekkelenkamp, M.B.; Cantón, R.; Díez-Aguilar, M.; Tunney, M.M.; Gilpin, D.F.; Bernardini, F.; Dale, G.E.; Elborn, J.S.; Bayjanov, J.R.; Fluit, A. Susceptibility of Pseudomonas Aeruginosa Recovered from Cystic Fibrosis Patients to Murepavadin and 13 Comparator Antibiotics. Antimicrob. Agents Chemother. 2020, 64, e01541-19. [Google Scholar] [CrossRef]
  117. Grönberg, A.; Mahlapuu, M.; Ståhle, M.; Whately-Smith, C.; Rollman, O. Treatment with LL-37 Is Safe and Effective in Enhancing Healing of Hard-to-Heal Venous Leg Ulcers: A Randomized, Placebo-Controlled Clinical Trial. Wound Repair Regen. 2014, 22, 613–621. [Google Scholar] [CrossRef]
  118. Sierra, J.M.; Fusté, E.; Rabanal, F.; Vinuesa, T.; Viñas, M. An Overview of Antimicrobial Peptides and the Latest Advances in Their Development. Expert Opin. Biol. Ther. 2017, 17, 663–676. [Google Scholar] [CrossRef]
  119. Koo, H.B.; Seo, J. Antimicrobial Peptides under Clinical Investigation. Pept. Sci. 2019, 111, e24122. [Google Scholar] [CrossRef]
  120. Ge, Y.; MacDonald, D.L.; Holroyd, K.J.; Thornsberry, C.; Wexler, H.; Zasloff, M. In Vitro Antibacterial Properties of Pexiganan, an Analog of Magainin. Antimicrob. Agents Chemother. 1999, 43, 782–788. [Google Scholar] [CrossRef]
  121. Gottler, L.M.; Ramamoorthy, A. Structure, Membrane Orientation, Mechanism, and Function of Pexiganan — A Highly Potent Antimicrobial Peptide Designed from Magainin. Biochim. Biophys. Acta BBA—Biomembr. 2009, 1788, 1680–1686. [Google Scholar] [CrossRef]
  122. Willcox, M.D.P.; Hume, E.B.H.; Aliwarga, Y.; Kumar, N.; Cole, N. A Novel Cationic-peptide Coating for the Prevention of Microbial Colonization on Contact Lenses. J. Appl. Microbiol. 2008, 105, 1817–1825. [Google Scholar] [CrossRef]
  123. Rubinchik, E.; Dugourd, D.; Algara, T.; Pasetka, C.; Friedland, H.D. Antimicrobial and Antifungal Activities of a Novel Cationic Antimicrobial Peptide, Omiganan, in Experimental Skin Colonisation Models. Int. J. Antimicrob. Agents 2009, 34, 457–461. [Google Scholar] [CrossRef]
  124. Sinha, M.; Kaushik, S.; Kaur, P.; Sharma, S.; Singh, T.P. Antimicrobial Lactoferrin Peptides: The Hidden Players in the Protective Function of a Multifunctional Protein. Int. J. Pept. 2013, 2013, 390230. [Google Scholar] [CrossRef]
  125. Mensa, B.; Howell, G.L.; Scott, R.; DeGrado, W.F. Comparative Mechanistic Studies of Brilacidin, Daptomycin, and the Antimicrobial Peptide LL16. Antimicrob. Agents Chemother. 2014, 58, 5136–5145. [Google Scholar] [CrossRef]
  126. Cole, J.N.; Nizet, V. Bacterial Evasion of Host Antimicrobial Peptide Defenses. Microbiol. Spectr. 2016, 4, 4.1.04. [Google Scholar] [CrossRef]
  127. Hooper, N.M. Proteases: A Primer. Essays Biochem. 2002, 38, 1–8. [Google Scholar] [CrossRef]
  128. Guina, T.; Yi, E.C.; Wang, H.; Hackett, M.; Miller, S.I. A PhoP-Regulated Outer Membrane Protease of Salmonella Enterica Serovar Typhimurium Promotes Resistance to Alpha-Helical Antimicrobial Peptides. J. Bacteriol. 2000, 182, 4077–4086. [Google Scholar] [CrossRef]
  129. Sugimura, K.; Nishihara, T. Purification, Characterization, and Primary Structure of Escherichia Coli Protease VII with Specificity for Paired Basic Residues: Identity of Protease VII and OmpT. J. Bacteriol. 1988, 170, 5625–5632. [Google Scholar] [CrossRef]
  130. Sodeinde, O.A.; Subrahmanyam, Y.V.B.K.; Stark, K.; Quan, T.; Bao, Y.; Goguen, J.D. A Surface Protease and the Invasive Character of Plague. Science 1992, 258, 1004–1007. [Google Scholar] [CrossRef]
  131. Stumpe, S.; Bakker, E.P. Requirement of a Large K+-Uptake Capacity and of Extracytoplasmic Protease Activity for Protamine Resistance of Escherichia Coli. Arch. Microbiol. 1997, 167, 126–136. [Google Scholar] [CrossRef]
  132. Boll, J.M.; Tucker, A.T.; Klein, D.R.; Beltran, A.M.; Brodbelt, J.S.; Davies, B.W.; Trent, M.S. Reinforcing Lipid A Acylation on the Cell Surface of Acinetobacter Baumannii Promotes Cationic Antimicrobial Peptide Resistance and Desiccation Survival. mBio 2015, 6, e00478-15. [Google Scholar] [CrossRef]
  133. Kamoshida, G.; Akaji, T.; Takemoto, N.; Suzuki, Y.; Sato, Y.; Kai, D.; Hibino, T.; Yamaguchi, D.; Kikuchi-Ueda, T.; Nishida, S.; et al. Lipopolysaccharide-Deficient Acinetobacter Baumannii Due to Colistin Resistance Is Killed by Neutrophil-Produced Lysozyme. Front. Microbiol. 2020, 11, 573. [Google Scholar] [CrossRef]
  134. Moffatt, J.H.; Harper, M.; Mansell, A.; Crane, B.; Fitzsimons, T.C.; Nation, R.L.; Li, J.; Adler, B.; Boyce, J.D. Lipopolysaccharide-Deficient Acinetobacter Baumannii Shows Altered Signaling through Host Toll-Like Receptors and Increased Susceptibility to the Host Antimicrobial Peptide LL-37. Infect. Immun. 2013, 81, 684–689. [Google Scholar] [CrossRef]
  135. Vila-Farrés, X.; Ferrer-Navarro, M.; Callarisa, A.E.; Martí, S.; Espinal, P.; Gupta, S.; Rolain, J.-M.; Giralt, E.; Vila, J. Loss of LPS Is Involved in the Virulence and Resistance to Colistin of Colistin-Resistant Acinetobacter Nosocomialis Mutants Selected In Vitro. J. Antimicrob. Chemother. 2015, 70, 2981–2986. [Google Scholar] [CrossRef]
  136. Chin, C.-Y.; Gregg, K.A.; Napier, B.A.; Ernst, R.K.; Weiss, D.S. A PmrB-Regulated Deacetylase Required for Lipid A Modification and Polymyxin Resistance in Acinetobacter Baumannii. Antimicrob. Agents Chemother. 2015, 59, 7911–7914. [Google Scholar] [CrossRef]
  137. Bartholomew, T.L.; Kidd, T.J.; Sá Pessoa, J.; Conde Álvarez, R.; Bengoechea, J.A. 2-Hydroxylation of Acinetobacter Baumannii Lipid A Contributes to Virulence. Infect. Immun. 2019, 87, e00066-19. [Google Scholar] [CrossRef]
  138. Srisakul, S.; Wannigama, D.L.; Higgins, P.G.; Hurst, C.; Abe, S.; Hongsing, P.; Saethang, T.; Luk-in, S.; Liao, T.; Kueakulpattana, N.; et al. Overcoming Addition of Phosphoethanolamine to Lipid A Mediated Colistin Resistance in Acinetobacter Baumannii Clinical Isolates with Colistin–Sulbactam Combination Therapy. Sci. Rep. 2022, 12, 11390. [Google Scholar] [CrossRef]
  139. Giles, S.K.; Stroeher, U.H.; Papudeshi, B.; Edwards, R.A.; Carlson-Jones, J.A.; Roach, M.; Brown, M.H. The StkSR Two-Component System Influences Colistin Resistance in Acinetobacter Baumannii. Microorganisms 2022, 10, 985. [Google Scholar] [CrossRef]
  140. Trebosc, V.; Gartenmann, S.; Tötzl, M.; Lucchini, V.; Schellhorn, B.; Pieren, M.; Lociuro, S.; Gitzinger, M.; Tigges, M.; Bumann, D.; et al. Dissecting Colistin Resistance Mechanisms in Extensively Drug-Resistant Acinetobacter Baumannii Clinical Isolates. mBio 2019, 10, e01083-19. [Google Scholar] [CrossRef]
  141. Adams, M.D.; Nickel, G.C.; Bajaksouzian, S.; Lavender, H.; Murthy, A.R.; Jacobs, M.R.; Bonomo, R.A. Resistance to Colistin in Acinetobacter Baumannii Associated with Mutations in the PmrAB Two-Component System. Antimicrob. Agents Chemother. 2009, 53, 3628–3634. [Google Scholar] [CrossRef]
  142. Martins-Sorenson, N.; Snesrud, E.; Xavier, D.E.; Cacci, L.C.; Iavarone, A.T.; McGann, P.; Riley, L.W.; Moreira, B.M. A Novel Plasmid-Encoded Mcr-4.3 Gene in a Colistin-Resistant Acinetobacter Baumannii Clinical Strain. J. Antimicrob. Chemother. 2020, 75, 60–64. [Google Scholar] [CrossRef]
  143. Mills, G.; Dumigan, A.; Kidd, T.; Hobley, L.; Bengoechea, J.A. Identification and Characterization of Two Klebsiella Pneumoniae lpxL Lipid A Late Acyltransferases and Their Role in Virulence. Infect. Immun. 2017, 85, e00068-17. [Google Scholar] [CrossRef]
  144. Tsai, Y.-K.; Liou, C.-H.; Lin, J.-C.; Fung, C.-P.; Chang, F.-Y.; Siu, L.K. Effects of Different Resistance Mechanisms on Antimicrobial Resistance in Acinetobacter Baumannii: A Strategic System for Screening and Activity Testing of New Antibiotics. Int. J. Antimicrob. Agents 2020, 55, 105918. [Google Scholar] [CrossRef]
  145. Uppalapati, S.R.; Sett, A.; Pathania, R. The Outer Membrane Proteins OmpA, CarO, and OprD of Acinetobacter Baumannii Confer a Two-Pronged Defense in Facilitating Its Success as a Potent Human Pathogen. Front. Microbiol. 2020, 11, 589234. [Google Scholar] [CrossRef]
  146. Lin, M.-F.; Lin, Y.-Y.; Lan, C.-Y. Contribution of EmrAB Efflux Pumps to Colistin Resistance in Acinetobacter Baumannii. J. Microbiol. Seoul Korea 2017, 55, 130–136. [Google Scholar] [CrossRef]
  147. Machado, D.; Antunes, J.; Simões, A.; Perdigão, J.; Couto, I.; McCusker, M.; Martins, M.; Portugal, I.; Pacheco, T.; Batista, J.; et al. Contribution of Efflux to Colistin Heteroresistance in a Multidrug Resistant Acinetobacter Baumannii Clinical Isolate. J. Med. Microbiol. 2018, 67, 740–749. [Google Scholar] [CrossRef]
  148. Yoon, E.-J.; Balloy, V.; Fiette, L.; Chignard, M.; Courvalin, P.; Grillot-Courvalin, C. Contribution of the Ade Resistance-Nodulation-Cell Division-Type Efflux Pumps to Fitness and Pathogenesis of Acinetobacter Baumannii. mBio 2016, 7, e00697-16. [Google Scholar] [CrossRef]
  149. Chen, X.; Shao, Z.; Wu, L.; He, B.; Yang, W.; Chen, J.; Jin, E.; Huang, Q.; Lei, L.; Xu, J.; et al. Involvement of the Actinobacillus Pleuropneumoniae ompW Gene in Confrontation of Environmental Pressure. Front. Vet. Sci. 2022, 9, 846322. [Google Scholar] [CrossRef]
  150. Xie, F.; Wang, Y.; Li, G.; Liu, S.; Cui, N.; Liu, S.; Langford, P.R.; Wang, C. The SapA Protein Is Involved in Resistance to Antimicrobial Peptide PR-39 and Virulence of Actinobacillus Pleuropneumoniae. Front. Microbiol. 2017, 8, 811. [Google Scholar] [CrossRef]
  151. Banemann, A.; Deppisch, H.; Gross, R. The Lipopolysaccharide of Bordetella Bronchiseptica Acts as a Protective Shield against Antimicrobial Peptides. Infect. Immun. 1998, 66, 5607–5612. [Google Scholar] [CrossRef]
  152. Spears, P.A.; Temple, L.M.; Orndorff, P.E. A Role for Lipopolysaccharide in Turkey Tracheal Colonization by Bordetella Avium as Demonstrated in Vivo and in Vitro. Mol. Microbiol. 2000, 36, 1425–1435. [Google Scholar] [CrossRef]
  153. Hittle, L.E.; Jones, J.W.; Hajjar, A.M.; Ernst, R.K.; Preston, A. Bordetella Parapertussis PagP Mediates the Addition of Two Palmitates to the Lipopolysaccharide Lipid A. J. Bacteriol. 2015, 197, 572–580. [Google Scholar] [CrossRef]
  154. Geurtsen, J.; Angevaare, E.; Janssen, M.; Hamstra, H.-J.; ten Hove, J.; de Haan, A.; Kuipers, B.; Tommassen, J.; van der Ley, P. A Novel Secondary Acyl Chain in the Lipopolysaccharide of Bordetella Pertussis Required for Efficient Infection of Human Macrophages. J. Biol. Chem. 2007, 282, 37875–37884. [Google Scholar] [CrossRef]
  155. Conde-Álvarez, R.; Palacios-Chaves, L.; Gil-Ramírez, Y.; Salvador-Bescós, M.; Bárcena-Varela, M.; Aragón-Aranda, B.; Martínez-Gómez, E.; Zúñiga-Ripa, A.; De Miguel, M.J.; Bartholomew, T.L.; et al. Identification of lptA, lpxE, and lpxO, Three Genes Involved in the Remodeling of Brucella Cell Envelope. Front. Microbiol. 2018, 8, 2657. [Google Scholar] [CrossRef]
  156. Wang, Z.; Bie, P.; Cheng, J.; Lu, L.; Cui, B.; Wu, Q. The ABC Transporter YejABEF Is Required for Resistance to Antimicrobial Peptides and the Virulence of Brucella Melitensis. Sci. Rep. 2016, 6, 31876. [Google Scholar] [CrossRef]
  157. Kooi, C.; Subsin, B.; Chen, R.; Pohorelic, B.; Sokol, P.A. Burkholderia Cenocepacia ZmpB Is a Broad-Specificity Zinc Metalloprotease Involved in Virulence. Infect. Immun. 2006, 74, 4083–4093. [Google Scholar] [CrossRef] [PubMed]
  158. Kooi, C.; Sokol, P.A. Burkholderia Cenocepacia Zinc Metalloproteases Influence Resistance to Antimicrobial Peptides. Microbiology 2009, 155, 2818–2825. [Google Scholar] [CrossRef] [PubMed]
  159. Foschiatti, M.; Cescutti, P.; Tossi, A.; Rizzo, R. Inhibition of Cathelicidin Activity by Bacterial Exopolysaccharides. Mol. Microbiol. 2009, 72, 1137–1146. [Google Scholar] [CrossRef] [PubMed]
  160. Loutet, S.A.; Flannagan, R.S.; Kooi, C.; Sokol, P.A.; Valvano, M.A. A Complete Lipopolysaccharide Inner Core Oligosaccharide Is Required for Resistance of Burkholderia Cenocepacia to Antimicrobial Peptides and Bacterial Survival in Vivo. J. Bacteriol. 2006, 188, 2073–2080. [Google Scholar] [CrossRef] [PubMed]
  161. Fehlner-Gardiner, C.C.; Valvano, M.A. Cloning and Characterization of the Burkholderia Vietnamiensis norM Gene Encoding a Multi-Drug Efflux Protein. FEMS Microbiol. Lett. 2002, 215, 279–283. [Google Scholar] [CrossRef] [PubMed]
  162. Keo, T.; Collins, J.; Kunwar, P.; Blaser, M.J.; Iovine, N.M. Campylobacter Capsule and Lipooligosaccharide Confer Resistance to Serum and Cationic Antimicrobials. Virulence 2011, 2, 30–40. [Google Scholar] [CrossRef]
  163. Kanipes, M.I.; Holder, L.C.; Corcoran, A.T.; Moran, A.P.; Guerry, P. A Deep-Rough Mutant of Campylobacter Jejuni 81-176 Is Noninvasive for Intestinal Epithelial Cells. Infect. Immun. 2004, 72, 2452–2455. [Google Scholar] [CrossRef]
  164. Lin, J.; Wang, Y.; Van Hoang, K. Systematic Identification of Genetic Loci Required for Polymyxin Resistance in Campylobacter jejuni Using an Efficient In Vivo Transposon Mutagenesis System. Foodborne Pathog. Dis. 2009, 6, 173–185. [Google Scholar] [CrossRef] [PubMed]
  165. Akiba, M.; Lin, J.; Barton, Y.-W.; Zhang, Q. Interaction of CmeABC and CmeDEF in Conferring Antimicrobial Resistance and Maintaining Cell Viability in Campylobacter jejuni. J. Antimicrob. Chemother. 2006, 57, 52–60. [Google Scholar] [CrossRef]
  166. Vieira, A.; Ramesh, A.; Seddon, A.M.; Karlyshev, A.V. CmeABC Multidrug Efflux Pump Contributes to Antibiotic Resistance and Promotes Campylobacter jejuni Survival and Multiplication in Acanthamoeba Polyphaga. Appl. Environ. Microbiol. 2017, 83, e01600-17. [Google Scholar] [CrossRef]
  167. Renzi, F.; Zähringer, U.; Chandler, C.E.; Ernst, R.K.; Cornelis, G.R.; Ittig, S.J. Modification of the 1-Phosphate Group during Biosynthesis of Capnocytophaga Canimorsus Lipid A. Infect. Immun. 2016, 84, 550–561. [Google Scholar] [CrossRef] [PubMed]
  168. Telke, A.A.; Olaitan, A.O.; Morand, S.; Rolain, J.-M. soxRS Induces Colistin Hetero-Resistance in Enterobacter Asburiae and Enterobacter Cloacae by Regulating the acrAB-tolC Efflux Pump. J. Antimicrob. Chemother. 2017, 72, 2715–2721. [Google Scholar] [CrossRef]
  169. Huang, L.; Feng, Y.; Zong, Z. Heterogeneous Resistance to Colistin in Enterobacter Cloacae Complex Due to a New Small Transmembrane Protein. J. Antimicrob. Chemother. 2019, 74, 2551–2558. [Google Scholar] [CrossRef]
  170. Pérez, A.; Poza, M.; Fernández, A.; del Carmen Fernández, M.; Mallo, S.; Merino, M.; Rumbo-Feal, S.; Cabral, M.P.; Bou, G. Involvement of the AcrAB-TolC Efflux Pump in the Resistance, Fitness, and Virulence of Enterobacter Cloacae. Antimicrob. Agents Chemother. 2012, 56, 2084–2090. [Google Scholar] [CrossRef]
  171. Kang, K.N.; Klein, D.R.; Kazi, M.I.; Guérin, F.; Cattoir, V.; Brodbelt, J.S.; Boll, J.M. Colistin Heteroresistance in Enterobacter Cloacae Is Regulated by PhoPQ-Dependent 4-Amino-4-Deoxy-l-Arabinose Addition to Lipid A. Mol. Microbiol. 2019, 111, 1604–1616. [Google Scholar] [CrossRef]
  172. Doijad, S.P.; Gisch, N.; Frantz, R.; Kumbhar, B.V.; Falgenhauer, J.; Imirzalioglu, C.; Falgenhauer, L.; Mischnik, A.; Rupp, J.; Behnke, M.; et al. Resolving Colistin Resistance and Heteroresistance in Enterobacter Species. Nat. Commun. 2023, 14, 140. [Google Scholar] [CrossRef] [PubMed]
  173. Xedzro, C.; Shimamoto, T.; Yu, L.; Zuo, H.; Sugawara, Y.; Sugai, M.; Shimamoto, T. Emergence of Colistin-Resistant Enterobacter Cloacae and Raoultella Ornithinolytica Carrying the Phosphoethanolamine Transferase Gene, Mcr-9, Derived from Vegetables in Japan. Microbiol. Spectr. 2023, 11, e0106323. [Google Scholar] [CrossRef] [PubMed]
  174. López-Solanilla, E.; García-Olmedo, F.; Rodríguez-Palenzuela, P. Inactivation of the sapA to sapF Locus of Erwinia Chrysanthemi Reveals Common Features in Plant and Animal Bacterial Pathogenesis. Plant Cell 1998, 10, 917–924. [Google Scholar] [CrossRef]
  175. Ulvatne, H.; Haukland, H.H.; Samuelsen, Ø.; Krämer, M.; Vorland, L.H. Proteases in Escherichia Coli and Staphylococcus Aureus Confer Reduced Susceptibility to Lactoferricin B. J. Antimicrob. Chemother. 2002, 50, 461–467. [Google Scholar] [CrossRef]
  176. Stumpe, S.; Schmid, R.; Stephens, D.L.; Georgiou, G.; Bakker, E.P. Identification of OmpT as the Protease That Hydrolyzes the Antimicrobial Peptide Protamine before It Enters Growing Cells of Escherichia coli. J. Bacteriol. 1998, 180, 4002–4006. [Google Scholar] [CrossRef]
  177. Thomassin, J.-L.; Brannon, J.R.; Gibbs, B.F.; Gruenheid, S.; Le Moual, H. OmpT Outer Membrane Proteases of Enterohemorrhagic and Enteropathogenic Escherichia coli Contribute Differently to the Degradation of Human LL-37. Infect. Immun. 2012, 80, 483–492. [Google Scholar] [CrossRef] [PubMed]
  178. Desloges, I.; Taylor, J.A.; Leclerc, J.-M.; Brannon, J.R.; Portt, A.; Spencer, J.D.; Dewar, K.; Marczynski, G.T.; Manges, A.; Gruenheid, S.; et al. Identification and Characterization of OmpT-like Proteases in Uropathogenic Escherichia coli Clinical Isolates. MicrobiologyOpen 2019, 8, e915. [Google Scholar] [CrossRef] [PubMed]
  179. Wang, J.; Ma, W.; Fang, Y.; Liang, H.; Yang, H.; Wang, Y.; Dong, X.; Zhan, Y.; Wang, X. Core Oligosaccharide Portion of Lipopolysaccharide Plays Important Roles in Multiple Antibiotic Resistance in Escherichia coli. Antimicrob. Agents Chemother. 2021, 65, e00341-21. [Google Scholar] [CrossRef] [PubMed]
  180. Smith, A.E.; Kim, S.-H.; Liu, F.; Jia, W.; Vinogradov, E.; Gyles, C.L.; Bishop, R.E. PagP Activation in the Outer Membrane Triggers R3 Core Oligosaccharide Truncation in the Cytoplasm of Escherichia coli O157:H7*. J. Biol. Chem. 2008, 283, 4332–4343. [Google Scholar] [CrossRef] [PubMed]
  181. Yang, H.; Fu, J.; Zhao, Y.; Shi, H.; Hu, H.; Wang, H. Escherichia Coli PagP Enzyme-Based De Novo Design and In Vitro Activity of Antibacterial Peptide LL-37. Med. Sci. Monit. Int. Med. J. Exp. Clin. Res. 2017, 23, 2558–2564. [Google Scholar] [CrossRef] [PubMed]
  182. Tran, A.X.; Lester, M.E.; Stead, C.M.; Raetz, C.R.H.; Maskell, D.J.; McGrath, S.C.; Cotter, R.J.; Trent, M.S. Resistance to the Antimicrobial Peptide Polymyxin Requires Myristoylation of Escherichia coli and Salmonella typhimurium Lipid A*. J. Biol. Chem. 2005, 280, 28186–28194. [Google Scholar] [CrossRef] [PubMed]
  183. Chen, H.D.; Groisman, E.A. The PmrA/PmrB Two-Component System: The Major Regulator of LPS Modifications. Annu. Rev. Microbiol. 2013, 67, 83–112. [Google Scholar] [CrossRef] [PubMed]
  184. Lippa, A.M.; Goulian, M. Feedback Inhibition in the PhoQ/PhoP Signaling System by a Membrane Peptide. PLoS Genet. 2009, 5, e1000788. [Google Scholar] [CrossRef] [PubMed]
  185. Gerster, T.; Wröbel, M.; Hofstaedter, C.E.; Schwudke, D.; Ernst, R.K.; Ranf, S.; Gisch, N. Remodeling of Lipid A in Pseudomonas Syringae Pv. Phaseolicola In Vitro. Int. J. Mol. Sci. 2022, 23, 1996. [Google Scholar] [CrossRef]
  186. Salazar, J.; Alarcón, M.; Huerta, J.; Navarro, B.; Aguayo, D. Phosphoethanolamine Addition to the Heptose I of the Lipopolysaccharide Modifies the Inner Core Structure and Has an Impact on the Binding of Polymyxin B to the Escherichia coli Outer Membrane. Arch. Biochem. Biophys. 2017, 620, 28–34. [Google Scholar] [CrossRef]
  187. Bredin, J.; Simonet, V.; Iyer, R.; Delcour, A.H.; Pagès, J.-M. Colicins, Spermine and Cephalosporins: A Competitive Interaction with the OmpF Eyelet. Biochem. J. 2003, 376, 245–252. [Google Scholar] [CrossRef] [PubMed]
  188. Necula, G.; Bacalum, M.; Radu, M. Interaction of Tryptophan- and Arginine-Rich Antimicrobial Peptide with E. Coli Outer Membrane—A Molecular Simulation Approach. Int. J. Mol. Sci. 2023, 24, 2005. [Google Scholar] [CrossRef] [PubMed]
  189. Weatherspoon-Griffin, N.; Zhao, G.; Kong, W.; Kong, Y.; Morigen; Andrews-Polymenis, H.; McClelland, M.; Shi, Y. The CpxR/CpxA Two-Component System Up-Regulates Two Tat-Dependent Peptidoglycan Amidases to Confer Bacterial Resistance to Antimicrobial Peptide *. J. Biol. Chem. 2011, 286, 5529–5539. [Google Scholar] [CrossRef] [PubMed]
  190. Crow, A.; Greene, N.P.; Kaplan, E.; Koronakis, V. Structure and Mechanotransmission Mechanism of the MacB ABC Transporter Superfamily. Proc. Natl. Acad. Sci. USA 2017, 114, 12572–12577. [Google Scholar] [CrossRef] [PubMed]
  191. Weatherspoon-Griffin, N.; Yang, D.; Kong, W.; Hua, Z.; Shi, Y. The CpxR/CpxA Two-Component Regulatory System up-Regulates the Multidrug Resistance Cascade to Facilitate Escherichia coli Resistance to a Model Antimicrobial Peptide. J. Biol. Chem. 2014, 289, 32571–32582. [Google Scholar] [CrossRef] [PubMed]
  192. Chakraborty, K.; Ghosh, S.; Koley, H.; Mukhopadhyay, A.K.; Ramamurthy, T.; Saha, D.R.; Mukhopadhyay, D.; Roychowdhury, S.; Hamabata, T.; Takeda, Y.; et al. Bacterial Exotoxins Downregulate Cathelicidin (hCAP-18/LL-37) and Human Beta-Defensin 1 (HBD-1) Expression in the Intestinal Epithelial Cells. Cell. Microbiol. 2008, 10, 2520–2537. [Google Scholar] [CrossRef] [PubMed]
  193. Wang, X.; McGrath, S.C.; Cotter, R.J.; Raetz, C.R.H. Expression Cloning and Periplasmic Orientation of the Francisella novicida Lipid A 4′-Phosphatase LpxF*. J. Biol. Chem. 2006, 281, 9321–9330. [Google Scholar] [CrossRef] [PubMed]
  194. Kanistanon, D.; Powell, D.A.; Hajjar, A.M.; Pelletier, M.R.; Cohen, I.E.; Way, S.S.; Skerrett, S.J.; Wang, X.; Raetz, C.R.H.; Ernst, R.K. Role of Francisella Lipid A Phosphate Modification in Virulence and Long-Term Protective Immune Responses. Infect. Immun. 2012, 80, 943–951. [Google Scholar] [CrossRef] [PubMed]
  195. Llewellyn, A.C.; Zhao, J.; Song, F.; Parvathareddy, J.; Xu, Q.; Napier, B.A.; Laroui, H.; Merlin, D.; Bina, J.E.; Cotter, P.A.; et al. NaxD Is a Deacetylase Required for Lipid A Modification and F Rancisella Pathogenesis. Mol. Microbiol. 2012, 86, 611–627. [Google Scholar] [CrossRef]
  196. Mount, K.L.B.; Townsend, C.A.; Rinker, S.D.; Gu, X.; Fortney, K.R.; Zwickl, B.W.; Janowicz, D.M.; Spinola, S.M.; Katz, B.P.; Bauer, M.E. Haemophilus Ducreyi SapA Contributes to Cathelicidin Resistance and Virulence in Humans. Infect. Immun. 2010, 78, 1176–1184. [Google Scholar] [CrossRef]
  197. Rinker, S.D.; Gu, X.; Fortney, K.R.; Zwickl, B.W.; Katz, B.P.; Janowicz, D.M.; Spinola, S.M.; Bauer, M.E. Permeases of the Sap Transporter Are Required for Cathelicidin Resistance and Virulence of Haemophilus ducreyi in Humans. J. Infect. Dis. 2012, 206, 1407–1414. [Google Scholar] [CrossRef] [PubMed]
  198. Starner, T.D.; Swords, W.E.; Apicella, M.A.; McCray, P.B. Susceptibility of Nontypeable Haemophilus Influenzae to Human β-Defensins Is Influenced by Lipooligosaccharide Acylation. Infect. Immun. 2002, 70, 5287–5289. [Google Scholar] [CrossRef] [PubMed]
  199. Swords, W.E.; Chance, D.L.; Cohn, L.A.; Shao, J.; Apicella, M.A.; Smith, A.L. Acylation of the Lipooligosaccharide of Haemophilus Influenzae and Colonization: An htrB Mutation Diminishes the Colonization of Human Airway Epithelial Cells. Infect. Immun. 2002, 70, 4661–4668. [Google Scholar] [CrossRef] [PubMed]
  200. Mason, K.M.; Munson, R.S.; Bakaletz, L.O. A Mutation in the Sap Operon Attenuates Survival of Nontypeable Haemophilus Influenzae in a Chinchilla Model of Otitis Media. Infect. Immun. 2005, 73, 599–608. [Google Scholar] [CrossRef] [PubMed]
  201. Cullen, T.W.; Giles, D.K.; Wolf, L.N.; Ecobichon, C.; Boneca, I.G.; Trent, M.S. Helicobacter Pylori versus the Host: Remodeling of the Bacterial Outer Membrane Is Required for Survival in the Gastric Mucosa. PLoS Pathog. 2011, 7, e1002454. [Google Scholar] [CrossRef] [PubMed]
  202. Tran, A.X.; Whittimore, J.D.; Wyrick, P.B.; McGrath, S.C.; Cotter, R.J.; Trent, M.S. The Lipid A 1-Phosphatase of Helicobacter Pylori Is Required for Resistance to the Antimicrobial Peptide Polymyxin. J. Bacteriol. 2006, 188, 4531–4541. [Google Scholar] [CrossRef] [PubMed]
  203. Wang, G.; Lo, L.F.; Forsberg, L.S.; Maier, R.J. Helicobacter Pylori Peptidoglycan Modifications Confer Lysozyme Resistance and Contribute to Survival in the Host. mBio 2012, 3, e00409-12. [Google Scholar] [CrossRef] [PubMed]
  204. Campos, M.A.; Vargas, M.A.; Regueiro, V.; Llompart, C.M.; Albertí, S.; Bengoechea, J.A. Capsule Polysaccharide Mediates Bacterial Resistance to Antimicrobial Peptides. Infect. Immun. 2004, 72, 7107–7114. [Google Scholar] [CrossRef] [PubMed]
  205. Chaput, C.; Spindler, E.; Gill, R.T.; Zychlinsky, A. O-Antigen Protects Gram-Negative Bacteria from Histone Killing. PLoS ONE 2013, 8, e71097. [Google Scholar] [CrossRef] [PubMed]
  206. Clements, A.; Tull, D.; Jenney, A.W.; Farn, J.L.; Kim, S.-H.; Bishop, R.E.; McPhee, J.B.; Hancock, R.E.W.; Hartland, E.L.; Pearse, M.J.; et al. Secondary Acylation of Klebsiella Pneumoniae Lipopolysaccharide Contributes to Sensitivity to Antibacterial Peptides. J. Biol. Chem. 2007, 282, 15569–15577. [Google Scholar] [CrossRef]
  207. Sun, L.; Zhang, Y.; Cai, T.; Li, X.; Li, N.; Xie, Z.; Yang, F.; You, X. CrrAB Regulates PagP-Mediated Glycerophosphoglycerol Palmitoylation in the Outer Membrane of Klebsiella Pneumoniae. J. Lipid Res. 2022, 63, 100251. [Google Scholar] [CrossRef]
  208. Llobet, E.; Martínez-Moliner, V.; Moranta, D.; Dahlström, K.M.; Regueiro, V.; Tomás, A.; Cano, V.; Pérez-Gutiérrez, C.; Frank, C.G.; Fernández-Carrasco, H.; et al. Deciphering Tissue-Induced Klebsiella Pneumoniae Lipid A Structure. Proc. Natl. Acad. Sci. USA 2015, 112, E6369–E6378. [Google Scholar] [CrossRef] [PubMed]
  209. Cannatelli, A.; D’Andrea, M.M.; Giani, T.; Di Pilato, V.; Arena, F.; Ambretti, S.; Gaibani, P.; Rossolini, G.M. In Vivo Emergence of Colistin Resistance in Klebsiella Pneumoniae Producing KPC-Type Carbapenemases Mediated by Insertional Inactivation of the PhoQ/PhoP mgrB Regulator. Antimicrob. Agents Chemother. 2013, 57, 5521–5526. [Google Scholar] [CrossRef] [PubMed]
  210. Mirshekar, M.; Zadeh, R.G.; Moghadam, M.T.; Shahbazi, S.; Masjedian Jazi, F. Upregulation of pmrA, pmrB, pmrC, phoQ, phoP, and arnT Genes Contributing to Resistance to Colistin in Superbug Klebsiella Pneumoniae Isolates from Human Clinical Samples in Tehran, Iran. New Microbes New Infect. 2024, 59, 101275. [Google Scholar] [CrossRef]
  211. Singh, S.; Pathak, A.; Kumar, A.; Rahman, M.; Singh, A.; Gonzalez-Zorn, B.; Prasad, K.N. Emergence of Chromosome-Borne Colistin Resistance Gene Mcr-1 in Clinical Isolates of Klebsiella Pneumoniae from India. Antimicrob. Agents Chemother. 2018, 62, e01885-17. [Google Scholar] [CrossRef] [PubMed]
  212. Llobet, E.; March, C.; Giménez, P.; Bengoechea, J.A. Klebsiella Pneumoniae OmpA Confers Resistance to Antimicrobial Peptides. Antimicrob. Agents Chemother. 2009, 53, 298–302. [Google Scholar] [CrossRef] [PubMed]
  213. Hsu, C.-R.; Chang, I.-W.; Hsieh, P.-F.; Lin, T.-L.; Liu, P.-Y.; Huang, C.-H.; Li, K.-T.; Wang, J.-T. A Novel Role for the Klebsiella Pneumoniae Sap (Sensitivity to Antimicrobial Peptides) Transporter in Intestinal Cell Interactions, Innate Immune Responses, Liver Abscess, and Virulence. J. Infect. Dis. 2019, 219, 1294–1306. [Google Scholar] [CrossRef] [PubMed]
  214. Padilla, E.; Llobet, E.; Doménech-Sánchez, A.; Martínez-Martínez, L.; Bengoechea, J.A.; Albertí, S. Klebsiella Pneumoniae AcrAB Efflux Pump Contributes to Antimicrobial Resistance and Virulence. Antimicrob. Agents Chemother. 2010, 54, 177–183. [Google Scholar] [CrossRef] [PubMed]
  215. Cheng, Y.-H.; Lin, T.-L.; Lin, Y.-T.; Wang, J.-T. A Putative RND-Type Efflux Pump, H239_3064, Contributes to Colistin Resistance through CrrB in Klebsiella Pneumoniae. J. Antimicrob. Chemother. 2018, 73, 1509–1516. [Google Scholar] [CrossRef]
  216. Robey, M.; O’Connell, W.; Cianciotto, N.P. Identification of Legionella Pneumophila Rcp, a pagP-Like Gene That Confers Resistance to Cationic Antimicrobial Peptides and Promotes Intracellular Infection. Infect. Immun. 2001, 69, 4276–4286. [Google Scholar] [CrossRef]
  217. Morgenthau, A.; Livingstone, M.; Adamiak, P.; Schryvers, A.B. The Role of Lactoferrin Binding Protein B in Mediating Protection against Human Lactoferricin. Biochem. Cell Biol. 2012, 90, 417–423. [Google Scholar] [CrossRef] [PubMed]
  218. Morgenthau, A.; Beddek, A.; Schryvers, A.B. The Negatively Charged Regions of Lactoferrin Binding Protein B, an Adaptation against Anti-Microbial Peptides. PloS ONE 2014, 9, e86243. [Google Scholar] [CrossRef] [PubMed]
  219. Hu, L.I.; Stohl, E.A.; Seifert, H.S. The Neisseria Gonorrhoeae Type IV Pilus Promotes Resistance to Hydrogen Peroxide- and LL-37-Mediated Killing by Modulating the Availability of Intracellular, Labile Iron. PLoS Pathog. 2022, 18, e1010561. [Google Scholar] [CrossRef] [PubMed]
  220. Rudel, T.; van Putten, J.P.; Gibbs, C.P.; Haas, R.; Meyer, T.F. Interaction of Two Variable Proteins (PilE and PilC) Required for Pilus-Mediated Adherence of Neisseria Gonorrhoeae to Human Epithelial Cells. Mol. Microbiol. 1992, 6, 3439–3450. [Google Scholar] [CrossRef] [PubMed]
  221. Warner, D.M.; Shafer, W.M.; Jerse, A.E. Clinically Relevant Mutations That Cause Derepression of the Neisseria Gonorrhoeae MtrC-MtrD-MtrE Efflux Pump System Confer Different Levels of Antimicrobial Resistance and in Vivo Fitness. Mol. Microbiol. 2008, 70, 462–478. [Google Scholar] [CrossRef] [PubMed]
  222. Chitsaz, M.; Booth, L.; Blyth, M.T.; O’Mara, M.L.; Brown, M.H. Multidrug Resistance in Neisseria Gonorrhoeae: Identification of Functionally Important Residues in the MtrD Efflux Protein. mBio 2019, 10, e02277-19. [Google Scholar] [CrossRef]
  223. Shafer, W.M.; Qu, X.-D.; Waring, A.J.; Lehrer, R.I. Modulation of Neisseria Gonorrhoeae Susceptibility to Vertebrate Antibacterial Peptides Due to a Member of the Resistance/Nodulation/Division Efflux Pump Family. Proc. Natl. Acad. Sci. USA 1998, 95, 1829–1833. [Google Scholar] [CrossRef] [PubMed]
  224. Handing, J.W.; Ragland, S.A.; Bharathan, U.V.; Criss, A.K. The MtrCDE Efflux Pump Contributes to Survival of Neisseria Gonorrhoeae From Human Neutrophils and Their Antimicrobial Components. Front. Microbiol. 2018, 9, 2688. [Google Scholar] [CrossRef] [PubMed]
  225. Roussel-Jazédé, V.; Jongerius, I.; Bos, M.P.; Tommassen, J.; van Ulsen, P. NalP-Mediated Proteolytic Release of Lactoferrin-Binding Protein B from the Meningococcal Cell Surface. Infect. Immun. 2010, 78, 3083–3089. [Google Scholar] [CrossRef] [PubMed]
  226. Spinosa, M.R.; Progida, C.; Talà, A.; Cogli, L.; Alifano, P.; Bucci, C. The Neisseria Meningitidis Capsule Is Important for Intracellular Survival in Human Cells. Infect. Immun. 2007, 75, 3594–3603. [Google Scholar] [CrossRef]
  227. Jones, A.; Geörg, M.; Maudsdotter, L.; Jonsson, A.-B. Endotoxin, Capsule, and Bacterial Attachment Contribute to Neisseria meningitidis Resistance to the Human Antimicrobial Peptide LL-37. J. Bacteriol. 2009, 191, 3861–3868. [Google Scholar] [CrossRef] [PubMed]
  228. Tzeng, Y.-L.; Stephens, D.S. Antimicrobial Peptide Resistance in Neisseria Meningitidis. Biochim. Biophys. Acta 2015, 1848, 3026–3031. [Google Scholar] [CrossRef] [PubMed]
  229. Tzeng, Y.-L.; Ambrose, K.D.; Zughaier, S.; Zhou, X.; Miller, Y.K.; Shafer, W.M.; Stephens, D.S. Cationic Antimicrobial Peptide Resistance in Neisseria Meningitidis. J. Bacteriol. 2005, 187, 5387–5396. [Google Scholar] [CrossRef]
  230. Piek, S.; Wang, Z.; Ganguly, J.; Lakey, A.M.; Bartley, S.N.; Mowlaboccus, S.; Anandan, A.; Stubbs, K.A.; Scanlon, M.J.; Vrielink, A.; et al. The Role of Oxidoreductases in Determining the Function of the Neisserial Lipid A Phosphoethanolamine Transferase Required for Resistance to Polymyxin. PLoS ONE 2014, 9, e106513. [Google Scholar] [CrossRef]
  231. Abi Khattar, Z.; Lanois, A.; Hadchity, L.; Gaudriault, S.; Givaudan, A. Spatiotemporal Expression of the Putative MdtABC Efflux Pump of Phtotorhabdus Luminescens Occurs in a Protease-Dependent Manner during Insect Infection. PLoS ONE 2019, 14, e0212077. [Google Scholar] [CrossRef] [PubMed]
  232. Hadchity, L.; Lanois, A.; Kiwan, P.; Nassar, F.; Givaudan, A.; Khattar, Z.A. AcrAB, the Major RND-Type Efflux Pump of Photorhabdus Laumondii, Confers Intrinsic Multidrug-Resistance and Contributes to Virulence in Insects. Environ. Microbiol. Rep. 2021, 13, 637–648. [Google Scholar] [CrossRef] [PubMed]
  233. Derzelle, S.; Turlin, E.; Duchaud, E.; Pages, S.; Kunst, F.; Givaudan, A.; Danchin, A. The PhoP-PhoQ Two-Component Regulatory System of Photorhabdus Luminescens Is Essential for Virulence in Insects. J. Bacteriol. 2004, 186, 1270–1279. [Google Scholar] [CrossRef] [PubMed]
  234. Bennett, H.P.J.; Clarke, D.J. The pbgPE Operon in Photorhabdus Luminescens Is Required for Pathogenicity and Symbiosis. J. Bacteriol. 2005, 187, 77–84. [Google Scholar] [CrossRef] [PubMed]
  235. Coats, S.R.; Jones, J.W.; Do, C.T.; Braham, P.H.; Bainbridge, B.W.; To, T.T.; Goodlett, D.R.; Ernst, R.K.; Darveau, R.P. Human Toll-like Receptor 4 Responses to P. Gingivalis Are Regulated by Lipid A 1- and 4′-Phosphatase Activities. Cell. Microbiol. 2009, 11, 1587–1599. [Google Scholar] [CrossRef]
  236. Horie, T.; Inomata, M.; Into, T. OmpA-Like Proteins of Porphyromonas Gingivalis Mediate Resistance to the Antimicrobial Peptide LL-37. J. Pathog. 2018, 2018, 2068435. [Google Scholar] [CrossRef]
  237. Devine, D.A.; Marsh, P.D.; Percival, R.S.; Rangarajan, M.; Curtis, M.A. Modulation of Antibacterial Peptide Activity by Products of Porphyromonas Gingivalis and Prevotella Spp. Microbiology 1999, 145, 965–971. [Google Scholar] [CrossRef]
  238. Belas, R.; Manos, J.; Suvanasuthi, R. Proteus Mirabilis ZapA Metalloprotease Degrades a Broad Spectrum of Substrates, Including Antimicrobial Peptides. Infect. Immun. 2004, 72, 5159–5167. [Google Scholar] [CrossRef]
  239. McCoy, A.J.; Liu, H.; Falla, T.J.; Gunn, J.S. Identification of Proteus Mirabilis Mutants with Increased Sensitivity to Antimicrobial Peptides. Antimicrob. Agents Chemother. 2001, 45, 2030–2037. [Google Scholar] [CrossRef] [PubMed]
  240. Schmidtchen, A.; Frick, I.-M.; Andersson, E.; Tapper, H.; Björck, L. Proteinases of Common Pathogenic Bacteria Degrade and Inactivate the Antibacterial Peptide LL-37. Mol. Microbiol. 2002, 46, 157–168. [Google Scholar] [CrossRef] [PubMed]
  241. Llobet, E.; Tomás, J.M.; Bengoechea, J.A. Capsule Polysaccharide Is a Bacterial Decoy for Antimicrobial Peptides. Microbiol. Read. Engl. 2008, 154, 3877–3886. [Google Scholar] [CrossRef] [PubMed]
  242. Park, P.W.; Pier, G.B.; Hinkes, M.T.; Bernfield, M. Exploitation of Syndecan-1 Shedding by Pseudomonas Aeruginosa Enhances Virulence. Nature 2001, 411, 98–102. [Google Scholar] [CrossRef]
  243. Park, P.W.; Pier, G.B.; Preston, M.J.; Goldberger, O.; Fitzgerald, M.L.; Bernfield, M. Syndecan-1 Shedding Is Enhanced by LasA, a Secreted Virulence Factor of Pseudomonas Aeruginosa*. J. Biol. Chem. 2000, 275, 3057–3064. [Google Scholar] [CrossRef]
  244. Lo Sciuto, A.; Cervoni, M.; Stefanelli, R.; Spinnato, M.C.; Di Giamberardino, A.; Mancone, C.; Imperi, F. Genetic Basis and Physiological Effects of Lipid A Hydroxylation in Pseudomonas Aeruginosa PAO1. Pathogens 2019, 8, 291. [Google Scholar] [CrossRef]
  245. Hofstaedter, C.E.; Chandler, C.E.; Met, C.M.; Gillespie, J.J.; Harro, J.M.; Goodlett, D.R.; Rasko, D.A.; Ernst, R.K. Divergent Pseudomonas Aeruginosa LpxO Enzymes Perform Site-Specific Lipid A 2-Hydroxylation. mBio 2023, 15, e02823-23. [Google Scholar] [CrossRef]
  246. Poirel, L.; Jayol, A.; Nordmann, P. Polymyxins: Antibacterial Activity, Susceptibility Testing, and Resistance Mechanisms Encoded by Plasmids or Chromosomes. Clin. Microbiol. Rev. 2017, 30, 557–596. [Google Scholar] [CrossRef]
  247. Ernst, R.K.; Yi, E.C.; Guo, L.; Lim, K.B.; Burns, J.L.; Hackett, M.; Miller, S.I. Specific Lipopolysaccharide Found in Cystic Fibrosis Airway Pseudomonas Aeruginosa. Science 1999, 286, 1561–1565. [Google Scholar] [CrossRef] [PubMed]
  248. Moskowitz, S.M.; Ernst, R.K.; Miller, S.I. PmrAB, a Two-Component Regulatory System of Pseudomonas Aeruginosa That Modulates Resistance to Cationic Antimicrobial Peptides and Addition of Aminoarabinose to Lipid A. J. Bacteriol. 2004, 186, 575–579. [Google Scholar] [CrossRef] [PubMed]
  249. Snesrud, E.; Maybank, R.; Kwak, Y.I.; Jones, A.R.; Hinkle, M.K.; McGann, P. Chromosomally Encoded Mcr-5 in Colistin-Nonsusceptible Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 2018, 62, e00679-18. [Google Scholar] [CrossRef] [PubMed]
  250. Cervoni, M.; Sposato, D.; Lo Sciuto, A.; Imperi, F. Regulatory Landscape of the Pseudomonas Aeruginosa Phosphoethanolamine Transferase Gene eptA in the Context of Colistin Resistance. Antibiotics 2023, 12, 200. [Google Scholar] [CrossRef] [PubMed]
  251. Cervoni, M.; Lo Sciuto, A.; Bianchini, C.; Mancone, C.; Imperi, F. Exogenous and Endogenous Phosphoethanolamine Transferases Differently Affect Colistin Resistance and Fitness in Pseudomonas Aeruginosa. Front. Microbiol. 2021, 12, 778968. [Google Scholar] [CrossRef] [PubMed]
  252. Klein, S.; Lorenzo, C.; Hoffmann, S.; Walther, J.M.; Storbeck, S.; Piekarski, T.; Tindall, B.J.; Wray, V.; Nimtz, M.; Moser, J. Adaptation of Pseudomonas Aeruginosa to Various Conditions Includes tRNA-Dependent Formation of Alanyl-Phosphatidylglycerol. Mol. Microbiol. 2009, 71, 551–565. [Google Scholar] [CrossRef] [PubMed]
  253. Roy, H.; Dare, K.; Ibba, M. Adaptation of the Bacterial Membrane to Changing Environments Using Aminoacylated Phospholipids. Mol. Microbiol. 2009, 71, 547–550. [Google Scholar] [CrossRef]
  254. Pamp, S.J.; Gjermansen, M.; Johansen, H.K.; Tolker-Nielsen, T. Tolerance to the Antimicrobial Peptide Colistin in Pseudomonas Aeruginosa Biofilms Is Linked to Metabolically Active Cells, and Depends on the Pmr and mexAB-oprM Genes. Mol. Microbiol. 2008, 68, 223–240. [Google Scholar] [CrossRef] [PubMed]
  255. Masuda, N.; Sakagawa, E.; Ohya, S.; Gotoh, N.; Tsujimoto, H.; Nishino, T. Substrate Specificities of MexAB-OprM, MexCD-OprJ, and MexXY-OprM Efflux Pumps in Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 2000, 44, 3322–3327. [Google Scholar] [CrossRef]
  256. Fernandes, S.E.; Jorth, P. A Brief Update on the Controversial and Opposing Roles of Pseudomonas Aeruginosa Efflux Pumps in Virulence Regulation. Front. Bacteriol. 2023, 2, 1231657. [Google Scholar] [CrossRef]
  257. Taggart, C.C.; Greene, C.M.; Smith, S.G.; Levine, R.L.; McCray, P.B., Jr.; O’Neill, S.; McElvaney, N.G. Inactivation of Human β-Defensins 2 and 3 by Elastolytic Cathepsins1. J. Immunol. 2003, 171, 931–937. [Google Scholar] [CrossRef] [PubMed]
  258. Andrault, P.-M.; Samsonov, S.A.; Weber, G.; Coquet, L.; Nazmi, K.; Bolscher, J.G.M.; Lalmanach, A.-C.; Jouenne, T.; Brömme, D.; Pisabarro, M.T.; et al. Antimicrobial Peptide LL-37 Is Both a Substrate of Cathepsins S and K and a Selective Inhibitor of Cathepsin L. Biochemistry 2015, 54, 2785–2798. [Google Scholar] [CrossRef] [PubMed]
  259. Dorrer, E.; Teuber, M. Induction of Polymyxin Resistance in Pseudomonas Fluorescens by Phosphate Limitation. Arch. Microbiol. 1977, 114, 87–89. [Google Scholar] [CrossRef] [PubMed]
  260. Guo, L.; Lim, K.B.; Gunn, J.S.; Bainbridge, B.; Darveau, R.P.; Hackett, M.; Miller, S.I. Regulation of Lipid A Modifications by Salmonella Typhimurium Virulence Genes phoP-phoQ. Science 1997, 276, 250–253. [Google Scholar] [CrossRef] [PubMed]
  261. Kawasaki, K.; China, K.; Nishijima, M. Release of the Lipopolysaccharide Deacylase PagL from Latency Compensates for a Lack of Lipopolysaccharide Aminoarabinose Modification-Dependent Resistance to the Antimicrobial Peptide Polymyxin B in Salmonella Enterica. J. Bacteriol. 2007, 189, 4911–4919. [Google Scholar] [CrossRef] [PubMed]
  262. Strandberg, K.L.; Richards, S.M.; Gunn, J.S. Cathelicidin Antimicrobial Peptide Expression Is Not Induced or Required for Bacterial Clearance during Salmonella Enterica Infection of Human Monocyte-Derived Macrophages. Infect. Immun. 2012, 80, 3930–3938. [Google Scholar] [CrossRef]
  263. Otton, L.M.; da Silva Campos, M.; Meneghetti, K.L.; Corção, G. Influence of Twitching and Swarming Motilities on Biofilm Formation in Pseudomonas Strains. Arch. Microbiol. 2017, 199, 677–682. [Google Scholar] [CrossRef] [PubMed]
  264. Tamayo, R.; Choudhury, B.; Septer, A.; Merighi, M.; Carlson, R.; Gunn, J.S. Identification of cptA, a PmrA-Regulated Locus Required for Phosphoethanolamine Modification of the Salmonella Enterica Serovar Typhimurium Lipopolysaccharide Core. J. Bacteriol. 2005, 187, 3391–3399. [Google Scholar] [CrossRef]
  265. Bogomolnaya, L.M.; Andrews, K.D.; Talamantes, M.; Maple, A.; Ragoza, Y.; Vazquez-Torres, A.; Andrews-Polymenis, H. The ABC-Type Efflux Pump MacAB Protects Salmonella Enterica Serovar Typhimurium from Oxidative Stress. mBio 2013, 4, e00630-13. [Google Scholar] [CrossRef]
  266. Honeycutt, J.D.; Wenner, N.; Li, Y.; Brewer, S.M.; Massis, L.M.; Brubaker, S.W.; Chairatana, P.; Owen, S.V.; Canals, R.; Hinton, J.C.D.; et al. Genetic Variation in the MacAB-TolC Efflux Pump Influences Pathogenesis of Invasive Salmonella Isolates from Africa. PLOS Pathog. 2020, 16, e1008763. [Google Scholar] [CrossRef]
  267. Parra-Lopez, C.; Baer, M.T.; Groisman, E.A. Molecular Genetic Analysis of a Locus Required for Resistance to Antimicrobial Peptides in Salmonella Typhimurium. EMBO J. 1993, 12, 4053–4062. [Google Scholar] [CrossRef] [PubMed]
  268. Eswarappa, S.M.; Panguluri, K.K.; Hensel, M.; Chakravortty, D. The yejABEF Operon of Salmonella Confers Resistance to Antimicrobial Peptides and Contributes to Its Virulence. Microbiology 2008, 154, 666–678. [Google Scholar] [CrossRef] [PubMed]
  269. Crouch, M.-L.; Becker, L.A.; Bang, I.-S.; Tanabe, H.; Ouellette, A.J.; Fang, F.C. The Alternative Sigma Factor Sigma Is Required for Resistance of Salmonella Enterica Serovar Typhimurium to Anti-Microbial Peptides. Mol. Microbiol. 2005, 56, 789–799. [Google Scholar] [CrossRef] [PubMed]
  270. Humphreys, S.; Stevenson, A.; Bacon, A.; Weinhardt, A.B.; Roberts, M. The Alternative Sigma Factor, sigmaE, Is Critically Important for the Virulence of Salmonella Typhimurium. Infect. Immun. 1999, 67, 1560–1568. [Google Scholar] [CrossRef] [PubMed]
  271. Islam, D.; Bandholtz, L.; Nilsson, J.; Wigzell, H.; Christensson, B.; Agerberth, B.; Gudmundsson, G. Downregulation of Bactericidal Peptides in Enteric Infections: A Novel Immune Escape Mechanism with Bacterial DNA as a Potential Regulator. Nat. Med. 2001, 7, 180–185. [Google Scholar] [CrossRef]
  272. West, N.P.; Sansonetti, P.; Mounier, J.; Exley, R.M.; Parsot, C.; Guadagnini, S.; Prévost, M.-C.; Prochnicka-Chalufour, A.; Delepierre, M.; Tanguy, M.; et al. Optimization of Virulence Functions Through Glucosylation of Shigella LPS. Science 2005, 307, 1313–1317. [Google Scholar] [CrossRef]
  273. Arbibe, L.; Kim, D.W.; Batsche, E.; Pedron, T.; Mateescu, B.; Muchardt, C.; Parsot, C.; Sansonetti, P.J. An Injected Bacterial Effector Targets Chromatin Access for Transcription Factor NF-kappaB to Alter Transcription of Host Genes Involved in Immune Responses. Nat. Immunol. 2007, 8, 47–56. [Google Scholar] [CrossRef]
  274. Xiao, L.; Crabb, D.M.; Dai, Y.; Chen, Y.; Waites, K.B.; Atkinson, T.P. Suppression of Antimicrobial Peptide Expression by Ureaplasma Species. Infect. Immun. 2014, 82, 1657–1665. [Google Scholar] [CrossRef]
  275. Hankins, J.V.; Madsen, J.A.; Giles, D.K.; Childers, B.M.; Klose, K.E.; Brodbelt, J.S.; Trent, M.S. Elucidation of a Novel Vibrio Cholerae Lipid A Secondary Hydroxy-Acyltransferase and Its Role in Innate Immune Recognition. Mol. Microbiol. 2011, 81, 1313–1329. [Google Scholar] [CrossRef]
  276. Mathur, J.; Waldor, M.K. The Vibrio Cholerae ToxR-Regulated Porin OmpU Confers Resistance to Antimicrobial Peptides. Infect. Immun. 2004, 72, 3577–3583. [Google Scholar] [CrossRef]
  277. Mathur, J.; Davis, B.M.; Waldor, M.K. Antimicrobial Peptides Activate the Vibrio Cholerae sigmaE Regulon through an OmpU-Dependent Signalling Pathway. Mol. Microbiol. 2007, 63, 848–858. [Google Scholar] [CrossRef] [PubMed]
  278. Bina, X.R.; Provenzano, D.; Nguyen, N.; Bina, J.E. Vibrio Cholerae RND Family Efflux Systems Are Required for Antimicrobial Resistance, Optimal Virulence Factor Production, and Colonization of the Infant Mouse Small Intestine. Infect. Immun. 2008, 76, 3595–3605. [Google Scholar] [CrossRef] [PubMed]
  279. Chen, Y.-C.; Chuang, Y.-C.; Chang, C.-C.; Jeang, C.-L.; Chang, M.-C. A K+ Yptake Protein, TrkA, Is Required for Serum, Protamine, and Polymyxin B Resistance in Vibrio Vulnificus. Infect. Immun. 2004, 72, 629–636. [Google Scholar] [CrossRef] [PubMed]
  280. Pérez-Gutiérrez, C.; Llobet, E.; Llompart, C.M.; Reinés, M.; Bengoechea, J.A. Role of Lipid A Acylation in Yersinia enterocolitica Virulence. Infect. Immun. 2010, 78, 2768–2781. [Google Scholar] [CrossRef] [PubMed]
  281. Visser, L.G.; Hiemstra, P.S.; van den Barselaar, M.T.; Ballieux, P.A.; van Furth, R. Role of YadA in Resistance to Killing of Yersinia enterocolitica by Antimicrobial Polypeptides of Human Granulocytes. Infect. Immun. 1996, 64, 1653–1658. [Google Scholar] [CrossRef] [PubMed]
  282. Bengoechea, J.A.; Skurnik, M. Temperature-Regulated Efflux Pump/Potassium Antiporter System Mediates Resistance to Cationic Antimicrobial Peptides in Yersinia. Mol. Microbiol. 2000, 37, 67–80. [Google Scholar] [CrossRef] [PubMed]
  283. Lister, I.M.; Raftery, C.; Mecsas, J.; Levy, S.B. Yersinia Pestis acrAB-tolC in Antibiotic Resistance and Virulence. Antimicrob. Agents Chemother. 2012, 56, 1120–1123. [Google Scholar] [CrossRef] [PubMed]
  284. Rebeil, R.; Ernst, R.K.; Gowen, B.B.; Miller, S.I.; Hinnebusch, B.J. Variation in Lipid A Structure in the Pathogenic Yersiniae. Mol. Microbiol. 2004, 52, 1363–1373. [Google Scholar] [CrossRef]
  285. Thwaite, J.E.; Hibbs, S.; Titball, R.W.; Atkins, T.P. Proteolytic Degradation of Human Antimicrobial Peptide LL-37 by Bacillus Anthracis May Contribute to Virulence. Antimicrob. Agents Chemother. 2006, 50, 2316–2322. [Google Scholar] [CrossRef] [PubMed]
  286. McGillivray, S.M.; Ebrahimi, C.M.; Fisher, N.; Sabet, M.; Zhang, D.X.; Chen, Y.; Haste, N.M.; Aroian, R.V.; Gallo, R.L.; Guiney, D.G.; et al. ClpX Contributes to Innate Defense Peptide Resistance and Virulence Phenotypes of Bacillus Anthracis. J. Innate Immun. 2009, 1, 494–506. [Google Scholar] [CrossRef]
  287. Drysdale, M.; Bourgogne, A.; Hilsenbeck, S.G.; Koehler, T.M. atxA Controls Bacillus Anthracis Capsule Synthesis via acpA and a Newly Discovered Regulator, acpB. J. Bacteriol. 2004, 186, 307–315. [Google Scholar] [CrossRef] [PubMed]
  288. Richter, S.; Anderson, V.J.; Garufi, G.; Lu, L.; Budzik, J.M.; Joachimiak, A.; He, C.; Schneewind, O.; Missiakas, D. Capsule Anchoring in Bacillus Anthracis Occurs by a Transpeptidation Reaction That Is Inhibited by Capsidin. Mol. Microbiol. 2009, 71, 404–420. [Google Scholar] [CrossRef] [PubMed]
  289. Fisher, N.; Shetron-Rama, L.; Herring-Palmer, A.; Heffernan, B.; Bergman, N.; Hanna, P. The dltABCD Operon of Bacillus Anthracis Sterne Is Required for Virulence and Resistance to Peptide, Enzymatic, and Cellular Mediators of Innate Immunity. J. Bacteriol. 2006, 188, 1301–1309. [Google Scholar] [CrossRef] [PubMed]
  290. Samant, S.; Hsu, F.-F.; Neyfakh, A.A.; Lee, H. The Bacillus Anthracis Protein MprF Is Required for Synthesis of Lysylphosphatidylglycerols and for Resistance to Cationic Antimicrobial Peptides. J. Bacteriol. 2009, 191, 1311–1319. [Google Scholar] [CrossRef] [PubMed]
  291. Abi Khattar, Z.; Rejasse, A.; Destoumieux-Garzón, D.; Escoubas, J.M.; Sanchis, V.; Lereclus, D.; Givaudan, A.; Kallassy, M.; Nielsen-Leroux, C.; Gaudriault, S. The Dlt Operon of Bacillus Cereus Is Required for Resistance to Cationic Antimicrobial Peptides and for Virulence in Insects. J. Bacteriol. 2009, 191, 7063–7073. [Google Scholar] [CrossRef] [PubMed]
  292. Haydar, A.; Tran, S.-L.; Guillemet, E.; Darrigo, C.; Perchat, S.; Lereclus, D.; Coquet, L.; Jouenne, T.; Ramarao, N. InhA1-Mediated Cleavage of the Metalloprotease NprA Allows Bacillus Cereus to Escape From Macrophages. Front. Microbiol. 2018, 9, 1063. [Google Scholar] [CrossRef] [PubMed]
  293. Guillemet, E.; Cadot, C.; Tran, S.-L.; Guinebretière, M.-H.; Lereclus, D.; Ramarao, N. The InhA Metalloproteases of Bacillus Cereus Contribute Concomitantly to Virulence. J. Bacteriol. 2010, 192, 286–294. [Google Scholar] [CrossRef] [PubMed]
  294. Cao, M.; Helmann, J.D. The Bacillus Subtilis Extracytoplasmic-Function σX Factor Regulates Modification of the Cell Envelope and Resistance to Cationic Antimicrobial Peptides. J. Bacteriol. 2004, 186, 1136–1146. [Google Scholar] [CrossRef]
  295. Perego, M.; Glaser, P.; Minutello, A.; Strauch, M.A.; Leopold, K.; Fischer, W. Incorporation of D-Alanine into Lipoteichoic Acid and Wall Teichoic Acid in Bacillus Subtilis. Identification of Genes and Regulation. J. Biol. Chem. 1995, 270, 15598–15606. [Google Scholar] [CrossRef]
  296. McBride, S.M.; Sonenshein, A.L. The Dlt Operon Confers Resistance to Cationic Antimicrobial Peptides in Clostridium Difficile. Microbiology 2011, 157, 1457–1465. [Google Scholar] [CrossRef]
  297. Singh, K.V.; Coque, T.M.; Weinstock, G.M.; Murray, B.E. In Vivo Testing of an Enterococcus Faecalis efaA Mutant and Use of efaA Homologs for Species Identification. FEMS Immunol. Med. Microbiol. 1998, 21, 323–331. [Google Scholar] [CrossRef] [PubMed]
  298. Nešuta, O.; Buděšínský, M.; Hadravová, R.; Monincová, L.; Humpoličková, J.; Čeřovský, V. How Proteases from Enterococcus Faecalis Contribute to Its Resistance to Short α-Helical Antimicrobial Peptides. Pathog. Dis. 2017, 75, ftx091. [Google Scholar] [CrossRef] [PubMed]
  299. Engelbert, M.; Mylonakis, E.; Ausubel, F.M.; Calderwood, S.B.; Gilmore, M.S. Contribution of Gelatinase, Serine Protease, and Fsr to the Pathogenesis of Enterococcus Faecalis Endophthalmitis. Infect. Immun. 2004, 72, 3628–3633. [Google Scholar] [CrossRef] [PubMed]
  300. Schmidtchen, A.; Frick, I.-M.; Björck, L. Dermatan Sulphate Is Released by Proteinases of Common Pathogenic Bacteria and Inactivates Antibacterial α-Defensin. Mol. Microbiol. 2001, 39, 708–713. [Google Scholar] [CrossRef] [PubMed]
  301. Fabretti, F.; Theilacker, C.; Baldassarri, L.; Kaczynski, Z.; Kropec, A.; Holst, O.; Huebner, J. Alanine Esters of Enterococcal Lipoteichoic Acid Play a Role in Biofilm Formation and Resistance to Antimicrobial Peptides. Infect. Immun. 2006, 74, 4164–4171. [Google Scholar] [CrossRef] [PubMed]
  302. Bao, Y.; Sakinc, T.; Laverde, D.; Wobser, D.; Benachour, A.; Theilacker, C.; Hartke, A.; Huebner, J. Role of mprF1 and mprF2 in the Pathogenicity of Enterococcus Faecalis. PLoS ONE 2012, 7, e38458. [Google Scholar] [CrossRef]
  303. Kandaswamy, K.; Liew, T.H.; Wang, C.Y.; Huston-Warren, E.; Meyer-Hoffert, U.; Hultenby, K.; Schröder, J.M.; Caparon, M.G.; Normark, S.; Henriques-Normark, B.; et al. Focal Targeting by Human β-Defensin 2 Disrupts Localized Virulence Factor Assembly Sites in Enterococcus Faecalis. Proc. Natl. Acad. Sci. USA 2013, 110, 20230–20235. [Google Scholar] [CrossRef] [PubMed]
  304. Nair, Z.J.; Gao, I.H.; Firras, A.; Chong, K.K.L.; Hill, E.D.; Choo, P.Y.; Colomer-Winter, C.; Chen, Q.; Manzano, C.; Pethe, K.; et al. An Essential Protease, FtsH, Influences Daptomycin Resistance Acquisition in Enterococcus faecalis. Mol. Microbiol. 2024, 121, 1021–1038. [Google Scholar] [CrossRef] [PubMed]
  305. Reyes, J.; Panesso, D.; Tran, T.T.; Mishra, N.N.; Cruz, M.R.; Munita, J.M.; Singh, K.V.; Yeaman, M.R.; Murray, B.E.; Shamoo, Y.; et al. A liaR Deletion Restores Susceptibility to Daptomycin and Antimicrobial Peptides in Multidrug-Resistant Enterococcus Faecalis. J. Infect. Dis. 2015, 211, 1317–1325. [Google Scholar] [CrossRef]
  306. Hébert, L.; Courtin, P.; Torelli, R.; Sanguinetti, M.; Chapot-Chartier, M.-P.; Auffray, Y.; Benachour, A. Enterococcus Faecalis Constitutes an Unusual Bacterial Model in Lysozyme Resistance. Infect. Immun. 2007, 75, 5390–5398. [Google Scholar] [CrossRef]
  307. Johansson, L.; Thulin, P.; Sendi, P.; Hertzén, E.; Linder, A.; Åkesson, P.; Low, D.E.; Agerberth, B.; Norrby-Teglund, A. Cathelicidin LL-37 in Severe Streptococcus Pyogenes Soft Tissue Infections in Humans. Infect. Immun. 2008, 76, 3399–3404. [Google Scholar] [CrossRef] [PubMed]
  308. Cole, J.N.; Pence, M.A.; von Köckritz-Blickwede, M.; Hollands, A.; Gallo, R.L.; Walker, M.J.; Nizet, V. M Protein and Hyaluronic Acid Capsule Are Essential for in Vivo Selection of covRS Mutations Characteristic of Invasive Serotype M1T1 Group A Streptococcus. mBio 2010, 1, e00191-10. [Google Scholar] [CrossRef] [PubMed]
  309. Lauth, X.; von Köckritz-Blickwede, M.; McNamara, C.W.; Myskowski, S.; Zinkernagel, A.S.; Beall, B.; Ghosh, P.; Gallo, R.L.; Nizet, V. M1 Protein Allows Group A Streptococcal Survival in Phagocyte Extracellular Traps through Cathelicidin Inhibition. J. Innate Immun. 2009, 1, 202–214. [Google Scholar] [CrossRef] [PubMed]
  310. Hollands, A.; Gonzalez, D.; Leire, E.; Donald, C.; Gallo, R.L.; Sanderson-Smith, M.; Dorrestein, P.C.; Nizet, V. A Bacterial Pathogen Co-Opts Host Plasmin to Resist Killing by Cathelicidin Antimicrobial Peptides. J. Biol. Chem. 2012, 287, 40891–40897. [Google Scholar] [CrossRef] [PubMed]
  311. Sun, H.; Xu, Y.; Sitkiewicz, I.; Ma, Y.; Wang, X.; Yestrepsky, B.D.; Huang, Y.; Lapadatescu, M.C.; Larsen, M.J.; Larsen, S.D.; et al. Inhibitor of Streptokinase Gene Expression Improves Survival after Group A Streptococcus Infection in Mice. Proc. Natl. Acad. Sci. USA 2012, 109, 3469–3474. [Google Scholar] [CrossRef] [PubMed]
  312. Frick, I.-M.; Åkesson, P.; Rasmussen, M.; Schmidtchen, A.; Björck, L. SIC, a Secreted Protein of Streptococcus pyogenesThat Inactivates Antibacterial Peptides *. J. Biol. Chem. 2003, 278, 16561–16566. [Google Scholar] [CrossRef]
  313. Barańska-Rybak, W.; Sonesson, A.; Nowicki, R.; Schmidtchen, A. Glycosaminoglycans Inhibit the Antibacterial Activity of LL-37 in Biological Fluids. J. Antimicrob. Chemother. 2006, 57, 260–265. [Google Scholar] [CrossRef] [PubMed]
  314. Nyberg, P.; Rasmussen, M.; Björck, L. A2-Macroglobulin-Proteinase Complexes Protect Streptococcus Pyogenes from Killing by the Antimicrobial Peptide LL-37 *. J. Biol. Chem. 2004, 279, 52820–52823. [Google Scholar] [CrossRef] [PubMed]
  315. Froehlich, B.J.; Bates, C.; Scott, J.R. Streptococcus Pyogenes CovRS Mediates Growth in Iron Starvation and in the Presence of the Human Cationic Antimicrobial Peptide LL-37. J. Bacteriol. 2009, 191, 673–677. [Google Scholar] [CrossRef]
  316. Tran-Winkler, H.J.; Love, J.F.; Gryllos, I.; Wessels, M.R. Signal Transduction through CsrRS Confers an Invasive Phenotype in Group A Streptococcus. PLoS Pathog. 2011, 7, e1002361. [Google Scholar] [CrossRef]
  317. Nizet, V.; Ohtake, T.; Lauth, X.; Trowbridge, J.; Rudisill, J.; Dorschner, R.A.; Pestonjamasp, V.; Piraino, J.; Huttner, K.; Gallo, R.L. Innate Antimicrobial Peptide Protects the Skin from Invasive Bacterial Infection. Nature 2001, 414, 454–457. [Google Scholar] [CrossRef]
  318. Kristian, S.A.; Datta, V.; Weidenmaier, C.; Kansal, R.; Fedtke, I.; Peschel, A.; Gallo, R.L.; Nizet, V. D-Alanylation of Teichoic Acids Promotes Group a Streptococcus Antimicrobial Peptide Resistance, Neutrophil Survival, and Epithelial Cell Invasion. J. Bacteriol. 2005, 187, 6719–6725. [Google Scholar] [CrossRef] [PubMed]
  319. Phelps, H.A.; Neely, M.N. SalY of the Streptococcus Pyogenes Lantibiotic Locus Is Required for Full Virulence and Intracellular Survival in Macrophages. Infect. Immun. 2007, 75, 4541–4551. [Google Scholar] [CrossRef] [PubMed]
  320. Dinulos, J.G.H.; Mentele, L.; Fredericks, L.P.; Dale, B.A.; Darmstadt, G.L. Keratinocyte Expression of Human β Defensin 2 Following Bacterial Infection: Role in Cutaneous Host Defense. Clin. Diagn. Lab. Immunol. 2003, 10, 161–166. [Google Scholar] [CrossRef] [PubMed]
  321. Klinzing, D.C.; Ishmael, N.; Hotopp, J.C.D.; Tettelin, H.; Shields, K.R.; Madoff, L.C.; Puopolo, K.M. The Two-Component Response Regulator LiaR Regulates Cell Wall Stress Responses, Pili Expression and Virulence in Group B Streptococcus. Microbiology 2013, 159, 1521–1534. [Google Scholar] [CrossRef] [PubMed]
  322. Quach, D.; van Sorge, N.M.; Kristian, S.A.; Bryan, J.D.; Shelver, D.W.; Doran, K.S. The CiaR Response Regulator in Group B Streptococcus Promotes Intracellular Survival and Resistance to Innate Immune Defenses. J. Bacteriol. 2009, 191, 2023–2032. [Google Scholar] [CrossRef] [PubMed]
  323. Hamilton, A.; Popham, D.L.; Carl, D.J.; Lauth, X.; Nizet, V.; Jones, A.L. Penicillin-Binding Protein 1a Promotes Resistance of Group B Streptococcus to Antimicrobial Peptides. Infect. Immun. 2006, 74, 6179–6187. [Google Scholar] [CrossRef] [PubMed]
  324. Jones, A.L.; Mertz, R.H.; Carl, D.J.; Rubens, C.E. A Streptococcal Penicillin-Binding Protein Is Critical for Resisting Innate Airway Defenses in the Neonatal Lung1. J. Immunol. 2007, 179, 3196–3202. [Google Scholar] [CrossRef] [PubMed]
  325. Maisey, H.C.; Doran, K.S.; Nizet, V. Recent Advances in Understanding the Molecular Basis of Group B Streptococcus Virulence. Expert Rev. Mol. Med. 2008, 10, e27. [Google Scholar] [CrossRef]
  326. Shabayek, S.; Spellerberg, B. Group B Streptococcal Colonization, Molecular Characteristics, and Epidemiology. Front. Microbiol. 2018, 9, 437. [Google Scholar] [CrossRef]
  327. Poyart, C.; Lamy, M.C.; Boumaila, C.; Fiedler, F.; Trieu-Cuot, P. Regulation of D-Alanyl-Lipoteichoic Acid Biosynthesis in Streptococcus Agalactiae Involves a Novel Two-Component Regulatory System. J. Bacteriol. 2001, 183, 6324–6334. [Google Scholar] [CrossRef] [PubMed]
  328. Saar-Dover, R.; Bitler, A.; Nezer, R.; Shmuel-Galia, L.; Firon, A.; Shimoni, E.; Trieu-Cuot, P.; Shai, Y. D-Alanylation of Lipoteichoic Acids Confers Resistance to Cationic Peptides in Group B Streptococcus by Increasing the Cell Wall Density. PLoS Pathog. 2012, 8, e1002891. [Google Scholar] [CrossRef] [PubMed]
  329. Boneca, I.G.; Dussurget, O.; Cabanes, D.; Nahori, M.-A.; Sousa, S.; Lecuit, M.; Psylinakis, E.; Bouriotis, V.; Hugot, J.-P.; Giovannini, M.; et al. A Critical Role for Peptidoglycan N-Deacetylation in Listeria Evasion from the Host Innate Immune System. Proc. Natl. Acad. Sci. USA 2007, 104, 997–1002. [Google Scholar] [CrossRef] [PubMed]
  330. Aubry, C.; Goulard, C.; Nahori, M.-A.; Cayet, N.; Decalf, J.; Sachse, M.; Boneca, I.G.; Cossart, P.; Dussurget, O. OatA, a Peptidoglycan O-Acetyltransferase Involved in Listeria Monocytogenes Immune Escape, Is Critical for Virulence. J. Infect. Dis. 2011, 204, 731–740. [Google Scholar] [CrossRef]
  331. Carvalho, F.; Atilano, M.L.; Pombinho, R.; Covas, G.; Gallo, R.L.; Filipe, S.R.; Sousa, S.; Cabanes, D. L-Rhamnosylation of Listeria Monocytogenes Wall Teichoic Acids Promotes Resistance to Antimicrobial Peptides by Delaying Interaction with the Membrane. PLOS Pathog. 2015, 11, e1004919. [Google Scholar] [CrossRef] [PubMed]
  332. Meireles, D.; Pombinho, R.; Carvalho, F.; Sousa, S.; Cabanes, D. Listeria Monocytogenes Wall Teichoic Acid Glycosylation Promotes Surface Anchoring of Virulence Factors, Resistance to Antimicrobial Peptides, and Decreased Susceptibility to Antibiotics. Pathogens 2020, 9, 290. [Google Scholar] [CrossRef]
  333. Abachin, E.; Poyart, C.; Pellegrini, E.; Milohanic, E.; Fiedler, F.; Berche, P.; Trieu-Cuot, P. Formation of D-Alanyl-Lipoteichoic Acid Is Required for Adhesion and Virulence of Listeria Monocytogenes. Mol. Microbiol. 2002, 43, 1–14. [Google Scholar] [CrossRef]
  334. Mandin, P.; Fsihi, H.; Dussurget, O.; Vergassola, M.; Milohanic, E.; Toledo-Arana, A.; Lasa, I.; Johansson, J.; Cossart, P. VirR, a Response Regulator Critical for Listeria Monocytogenes Virulence. Mol. Microbiol. 2005, 57, 1367–1380. [Google Scholar] [CrossRef]
  335. Thedieck, K.; Hain, T.; Mohamed, W.; Tindall, B.J.; Nimtz, M.; Chakraborty, T.; Wehland, J.; Jänsch, L. The MprF Protein Is Required for Lysinylation of Phospholipids in Listerial Membranes and Confers Resistance to Cationic Antimicrobial Peptides (CAMPs) on Listeria Monocytogenes. Mol. Microbiol. 2006, 62, 1325–1339. [Google Scholar] [CrossRef]
  336. Collins, B.; Curtis, N.; Cotter, P.D.; Hill, C.; Ross, R.P. The ABC Transporter AnrAB Contributes to the Innate Resistance of Listeria Monocytogenes to Nisin, Bacitracin, and Various β-Lactam Antibiotics. Antimicrob. Agents Chemother. 2010, 54, 4416–4423. [Google Scholar] [CrossRef]
  337. López-Solanilla, E.; González-Zorn, B.; Novella, S.; Vázquez-Boland, J.A.; Rodríguez-Palenzuela, P. Susceptibility of Listeria Monocytogenes to Antimicrobial Peptides. FEMS Microbiol. Lett. 2003, 226, 101–105. [Google Scholar] [CrossRef] [PubMed]
  338. Maloney, E.; Stankowska, D.; Zhang, J.; Fol, M.; Cheng, Q.-J.; Lun, S.; Bishai, W.R.; Rajagopalan, M.; Chatterjee, D.; Madiraju, M.V. The Two-Domain LysX Protein of Mycobacterium Tuberculosis Is Required for Production of Lysinylated Phosphatidylglycerol and Resistance to Cationic Antimicrobial Peptides. PLoS Pathog. 2009, 5, e1000534. [Google Scholar] [CrossRef] [PubMed]
  339. Gao, L.-Y.; Laval, F.; Lawson, E.H.; Groger, R.K.; Woodruff, A.; Morisaki, J.H.; Cox, J.S.; Daffe, M.; Brown, E.J. Requirement for kasB in Mycobacterium Mycolic Acid Biosynthesis, Cell Wall Impermeability and Intracellular Survival: Implications for Therapy. Mol. Microbiol. 2003, 49, 1547–1563. [Google Scholar] [CrossRef] [PubMed]
  340. Sieprawska-Lupa, M.; Mydel, P.; Krawczyk, K.; Wójcik, K.; Puklo, M.; Lupa, B.; Suder, P.; Silberring, J.; Reed, M.; Pohl, J.; et al. Degradation of Human Antimicrobial Peptide LL-37 by Staphylococcus Aureus-Derived Proteinases. Antimicrob. Agents Chemother. 2004, 48, 4673–4679. [Google Scholar] [CrossRef] [PubMed]
  341. Seo, S.; Jung, J.; Kim, C.Y.; Kang, H.; Lee, I.H. Antimicrobial Peptides Encounter Resistance of Aureolysin during Their Action on Staphylococcus Aureus Biofilm. Biotechnol. Bioprocess Eng. 2021, 26, 216–222. [Google Scholar] [CrossRef]
  342. Frey, A.M.; Chaput, D.; Shaw, L.N. Insight into the Human Pathodegradome of the V8 Protease from Staphylococcus Aureus. Cell Rep. 2021, 35, 108930. [Google Scholar] [CrossRef] [PubMed]
  343. Shaw, L.; Golonka, E.; Potempa, J.; Foster, S.J. The Role and Regulation of the Extracellular Proteases of Staphylococcus Aureus. Microbiol. Read. Engl. 2004, 150, 217–228. [Google Scholar] [CrossRef] [PubMed]
  344. Coulter, S.N.; Schwan, W.R.; Ng, E.Y.; Langhorne, M.H.; Ritchie, H.D.; Westbrock-Wadman, S.; Hufnagle, W.O.; Folger, K.R.; Bayer, A.S.; Stover, C.K. Staphylococcus Aureus Genetic Loci Impacting Growth and Survival in Multiple Infection Environments. Mol. Microbiol. 1998, 30, 393–404. [Google Scholar] [CrossRef]
  345. Clarke, S.R.; Foster, S.J. IsdA Protects Staphylococcus Aureus against the Bactericidal Protease Activity of Apolactoferrin. Infect. Immun. 2008, 76, 1518–1526. [Google Scholar] [CrossRef] [PubMed]
  346. Clarke, S.R.; Mohamed, R.; Bian, L.; Routh, A.F.; Kokai-Kun, J.F.; Mond, J.J.; Tarkowski, A.; Foster, S.J. The Staphylococcus Aureus Surface Protein IsdA Mediates Resistance to Innate Defenses of Human Skin. Cell Host Microbe 2007, 1, 199–212. [Google Scholar] [CrossRef]
  347. Cheng, A.G.; Kim, H.K.; Burts, M.L.; Krausz, T.; Schneewind, O.; Missiakas, D.M. Genetic Requirements for Staphylococcus Aureus Abscess Formation and Persistence in Host Tissues. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2009, 23, 3393–3404. [Google Scholar] [CrossRef] [PubMed]
  348. Recsei, P.; Kreiswirth, B.; O’Reilly, M.; Schlievert, P.; Gruss, A.; Novick, R.P. Regulation of Exoprotein Gene Expression in Staphylococcus Aureus by Agr. Mol. Gen. Genet. MGG 1986, 202, 58–61. [Google Scholar] [CrossRef] [PubMed]
  349. Jin, T.; Bokarewa, M.; Foster, T.; Mitchell, J.; Higgins, J.; Tarkowski, A. Staphylococcus Aureus Resists Human Defensins by Production of Staphylokinase, a Novel Bacterial Evasion Mechanism1. J. Immunol. 2004, 172, 1169–1176. [Google Scholar] [CrossRef] [PubMed]
  350. Nguyen, L.T.; Vogel, H.J. Staphylokinase Has Distinct Modes of Interaction with Antimicrobial Peptides, Modulating Its Plasminogen-Activation Properties. Sci. Rep. 2016, 6, 31817. [Google Scholar] [CrossRef] [PubMed]
  351. Kwiecinski, J.; Jacobsson, G.; Karlsson, M.; Zhu, X.; Wang, W.; Bremell, T.; Josefsson, E.; Jin, T. Staphylokinase Promotes the Establishment of Staphylococcus Aureus Skin Infections While Decreasing Disease Severity. J. Infect. Dis. 2013, 208, 990–999. [Google Scholar] [CrossRef] [PubMed]
  352. Jones, C.S.; Anderson, A.C.; Clarke, A.J. Mechanism of Staphylococcus Aureus Peptidoglycan O-Acetyltransferase A as an O-Acyltransferase. Proc. Natl. Acad. Sci. USA 2021, 118, e2103602118. [Google Scholar] [CrossRef] [PubMed]
  353. Bera, A.; Herbert, S.; Jakob, A.; Vollmer, W.; Götz, F. Why Are Pathogenic Staphylococci so Lysozyme Resistant? The Peptidoglycan O-Acetyltransferase OatA Is the Major Determinant for Lysozyme Resistance of Staphylococcus Aureus. Mol. Microbiol. 2005, 55, 778–787. [Google Scholar] [CrossRef] [PubMed]
  354. Bera, A.; Biswas, R.; Herbert, S.; Götz, F. The Presence of Peptidoglycan O-Acetyltransferase in Various Staphylococcal Species Correlates with Lysozyme Resistance and Pathogenicity. Infect. Immun. 2006, 74, 4598–4604. [Google Scholar] [CrossRef]
  355. Baranwal, G.; Mohammad, M.; Jarneborn, A.; Reddy, B.R.; Golla, A.; Chakravarty, S.; Biswas, L.; Götz, F.; Shankarappa, S.; Jin, T.; et al. Impact of Cell Wall Peptidoglycan O-Acetylation on the Pathogenesis of Staphylococcus aureus in Septic Arthritis. Int. J. Med. Microbiol. 2017, 307, 388–397. [Google Scholar] [CrossRef]
  356. Shimada, T.; Park, B.G.; Wolf, A.J.; Brikos, C.; Goodridge, H.S.; Becker, C.A.; Reyes, C.N.; Miao, E.A.; Aderem, A.; Götz, F.; et al. Staphylococcus Aureus Evades the Lysozyme-Based Digestion of Peptidoglycan That Links Phagocytosis and Macrophage IL-1β Secretion. Cell Host Microbe 2010, 7, 38. [Google Scholar] [CrossRef]
  357. Sanchez, M.; Kolar, S.L.; Müller, S.; Reyes, C.N.; Wolf, A.J.; Ogawa, C.; Singhania, R.; Carvalho, D.D.; Arditi, M.; Underhill, D.M.; et al. O-Acetylation of Peptidoglycan Limits Helper T Cell Priming and Permits Staphylococcus aureus Reinfection. Cell Host Microbe 2017, 22, 543–551.e4. [Google Scholar] [CrossRef] [PubMed]
  358. Koprivnjak, T.; Weidenmaier, C.; Peschel, A.; Weiss, J.P. Wall Teichoic Acid Deficiency in Staphylococcus Aureus Confers Selective Resistance to Mammalian Group IIA Phospholipase A2 and Human β-Defensin 3. Infect. Immun. 2008, 76, 2169–2176. [Google Scholar] [CrossRef] [PubMed]
  359. Weidenmaier, C.; Peschel, A.; Kempf, V.A.J.; Lucindo, N.; Yeaman, M.R.; Bayer, A.S. DltABCD- and MprF-Mediated Cell Envelope Modifications of Staphylococcus Aureus Confer Resistance to Platelet Microbicidal Proteins and Contribute to Virulence in a Rabbit Endocarditis Model. Infect. Immun. 2005, 73, 8033–8038. [Google Scholar] [CrossRef] [PubMed]
  360. Weidenmaier, C.; Peschel, A.; Xiong, Y.-Q.; Kristian, S.A.; Dietz, K.; Yeaman, M.R.; Bayer, A.S. Lack of Wall Teichoic Acids in Staphylococcus Aureus Leads to Reduced Interactions with Endothelial Cells and to Attenuated Virulence in a Rabbit Model of Endocarditis. J. Infect. Dis. 2005, 191, 1771–1777. [Google Scholar] [CrossRef] [PubMed]
  361. Weidenmaier, C.; Kokai-Kun, J.F.; Kristian, S.A.; Chanturiya, T.; Kalbacher, H.; Gross, M.; Nicholson, G.; Neumeister, B.; Mond, J.J.; Peschel, A. Role of Teichoic Acids in Staphylococcus Aureus Nasal Colonization, a Major Risk Factor in Nosocomial Infections. Nat. Med. 2004, 10, 243–245. [Google Scholar] [CrossRef] [PubMed]
  362. Gross, M.; Cramton, S.E.; Götz, F.; Peschel, A. Key Role of Teichoic Acid Net Charge in Staphylococcus Aureus Colonization of Artificial Surfaces. Infect. Immun. 2001, 69, 3423–3426. [Google Scholar] [CrossRef] [PubMed]
  363. Kaito, C.; Sekimizu, K. Colony Spreading in Staphylococcus Aureus. J. Bacteriol. 2007, 189, 2553–2557. [Google Scholar] [CrossRef] [PubMed]
  364. Collins, L.V.; Kristian, S.A.; Weidenmaier, C.; Faigle, M.; Van Kessel, K.P.M.; Van Strijp, J.A.G.; Götz, F.; Neumeister, B.; Peschel, A. Staphylococcus Aureus Strains Lacking D-Alanine Modifications of Teichoic Acids Are Highly Susceptible to Human Neutrophil Killing and Are Virulence Attenuated in Mice. J. Infect. Dis. 2002, 186, 214–219. [Google Scholar] [CrossRef]
  365. Peschel, A.; Otto, M.; Jack, R.W.; Kalbacher, H.; Jung, G.; Götz, F. Inactivation of the Dlt Operon in Staphylococcus Aureus Confers Sensitivity to Defensins, Protegrins, and Other Antimicrobial Peptides. J. Biol. Chem. 1999, 274, 8405–8410. [Google Scholar] [CrossRef]
  366. Dunman, P.M.; Murphy, E.; Haney, S.; Palacios, D.; Tucker-Kellogg, G.; Wu, S.; Brown, E.L.; Zagursky, R.J.; Shlaes, D.; Projan, S.J. Transcription Profiling-Based Identification of Staphylococcus aureus Genes Regulated by the Agr and/or sarA Loci. J. Bacteriol. 2001, 183, 7341–7353. [Google Scholar] [CrossRef]
  367. Kraus, D.; Herbert, S.; Kristian, S.A.; Khosravi, A.; Nizet, V.; Götz, F.; Peschel, A. The GraRS Regulatory System Controls Staphylococcus Aureus Susceptibility to Antimicrobial Host Defenses. BMC Microbiol. 2008, 8, 85. [Google Scholar] [CrossRef] [PubMed]
  368. Li, M.; Cha, D.J.; Lai, Y.; Villaruz, A.E.; Sturdevant, D.E.; Otto, M. The Antimicrobial Peptide-Sensing System Aps of Staphylococcus Aureus. Mol. Microbiol. 2007, 66, 1136–1147. [Google Scholar] [CrossRef] [PubMed]
  369. Li, M.; Lai, Y.; Villaruz, A.E.; Cha, D.J.; Sturdevant, D.E.; Otto, M. Gram-Positive Three-Component Antimicrobial Peptide-Sensing System. Proc. Natl. Acad. Sci. USA 2007, 104, 9469–9474. [Google Scholar] [CrossRef] [PubMed]
  370. Peleg, A.Y.; Miyakis, S.; Ward, D.V.; Earl, A.M.; Rubio, A.; Cameron, D.R.; Pillai, S.; Moellering, R.C.; Eliopoulos, G.M. Whole Genome Characterization of the Mechanisms of Daptomycin Resistance in Clinical and Laboratory Derived Isolates of Staphylococcus Aureus. PLoS ONE 2012, 7, e28316. [Google Scholar] [CrossRef] [PubMed]
  371. Peschel, A.; Jack, R.W.; Otto, M.; Collins, L.V.; Staubitz, P.; Nicholson, G.; Kalbacher, H.; Nieuwenhuizen, W.F.; Jung, G.; Tarkowski, A.; et al. Staphylococcus Aureus Resistance to Human Defensins and Evasion of Neutrophil Killing via the Novel Virulence Factor MprF Is Based on Modification of Membrane Lipids with L-Lysine. J. Exp. Med. 2001, 193, 1067–1076. [Google Scholar] [CrossRef] [PubMed]
  372. Shireen, T.; Singh, M.; Das, T.; Mukhopadhyay, K. Differential Adaptive Responses of Staphylococcus Aureus to In Vitro Selection with Different Antimicrobial Peptides. Antimicrob. Agents Chemother. 2013, 57, 5134–5137. [Google Scholar] [CrossRef] [PubMed]
  373. Nishi, H.; Komatsuzawa, H.; Fujiwara, T.; McCallum, N.; Sugai, M. Reduced Content of Lysyl-Phosphatidylglycerol in the Cytoplasmic Membrane Affects Susceptibility to Moenomycin, as Well as Vancomycin, Gentamicin, and Antimicrobial Peptides, in Staphylococcus Aureus. Antimicrob. Agents Chemother. 2004, 48, 4800–4807. [Google Scholar] [CrossRef] [PubMed]
  374. Cheung, G.Y.C.; Fisher, E.L.; McCausland, J.W.; Choi, J.; Collins, J.W.M.; Dickey, S.W.; Otto, M. Antimicrobial Peptide Resistance Mechanism Contributes to Staphylococcus Aureus Infection. J. Infect. Dis. 2018, 217, 1153–1159. [Google Scholar] [CrossRef]
  375. Kupferwasser, L.I.; Skurray, R.A.; Brown, M.H.; Firth, N.; Yeaman, M.R.; Bayer, A.S. Plasmid-Mediated Resistance to Thrombin-Induced Platelet Microbicidal Protein in Staphylococci: Role of the qacA Locus. Antimicrob. Agents Chemother. 1999, 43, 2395–2399. [Google Scholar] [CrossRef]
  376. Falord, M.; Karimova, G.; Hiron, A.; Msadek, T. GraXSR Proteins Interact with the VraFG ABC Transporter to Form a Five-Component System Required for Cationic Antimicrobial Peptide Sensing and Resistance in Staphylococcus aureus. Antimicrob. Agents Chemother. 2012, 56, 1047–1058. [Google Scholar] [CrossRef]
  377. Cao, R.; Su, H.; Wei, Z.; He, Z.; Pan, T.; Li, Y.; Sun, B. An Induced Mutation of ABC-Transporter Component VraF(K84E) Contributes to Vancomycin Resistance and Virulence in Staphylococcus Aureus Strain MW2. Int. J. Med. Microbiol. IJMM 2024, 315, 151624. [Google Scholar] [CrossRef] [PubMed]
  378. Vuong, C.; Voyich, J.M.; Fischer, E.R.; Braughton, K.R.; Whitney, A.R.; DeLeo, F.R.; Otto, M. Polysaccharide Intercellular Adhesin (PIA) Protects Staphylococcus Epidermidis against Major Components of the Human Innate Immune System. Cell. Microbiol. 2004, 6, 269–275. [Google Scholar] [CrossRef] [PubMed]
  379. Cheung, G.Y.C.; Otto, M. Understanding the Significance of Staphylococcus Epidermidis Bacteremia in Babies and Children. Curr. Opin. Infect. Dis. 2010, 23, 208–216. [Google Scholar] [CrossRef] [PubMed]
  380. Buchanan, J.T.; Stannard, J.A.; Lauth, X.; Ostland, V.E.; Powell, H.C.; Westerman, M.E.; Nizet, V. Streptococcus Iniae Phosphoglucomutase Is a Virulence Factor and a Target for Vaccine Development. Infect. Immun. 2005, 73, 6935–6944. [Google Scholar] [CrossRef]
  381. Allen, J.P.; Neely, M.N. CpsY Influences Streptococcus Iniae Cell Wall Adaptations Important for Neutrophil Intracellular Survival. Infect. Immun. 2012, 80, 1707–1715. [Google Scholar] [CrossRef]
  382. Ouyang, J.; Tian, X.-L.; Versey, J.; Wishart, A.; Li, Y.-H. The BceABRS Four-Component System Regulates the Bacitracin-Induced Cell Envelope Stress Response in Streptococcus Mutans. Antimicrob. Agents Chemother. 2010, 54, 3895–3906. [Google Scholar] [CrossRef]
  383. Mazda, Y.; Kawada-Matsuo, M.; Kanbara, K.; Oogai, Y.; Shibata, Y.; Yamashita, Y.; Miyawaki, S.; Komatsuzawa, H. Association of CiaRH with Resistance of Streptococcus Mutans to Antimicrobial Peptides in Biofilms. Mol. Oral Microbiol. 2012, 27, 124–135. [Google Scholar] [CrossRef]
  384. Du, J.; Huang, S.; Wu, M.; Chen, S.; Zhou, W.; Zhan, L.; Huang, X. Dlt Operon Regulates Physiological Function and Cariogenic Virulence in Streptococcus Mutans. Future Microbiol. 2023, 18, 225–233. [Google Scholar] [CrossRef]
  385. Shaper, M.; Hollingshead, S.K.; Benjamin, W.H.; Briles, D.E. PspA Protects Streptococcus Pneumoniae from Killing by Apolactoferrin, and Antibody to PspA Enhances Killing of Pneumococci by Apolactoferrin. Infect. Immun. 2004, 72, 5031–5040. [Google Scholar] [CrossRef]
  386. Kovács, M.; Halfmann, A.; Fedtke, I.; Heintz, M.; Peschel, A.; Vollmer, W.; Hakenbeck, R.; Brückner, R. A Functional Dlt Operon, Encoding Proteins Required for Incorporation of d-Alanine in Teichoic Acids in Gram-Positive Bacteria, Confers Resistance to Cationic Antimicrobial Peptides in Streptococcus Pneumoniae. J. Bacteriol. 2006, 188, 5797–5805. [Google Scholar] [CrossRef]
  387. Zafar, M.A.; Hammond, A.J.; Hamaguchi, S.; Wu, W.; Kono, M.; Zhao, L.; Weiser, J.N. Identification of Pneumococcal Factors Affecting Pneumococcal Shedding Shows That the Dlt Locus Promotes Inflammation and Transmission. mBio 2019, 10, e01032-19. [Google Scholar] [CrossRef] [PubMed]
  388. Wartha, F.; Beiter, K.; Albiger, B.; Fernebro, J.; Zychlinsky, A.; Normark, S.; Henriques-Normark, B. Capsule and D-Alanylated Lipoteichoic Acids Protect Streptococcus Pneumoniae against Neutrophil Extracellular Traps. Cell. Microbiol. 2007, 9, 1162–1171. [Google Scholar] [CrossRef] [PubMed]
  389. Majchrzykiewicz, J.A.; Kuipers, O.P.; Bijlsma, J.J.E. Generic and Specific Adaptive Responses of Streptococcus Pneumoniae to Challenge with Three Distinct Antimicrobial Peptides, Bacitracin, LL-37, and Nisin. Antimicrob. Agents Chemother. 2010, 54, 440–451. [Google Scholar] [CrossRef] [PubMed]
  390. Yang, H.-B.; Hou, W.-T.; Cheng, M.-T.; Jiang, Y.-L.; Chen, Y.; Zhou, C.-Z. Structure of a MacAB-like Efflux Pump from Streptococcus Pneumoniae—Nature Communications. Nat. Commun. 2018, 9, 196. [Google Scholar] [CrossRef] [PubMed]
  391. Zähner, D.; Zhou, X.; Chancey, S.T.; Pohl, J.; Shafer, W.M.; Stephens, D.S. Human Antimicrobial Peptide LL-37 Induces MefE/Mel-Mediated Macrolide Resistance in Streptococcus Pneumoniae. Antimicrob. Agents Chemother. 2010, 54, 3516–3519. [Google Scholar] [CrossRef] [PubMed]
  392. Davis, K.M.; Akinbi, H.T.; Standish, A.J.; Weiser, J.N. Resistance to Mucosal Lysozyme Compensates for the Fitness Deficit of Peptidoglycan Modifications by Streptococcus Pneumoniae. PLoS Pathog. 2008, 4, e1000241. [Google Scholar] [CrossRef] [PubMed]
  393. Xie, F.; Zan, Y.; Zhang, Y.; Zheng, N.; Yan, Q.; Zhang, W.; Zhang, H.; Jin, M.; Chen, F.; Zhang, X.; et al. The Cysteine Protease ApdS from Streptococcus Suis Promotes Evasion of Innate Immune Defenses by Cleaving the Antimicrobial Peptide Cathelicidin LL-37. J. Biol. Chem. 2019, 294, 17962–17977. [Google Scholar] [CrossRef] [PubMed]
  394. Wichgers Schreur, P.J.; van Weeghel, C.; Rebel, J.M.J.; Smits, M.A.; van Putten, J.P.M.; Smith, H.E. Lysozyme Resistance in Streptococcus Suis Is Highly Variable and Multifactorial. PLoS ONE 2012, 7, e36281. [Google Scholar] [CrossRef] [PubMed]
  395. LaRock, C.N.; Döhrmann, S.; Todd, J.; Corriden, R.; Olson, J.; Johannssen, T.; Lepenies, B.; Gallo, R.L.; Ghosh, P.; Nizet, V. Group A Streptococcal M1 Protein Sequesters Cathelicidin to Evade Innate Immune Killing. Cell Host Microbe 2015, 18, 471–477. [Google Scholar] [CrossRef]
  396. Yadav, R.; Govindan, S.; Daczkowski, C.; Mesecar, A.; Chakravarthy, S.; Noinaj, N. Structural Insight into the Dual Function of LbpB in Mediating Neisserial Pathogenesis. eLife 2021, 10, e71683. [Google Scholar] [CrossRef]
  397. Morgenthau, A.; Partha, S.K.; Adamiak, P.; Schryvers, A.B. The Specificity of Protection against Cationic Antimicrobial Peptides by Lactoferrin Binding Protein B. Biometals Int. J. Role Met. Ions Biol. Biochem. Med. 2014, 27, 923–933. [Google Scholar] [CrossRef]
  398. Ostan, N.K.H.; Cole, G.B.; Wang, F.Z.; Reichheld, S.E.; Moore, G.; Pan, C.; Yu, R.; Lai, C.C.-L.; Sharpe, S.; Lee, H.O.; et al. A Secreted Bacterial Protein Protects Bacteria from Cationic Antimicrobial Peptides by Entrapment in Phase-Separated Droplets. PNAS Nexus 2024, 3, pgae139. [Google Scholar] [CrossRef] [PubMed]
  399. Zheng, Y.; Zhang, X.; Wang, X.; Wang, L.; Zhang, J.; Yin, Y. ComE, an Essential Response Regulator, Negatively Regulates the Expression of the Capsular Polysaccharide Locus and Attenuates the Bacterial Virulence in Streptococcus Pneumoniae. Front. Microbiol. 2017, 8, 277. [Google Scholar] [CrossRef] [PubMed]
  400. Cress, B.F.; Englaender, J.A.; He, W.; Kasper, D.; Linhardt, R.J.; Koffas, M.A.G. Masquerading Microbial Pathogens: Capsular Polysaccharides Mimic Host-Tissue Molecules. FEMS Microbiol. Rev. 2014, 38, 660–697. [Google Scholar] [CrossRef]
  401. Gao, S.; Jin, W.; Quan, Y.; Li, Y.; Shen, Y.; Yuan, S.; Yi, L.; Wang, Y.; Wang, Y. Bacterial Capsules: Occurrence, Mechanism, and Function. Npj Biofilms Microbiomes 2024, 10, 21. [Google Scholar] [CrossRef] [PubMed]
  402. Fleeman, R.M.; Macias, L.A.; Brodbelt, J.S.; Davies, B.W. Defining Principles That Influence Antimicrobial Peptide Activity against Capsulated Klebsiella Pneumoniae. Proc. Natl. Acad. Sci. USA 2020, 117, 27620–27626. [Google Scholar] [CrossRef]
  403. Claessen, D.; Errington, J. Cell Wall Deficiency as a Coping Strategy for Stress. Trends Microbiol. 2019, 27, 1025–1033. [Google Scholar] [CrossRef]
  404. de Tejada, G.M.; Sánchez-Gómez, S.; Kowalski, I.; Kaconis, Y.; Andrä, J.; Schürholz, T.; Hornef, M.; Dupont, A.; Garidel, P.; Gutsmann, T.; et al. Bacterial Cell Wall Compounds as Promising Targets of Antimicrobial Agents I. Antimicrobial Peptides and Lipopolyamines. Curr. Drug Targets 2012, 13, 1121–1130. [Google Scholar] [CrossRef]
  405. Nikaido, H. Molecular Basis of Bacterial Outer Membrane Permeability Revisited. Microbiol. Mol. Biol. Rev. MMBR 2003, 67, 593–656. [Google Scholar] [CrossRef]
  406. Papo, N.; Shai, Y. A Molecular Mechanism for Lipopolysaccharide Protection of Gram-Negative Bacteria from Antimicrobial Peptides*. J. Biol. Chem. 2005, 280, 10378–10387. [Google Scholar] [CrossRef]
  407. Hancock, R.E.W.; Rozek, A. Role of Membranes in the Activities of Antimicrobial Cationic Peptides. FEMS Microbiol. Lett. 2002, 206, 143–149. [Google Scholar] [CrossRef]
  408. Caroff, M.; Novikov, A. Lipopolysaccharides: Structure, Function and Bacterial Identifications. OCL 2020, 27, 31. [Google Scholar] [CrossRef]
  409. Som, N.; Reddy, M. Cross-Talk between Phospholipid Synthesis and Peptidoglycan Expansion by a Cell Wall Hydrolase. Proc. Natl. Acad. Sci. USA 2023, 120, e2300784120. [Google Scholar] [CrossRef]
  410. Thomanek, N.; Arends, J.; Lindemann, C.; Barkovits, K.; Meyer, H.E.; Marcus, K.; Narberhaus, F. Intricate Crosstalk Between Lipopolysaccharide, Phospholipid and Fatty Acid Metabolism in Escherichia Coli Modulates Proteolysis of LpxC. Front. Microbiol. 2019, 9, 3285. [Google Scholar] [CrossRef] [PubMed]
  411. Wang, J.; Ma, W.; Wang, X. Insights into the Structure of Escherichia Coli Outer Membrane as the Target for Engineering Microbial Cell Factories. Microb. Cell Factories 2021, 20, 73. [Google Scholar] [CrossRef] [PubMed]
  412. Joo, H.; Eom, H.; Cho, Y.; Rho, M.; Song, W.J. Discovery and Characterization of Polymyxin-Resistance Genes pmrE and pmrF from Sediment and Seawater Microbiome. Microbiol. Spectr. 2023, 11, e02736-22. [Google Scholar] [CrossRef] [PubMed]
  413. Moran, A.P.; Prendergast, M.M.; Appelmelk, B.J. Molecular Mimicry of Host Structures by Bacterial Lipopolysaccharides and Its Contribution to Disease. FEMS Immunol. Med. Microbiol. 1996, 16, 105–115. [Google Scholar] [CrossRef]
  414. Peterson, A.A.; Haug, A.; McGroarty, E.J. Physical Properties of Short- and Long-O-Antigen-Containing Fractions of Lipopolysaccharide from Escherichia Coli 0111:B4. J. Bacteriol. 1986, 165, 116–122. [Google Scholar] [CrossRef]
  415. Stinavage, P.; Martin, L.E.; Spitznagel, J.K. O Antigen and Lipid A Phosphoryl Groups in Resistance of Salmonella Typhimurium LT-2 to Nonoxidative Killing in Human Polymorphonuclear Neutrophils. Infect. Immun. 1989, 57, 3894–3900. [Google Scholar] [CrossRef]
  416. Capodici, C.; Chen, S.; Sidorczyk, Z.; Elsbach, P.; Weiss, J. Effect of Lipopolysaccharide (LPS) Chain Length on Interactions of Bactericidal/Permeability-Increasing Protein and Its Bioactive 23-Kilodalton NH2-Terminal Fragment with Isolated LPS and Intact Proteus Mirabilis and Escherichia Coli. Infect. Immun. 1994, 62, 259–265. [Google Scholar] [CrossRef]
  417. Allen, C.A.; Adams, L.G.; Ficht, T.A. Transposon-Derived Brucella Abortus Rough Mutants Are Attenuated and Exhibit Reduced Intracellular Survival. Infect. Immun. 1998, 66, 1008–1016. [Google Scholar] [CrossRef] [PubMed]
  418. Clifton, L.A.; Ciesielski, F.; Skoda, M.W.A.; Paracini, N.; Holt, S.A.; Lakey, J.H. The Effect of Lipopolysaccharide Core Oligosaccharide Size on the Electrostatic Binding of Antimicrobial Proteins to Models of the Gram Negative Bacterial Outer Membrane. Langmuir 2016, 32, 3485–3494. [Google Scholar] [CrossRef] [PubMed]
  419. Adams, P.G.; Lamoureux, L.; Swingle, K.L.; Mukundan, H.; Montaño, G.A. Lipopolysaccharide-Induced Dynamic Lipid Membrane Reorganization: Tubules, Perforations, and Stacks. Biophys. J. 2014, 106, 2395–2407. [Google Scholar] [CrossRef] [PubMed]
  420. Wilkinson, S.G. Bacterial Lipopolysaccharides—Themes and Variations. Prog. Lipid Res. 1996, 35, 283–343. [Google Scholar] [CrossRef] [PubMed]
  421. Papahadjopoulos, D.; Portis, A.; Pangborn, W. Calcium-Induced Lipid Phase Transitions and Membrane Fusion. Ann. N. Y. Acad. Sci. 1978, 308, 50–66. [Google Scholar] [CrossRef] [PubMed]
  422. Brandenburg, K.; Koch, M.H.J.; Seydel, U. Phase Diagram of Lipid A from Salmonella Minnesota and Escherichia Coli Rough Mutant Lipopolysaccharide. J. Struct. Biol. 1990, 105, 11–21. [Google Scholar] [CrossRef] [PubMed]
  423. Seydel, U.; Koch, M.H.; Brandenburg, K. Structural Polymorphisms of Rough Mutant Lipopolysaccharides Rd to Ra from Salmonella Minnesota. J. Struct. Biol. 1993, 110, 232–243. [Google Scholar] [CrossRef]
  424. Jia, W.; Zoeiby, A.E.; Petruzziello, T.N.; Jayabalasingham, B.; Seyedirashti, S.; Bishop, R.E. Lipid Trafficking Controls Endotoxin Acylation in Outer Membranes of Escherichia coli*. J. Biol. Chem. 2004, 279, 44966–44975. [Google Scholar] [CrossRef] [PubMed]
  425. Hancock, R.E.W.; Scott, M.G. The Role of Antimicrobial Peptides in Animal Defenses. Proc. Natl. Acad. Sci. USA 2000, 97, 8856–8861. [Google Scholar] [CrossRef]
  426. Morrison, D.C.; Leive, L. Fractions of Lipopolysaccharide from Escherichia Coli O111:B4 Prepared by Two Extraction Procedures. J. Biol. Chem. 1975, 250, 2911–2919. [Google Scholar] [CrossRef]
  427. Brandenburg, K.; Seydel, U. Physical Aspects of Structure and Function of Membranes Made from Lipopolysaccharides and Free Lipid A. Biochim. Biophys. Acta BBA—Biomembr. 1984, 775, 225–238. [Google Scholar] [CrossRef]
  428. Yeh, H.Y.; Jacobs, D.M. Characterization of Lipopolysaccharide Fractions and Their Interactions with Cells and Model Membranes. J. Bacteriol. 1992, 174, 336–341. [Google Scholar] [CrossRef]
  429. Rana, F.R.; Blazyk, J. Interactions between the Antimicrobial Peptide, Magainin 2, and Salmonella Typhimurium Lipopolysaccharides. FEBS Lett. 1991, 293, 11–15. [Google Scholar] [CrossRef] [PubMed]
  430. Ginez, L.D.; Osorio, A.; Vázquez-Ramírez, R.; Arenas, T.; Mendoza, L.; Camarena, L.; Poggio, S. Changes in Fluidity of the E. Coli Outer Membrane in Response to Temperature, Divalent Cations and Polymyxin-B Show Two Different Mechanisms of Membrane Fluidity Adaptation. FEBS J. 2022, 289, 3550–3567. [Google Scholar] [CrossRef]
  431. Hac-Wydro, K.; Wydro, P. The Influence of Fatty Acids on Model Cholesterol/Phospholipid Membranes. Chem. Phys. Lipids 2007, 150, 66–81. [Google Scholar] [CrossRef] [PubMed]
  432. Vaara, M. Agents That Increase the Permeability of the Outer Membrane. Microbiol. Rev. 1992, 56, 395–411. [Google Scholar] [CrossRef]
  433. Seydel, U.; Labischinski, H.; Kastowsky, M.; Brandenburg, K. Phase Behavior, Supramolecular Structure, and Molecular Conformation of Lipopolysaccharide. Immunobiology 1993, 187, 191–211. [Google Scholar] [CrossRef]
  434. Cullis, P.R.; Hope, M.J. Physical Properties and Functional Roles of Lipids in Membranes. In Biochemistry of Lipids and Membranes; Vance, D.E., Vance, J.E., Eds.; Benjamin-Cummings: Menlo Park, CA, USA, 1985. [Google Scholar]
  435. Kim, S.; Patel, D.S.; Park, S.; Slusky, J.; Klauda, J.B.; Widmalm, G.; Im, W. Bilayer Properties of Lipid A from Various Gram-Negative Bacteria. Biophys. J. 2016, 111, 1750–1760. [Google Scholar] [CrossRef]
  436. Emiola, A.; George, J.; Andrews, S.S. A Complete Pathway Model for Lipid A Biosynthesis in Escherichia Coli. PLoS ONE 2015, 10, e0121216. [Google Scholar] [CrossRef]
  437. Clementz, T.; Bednarski, J.J.; Raetz, C.R. Function of the htrB High Temperature Requirement Gene of Escherichia Coli in the Acylation of Lipid A: HtrB Catalyzed Incorporation of Laurate. J. Biol. Chem. 1996, 271, 12095–12102. [Google Scholar] [CrossRef]
  438. Schilling, B.; Hunt, J.; Gibson, B.W.; Apicella, M.A. Site-Specific Acylation Changes in the Lipid A of Escherichia Coli lpxL Mutants Grown at High Temperatures. Innate Immun. 2014, 20, 269–282. [Google Scholar] [CrossRef]
  439. Steeghs, L.; de Cock, H.; Evers, E.; Zomer, B.; Tommassen, J.; Ley, P. van der Outer Membrane Composition of a Lipopolysaccharide-Deficient Neisseria Meningitidis Mutant. EMBO J. 2001, 20, 6937–6945. [Google Scholar] [CrossRef] [PubMed]
  440. Somerville, J.E.; Cassiano, L.; Bainbridge, B.; Cunningham, M.D.; Darveau, R.P. A Novel Escherichia Coli Lipid A Mutant That Produces an Antiinflammatory Lipopolysaccharide. J. Clin. Invest. 1996, 97, 359–365. [Google Scholar] [CrossRef]
  441. Khan, S.A.; Everest, P.; Servos, S.; Foxwell, N.; Zähringer, U.; Brade, H.; Rietschel, E.T.; Dougan, G.; Charles, I.G.; Maskell, D.J. A Lethal Role for Lipid A in Salmonella Infections. Mol. Microbiol. 1998, 29, 571–579. [Google Scholar] [CrossRef] [PubMed]
  442. d’Hauteville, H.; Khan, S.; Maskell, D.J.; Kussak, A.; Weintraub, A.; Mathison, J.; Ulevitch, R.J.; Wuscher, N.; Parsot, C.; Sansonetti, P.J. Two msbB Genes Encoding Maximal Acylation of Lipid A Are Required for Invasive Shigella Flexneri to Mediate Inflammatory Rupture and Destruction of the Intestinal Epithelium1. J. Immunol. 2002, 168, 5240–5251. [Google Scholar] [CrossRef]
  443. Murray, S.R.; Bermudes, D.; de Felipe, K.S.; Low, K.B. Extragenic Suppressors of Growth Defects in msbB Salmonella. J. Bacteriol. 2001, 183, 5554–5561. [Google Scholar] [CrossRef]
  444. Karow, M.; Fayet, O.; Georgopoulos, C. The Lethal Phenotype Caused by Null Mutations in the Escherichia Coli htrB Gene Is Suppressed by Mutations in the accBC Operon, Encoding Two Subunits of Acetyl Coenzyme A Carboxylase. J. Bacteriol. 1992, 174, 7407–7418. [Google Scholar] [CrossRef] [PubMed]
  445. Karow, M.; Georgopoulos, C. Isolation and Characterization of the Escherichia Coli msbB Gene, a Multicopy Suppressor of Null Mutations in the High-Temperature Requirement Gene htrB. J. Bacteriol. 1992, 174, 702–710. [Google Scholar] [CrossRef]
  446. Somerville, J.E.; Cassiano, L.; Darveau, R.P. Escherichia Coli msbB Gene as a Virulence Factor and a Therapeutic Target. Infect. Immun. 1999, 67, 6583–6590. [Google Scholar] [CrossRef]
  447. Vorachek-Warren, M.K.; Ramirez, S.; Cotter, R.J.; Raetz, C.R.H. A Triple Mutant of Escherichia Coli Lacking Secondary Acyl Chains on Lipid A. J. Biol. Chem. 2002, 277, 14194–14205. [Google Scholar] [CrossRef]
  448. Jiang, X.; Yang, K.; Han, M.-L.; Yuan, B.; Li, J.; Gong, B.; Velkov, T.; Schreiber, F.; Wang, L.; Li, J. Outer Membranes of Polymyxin-Resistant Acinetobacter Baumannii with Phosphoethanolamine-Modified Lipid A and Lipopolysaccharide Loss Display Different Atomic-Scale Interactions with Polymyxins. ACS Infect. Dis. 2020, 6, 2698–2708. [Google Scholar] [CrossRef] [PubMed]
  449. Wang, X.; Ribeiro, A.A.; Guan, Z.; McGrath, S.C.; Cotter, R.J.; Raetz, C.R.H. Structure and Biosynthesis of Free Lipid A Molecules That Replace Lipopolysaccharide in Francisella Novicida. Biochemistry 2006, 45, 14427–14440. [Google Scholar] [CrossRef]
  450. Akira, S. TLR Signaling. Curr. Top. Microbiol. Immunol. 2006, 311, 1–16. [Google Scholar] [CrossRef] [PubMed]
  451. Dalebroux, Z.D.; Matamouros, S.; Whittington, D.; Bishop, R.E.; Miller, S.I. PhoPQ Regulates Acidic Glycerophospholipid Content of the Salmonella Typhimurium Outer Membrane. Proc. Natl. Acad. Sci. USA 2014, 111, 1963–1968. [Google Scholar] [CrossRef] [PubMed]
  452. Pilione, M.R.; Pishko, E.J.; Preston, A.; Maskell, D.J.; Harvill, E.T. pagP Is Required for Resistance to Antibody-Mediated Complement Lysis during Bordetella Bronchiseptica Respiratory Infection. Infect. Immun. 2004, 72, 2837–2842. [Google Scholar] [CrossRef] [PubMed]
  453. Kawasaki, K.; Ernst, R.K.; Miller, S.I. Deacylation and Palmitoylation of Lipid A by Salmonellae Outer Membrane Enzymes Modulate Host Signaling through Toll-like Receptor 4. J. Endotoxin Res. 2004, 10, 439–444. [Google Scholar] [CrossRef] [PubMed]
  454. Lipid A Acylation and Bacterial Resistance against Vertebrate Antimicrobial Peptides. Cell 1998, 95, 189–198. [CrossRef] [PubMed]
  455. Preston, A.; Maxim, E.; Toland, E.; Pishko, E.J.; Harvill, E.T.; Caroff, M.; Maskell, D.J. Bordetella Bronchiseptica PagP Is a Bvg-Regulated Lipid A Palmitoyl Transferase That Is Required for Persistent Colonization of the Mouse Respiratory Tract. Mol. Microbiol. 2003, 48, 725–736. [Google Scholar] [CrossRef] [PubMed]
  456. Trent, M.S.; Pabich, W.; Raetz, C.R.H.; Miller, S.I. A PhoP/PhoQ-Induced Lipase (PagL) That Catalyzes 3-O-Deacylation of Lipid A Precursors in Membranes ofSalmonella Typhimurium *. J. Biol. Chem. 2001, 276, 9083–9092. [Google Scholar] [CrossRef]
  457. Rutten, L.; Mannie, J.-P.B.A.; Stead, C.M.; Raetz, C.R.H.; Reynolds, C.M.; Bonvin, A.M.J.J.; Tommassen, J.P.; Egmond, M.R.; Trent, M.S.; Gros, P. Active-Site Architecture and Catalytic Mechanism of the Lipid A Deacylase LpxR of Salmonella Typhimurium. Proc. Natl. Acad. Sci. USA 2009, 106, 1960–1964. [Google Scholar] [CrossRef]
  458. Gibbons, H.S.; Kalb, S.R.; Cotter, R.J.; Raetz, C.R.H. Role of Mg2+ and pH in the Modification of Salmonella Lipid A after Endocytosis by Macrophage Tumour Cells. Mol. Microbiol. 2005, 55, 425–440. [Google Scholar] [CrossRef] [PubMed]
  459. Gibbons, H.S.; Lin, S.; Cotter, R.J.; Raetz, C.R. Oxygen Requirement for the Biosynthesis of the S-2-Hydroxymyristate Moiety in Salmonella Typhimurium Lipid A. Function of LpxO, A New Fe2+/Alpha-Ketoglutarate-Dependent Dioxygenase Homologue. J. Biol. Chem. 2000, 275, 32940–32949. [Google Scholar] [CrossRef]
  460. MacArthur, I.; Jones, J.W.; Goodlett, D.R.; Ernst, R.K.; Preston, A. Role of pagL and lpxO in Bordetella Bronchiseptica Lipid A Biosynthesis. J. Bacteriol. 2011, 193, 4726–4735. [Google Scholar] [CrossRef] [PubMed]
  461. Ogunlana, L.; Kaur, D.; Shaw, L.P.; Jangir, P.; Walsh, T.; Uphoff, S.; MacLean, R.C. Regulatory Fine-Tuning of Mcr-1 Increases Bacterial Fitness and Stabilises Antibiotic Resistance in Agricultural Settings. ISME J. 2023, 17, 2058–2069. [Google Scholar] [CrossRef] [PubMed]
  462. Nang, S.C.; Li, J.; Velkov, T. The Rise and Spread of Mcr Plasmid-Mediated Polymyxin Resistance. Crit. Rev. Microbiol. 2019, 45, 131–161. [Google Scholar] [CrossRef] [PubMed]
  463. Shprung, T.; Wani, N.A.; Wilmes, M.; Mangoni, M.L.; Bitler, A.; Shimoni, E.; Sahl, H.-G.; Shai, Y. Opposing Effects of PhoPQ and PmrAB on the Properties of Salmonella Enterica Serovar Typhimurium: Implications on Resistance to Antimicrobial Peptides. Biochemistry 2021, 60, 2943–2955. [Google Scholar] [CrossRef]
  464. Rubin, E.J.; Herrera, C.M.; Crofts, A.A.; Trent, M.S. PmrD Is Required for Modifications to Escherichia Coli Endotoxin That Promote Antimicrobial Resistance. Antimicrob. Agents Chemother. 2015, 59, 2051–2061. [Google Scholar] [CrossRef]
  465. Huang, J.; Li, C.; Song, J.; Velkov, T.; Wang, L.; Zhu, Y.; Li, J. Regulating Polymyxin Resistance in Gram-Negative Bacteria: Roles of Two-Component Systems PhoPQ and PmrAB. Future Microbiol. 2020, 15, 445–459. [Google Scholar] [CrossRef]
  466. Gunn, J.S. Bacterial Modification of LPS and Resistance to Antimicrobial Peptides. J. Endotoxin Res. 2001, 7, 57–62. [Google Scholar] [CrossRef]
  467. Raetz, C.R.H.; Reynolds, C.M.; Trent, M.S.; Bishop, R.E. Lipid A Modification Systems in Gram-Negative Bacteria. Annu. Rev. Biochem. 2007, 76, 295–329. [Google Scholar] [CrossRef]
  468. Yan, A.; Guan, Z.; Raetz, C.R.H. An Undecaprenyl Phosphate-Aminoarabinose Flippase Required for Polymyxin Resistance in Escherichia Coli*. J. Biol. Chem. 2007, 282, 36077–36089. [Google Scholar] [CrossRef] [PubMed]
  469. Murray, S.R.; Ernst, R.K.; Bermudes, D.; Miller, S.I.; Low, K.B. PmrA(Con) Confers pmrHFIJKL-Dependent EGTA and Polymyxin Resistance on msbB Salmonella by Decorating Lipid A with Phosphoethanolamine. J. Bacteriol. 2007, 189, 5161–5169. [Google Scholar] [CrossRef] [PubMed]
  470. Shearer, J.; Jefferies, D.; Khalid, S. Outer Membrane Proteins OmpA, FhuA, OmpF, EstA, BtuB, and OmpX Have Unique Lipopolysaccharide Fingerprints. J. Chem. Theory Comput. 2019, 15, 2608–2619. [Google Scholar] [CrossRef] [PubMed]
  471. Blair, J.M.A.; Zeth, K.; Bavro, V.N.; Sancho-Vaello, E. The Role of Bacterial Transport Systems in the Removal of Host Antimicrobial Peptides in Gram-Negative Bacteria. FEMS Microbiol. Rev. 2022, 46, fuac032. [Google Scholar] [CrossRef] [PubMed]
  472. Zgurskaya, H.I.; Krishnamoorthy, G.; Ntreh, A.; Lu, S. Mechanism and Function of the Outer Membrane Channel TolC in Multidrug Resistance and Physiology of Enterobacteria. Front. Microbiol. 2011, 2, 189. [Google Scholar] [CrossRef] [PubMed]
  473. Housden, N.G.; Webby, M.N.; Lowe, E.D.; El-Baba, T.J.; Kaminska, R.; Redfield, C.; Robinson, C.V.; Kleanthous, C. Toxin Import through the Antibiotic Efflux Channel TolC. Nat. Commun. 2021, 12, 4625. [Google Scholar] [CrossRef] [PubMed]
  474. Zhang, P.; Ye, Z.; Ye, C.; Zou, H.; Gao, Z.; Pan, J. OmpW Is Positively Regulated by Iron via Fur, and Negatively Regulated by SoxS Contribution to Oxidative Stress Resistance in Escherichia coli. Microb. Pathog. 2020, 138, 103808. [Google Scholar] [CrossRef] [PubMed]
  475. Lee, K.; Yong, D.; Jeong, S.H.; Chong, Y. Multidrug-Resistant Acinetobacter Spp.: Increasingly Problematic Nosocomial Pathogens. Yonsei Med. J. 2011, 52, 879–891. [Google Scholar] [CrossRef] [PubMed]
  476. Rajagopal, M.; Walker, S. Envelope Structures of Gram-Positive Bacteria. Curr. Top. Microbiol. Immunol. 2017, 404, 1–44. [Google Scholar] [CrossRef]
  477. Kim, S.J.; Chang, J.; Singh, M. Peptidoglycan Architecture of Gram-Positive Bacteria by Solid-State NMR. Biochim. Biophys. Acta 2015, 1848, 350–362. [Google Scholar] [CrossRef]
  478. Silhavy, T.J.; Kahne, D.; Walker, S. The Bacterial Cell Envelope. Cold Spring Harb. Perspect. Biol. 2010, 2, a000414. [Google Scholar] [CrossRef] [PubMed]
  479. Neuhaus, F.C.; Baddiley, J. A Continuum of Anionic Charge: Structures and Functions of d-Alanyl-Teichoic Acids in Gram-Positive Bacteria. Microbiol. Mol. Biol. Rev. MMBR 2003, 67, 686–723. [Google Scholar] [CrossRef] [PubMed]
  480. Grangette, C.; Müller-Alouf, H.; Hols, P.; Goudercourt, D.; Delcour, J.; Turneer, M.; Mercenier, A. Enhanced Mucosal Delivery of Antigen with Cell Wall Mutants of Lactic Acid Bacteria. Infect. Immun. 2004, 72, 2731–2737. [Google Scholar] [CrossRef] [PubMed]
  481. de Vos, W.M. Lipotechoic Acid in Lactobacilli: D-Alanine Makes the Difference. Proc. Natl. Acad. Sci. USA 2005, 102, 10763–10764. [Google Scholar] [CrossRef] [PubMed]
  482. Kristian, S.A.; Lauth, X.; Nizet, V.; Goetz, F.; Neumeister, B.; Peschel, A.; Landmann, R. Alanylation of Teichoic Acids Protects Staphylococcus Aureus against Toll-like Receptor 2-Dependent Host Defense in a Mouse Tissue Cage Infection Model. J. Infect. Dis. 2003, 188, 414–423. [Google Scholar] [CrossRef] [PubMed]
  483. Fischer, W. Phosphocholine of Pneumococcal Teichoic Acids: Role in Bacterial Physiology and Pneumococcal Infection. Res. Microbiol. 2000, 151, 421–427. [Google Scholar] [CrossRef] [PubMed]
  484. Baltz, R.H.; Norris, F.H.; Matsushima, P.; DeHoff, B.S.; Rockey, P.; Porter, G.; Burgett, S.; Peery, R.; Hoskins, J.; Braverman, L.; et al. DNA Sequence Sampling of the Streptococcus Pneumoniae Genome to Identify Novel Targets for Antibiotic Development. Microb. Drug Resist. Larchmt. N 1998, 4, 1–9. [Google Scholar] [CrossRef] [PubMed]
  485. Wartha, F.; Beiter, K.; Normark, S.; Henriques-Normark, B. Neutrophil Extracellular Traps: Casting the NET over Pathogenesis. Curr. Opin. Microbiol. 2007, 10, 52–56. [Google Scholar] [CrossRef]
  486. Walter, J.; Loach, D.M.; Alqumber, M.; Rockel, C.; Hermann, C.; Pfitzenmaier, M.; Tannock, G.W. D-Alanyl Ester Depletion of Teichoic Acids in Lactobacillus Reuteri 100-23 Results in Impaired Colonization of the Mouse Gastrointestinal Tract. Environ. Microbiol. 2007, 9, 1750–1760. [Google Scholar] [CrossRef] [PubMed]
  487. Epand, R.F.; Savage, P.B.; Epand, R.M. Bacterial Lipid Composition and the Antimicrobial Efficacy of Cationic Steroid Compounds (Ceragenins). Biochim. Biophys. Acta BBA—Biomembr. 2007, 1768, 2500–2509. [Google Scholar] [CrossRef]
  488. Rashid, R.; Veleba, M.; Kline, K.A. Focal Targeting of the Bacterial Envelope by Antimicrobial Peptides. Front. Cell Dev. Biol. 2016, 4, 55. [Google Scholar] [CrossRef] [PubMed]
  489. Strahl, H.; Errington, J. Bacterial Membranes: Structure, Domains, and Function. Annu. Rev. Microbiol. 2017, 71, 519–538. [Google Scholar] [CrossRef]
  490. Short, S.A.; White, D.C. Metabolism of Phosphatidylglycerol, Lysylphosphatidylglycerol, and Cardiolipin of Staphylococcus Aureus. J. Bacteriol. 1971, 108, 219–226. [Google Scholar] [CrossRef] [PubMed]
  491. Nahaie, M.R.; Goodfellow, M.; Minnikin, D.E.; Hájek, V. Polar Lipid and Isoprenoid Quinone Composition in the Classification of Staphylococcus. J. Gen. Microbiol. 1984, 130, 2427–2437. [Google Scholar] [CrossRef]
  492. Kilelee, E.; Pokorny, A.; Yeaman, M.R.; Bayer, A.S. Lysyl-Phosphatidylglycerol Attenuates Membrane Perturbation Rather than Surface Association of the Cationic Antimicrobial Peptide 6W-RP-1 in a Model Membrane System: Implications for Daptomycin Resistance. Antimicrob. Agents Chemother. 2010, 54, 4476–4479. [Google Scholar] [CrossRef]
  493. Zhang, T.; Muraih, J.K.; Tishbi, N.; Herskowitz, J.; Victor, R.L.; Silverman, J.; Uwumarenogie, S.; Taylor, S.D.; Palmer, M.; Mintzer, E. Cardiolipin Prevents Membrane Translocation and Permeabilization by Daptomycin. J. Biol. Chem. 2014, 289, 11584–11591. [Google Scholar] [CrossRef]
  494. Bayer, A.S.; Prasad, R.; Chandra, J.; Koul, A.; Smriti, M.; Varma, A.; Skurray, R.A.; Firth, N.; Brown, M.H.; Koo, S.P.; et al. In Vitro Resistance of Staphylococcus Aureus to Thrombin-Induced Platelet Microbicidal Protein Is Associated with Alterations in Cytoplasmic Membrane Fluidity. Infect. Immun. 2000, 68, 3548–3553. [Google Scholar] [CrossRef] [PubMed]
  495. Ernst, C.M.; Staubitz, P.; Mishra, N.N.; Yang, S.-J.; Hornig, G.; Kalbacher, H.; Bayer, A.S.; Kraus, D.; Peschel, A. The Bacterial Defensin Resistance Protein MprF Consists of Separable Domains for Lipid Lysinylation and Antimicrobial Peptide Repulsion. PLoS Pathog. 2009, 5, e1000660. [Google Scholar] [CrossRef]
  496. Bayer, A.S.; Mishra, N.N.; Chen, L.; Kreiswirth, B.N.; Rubio, A.; Yang, S.-J. Frequency and Distribution of Single-Nucleotide Polymorphisms within mprF in Methicillin-Resistant Staphylococcus Aureus Clinical Isolates and Their Role in Cross-Resistance to Daptomycin and Host Defense Antimicrobial Peptides. Antimicrob. Agents Chemother. 2015, 59, 4930–4937. [Google Scholar] [CrossRef]
  497. Yang, S.-J.; Mishra, N.N.; Rubio, A.; Bayer, A.S. Causal Role of Single Nucleotide Polymorphisms within the mprF Gene of Staphylococcus Aureus in Daptomycin Resistance. Antimicrob. Agents Chemother. 2013, 57, 5658–5664. [Google Scholar] [CrossRef]
  498. Rubio, A.; Conrad, M.; Haselbeck, R.J.; Kedar, G.C.; Brown-Driver, V.; Finn, J.; Silverman, J.A. Regulation of mprF by Antisense RNA Restores Daptomycin Susceptibility to Daptomycin-Resistant Isolates of Staphylococcus Aureus. Antimicrob. Agents Chemother. 2011, 55, 364–367. [Google Scholar] [CrossRef] [PubMed]
  499. Koprivnjak, T.; Peschel, A.; Gelb, M.H.; Liang, N.S.; Weiss, J.P. Role of Charge Properties of Bacterial Envelope in Bactericidal Action of Human Group IIA Phospholipase A2 against Staphylococcus Aureus. J. Biol. Chem. 2002, 277, 47636–47644. [Google Scholar] [CrossRef] [PubMed]
  500. Song, D.; Jiao, H.; Liu, Z. Phospholipid Translocation Captured in a Bifunctional Membrane Protein MprF. Nat. Commun. 2021, 12, 2927. [Google Scholar] [CrossRef] [PubMed]
  501. Rashid, R.; Nair, Z.J.; Chia, D.M.H.; Chong, K.K.L.; Cazenave Gassiot, A.; Morley, S.A.; Allen, D.K.; Chen, S.L.; Chng, S.S.; Wenk, M.R.; et al. Depleting Cationic Lipids Involved in Antimicrobial Resistance Drives Adaptive Lipid Remodeling in Enterococcus Faecalis. mBio 2023, 14, e0307322. [Google Scholar] [CrossRef] [PubMed]
  502. Boldrin, F.; Cioetto Mazzabò, L.; Lanéelle, M.-A.; Rindi, L.; Segafreddo, G.; Lemassu, A.; Etienne, G.; Conflitti, M.; Daffé, M.; Garzino Demo, A.; et al. LysX2 Is a Mycobacterium Tuberculosis Membrane Protein with an Extracytoplasmic MprF-like Domain. BMC Microbiol. 2022, 22, 85. [Google Scholar] [CrossRef] [PubMed]
  503. Melo, M.N.; Ferre, R.; Castanho, M.A.R.B. Antimicrobial Peptides: Linking Partition, Activity and High Membrane-Bound Concentrations. Nat. Rev. Microbiol. 2009, 7, 245–250. [Google Scholar] [CrossRef] [PubMed]
  504. Nikaido, H. Multidrug Efflux Pumps of Gram-Negative Bacteria. J. Bacteriol. 1996, 178, 5853–5859. [Google Scholar] [CrossRef]
  505. Paulsen, I.T.; Brown, M.H.; Skurray, R.A. Proton-Dependent Multidrug Efflux Systems. Microbiol. Rev. 1996, 60, 575–608. [Google Scholar] [CrossRef]
  506. Li, X.-Z.; Nikaido, H. Efflux-Mediated Drug Resistance in Bacteria. Drugs 2004, 64, 159–204. [Google Scholar] [CrossRef]
  507. Clemens, R.; Zaschke-Kriesche, J.; Khosa, S.; Smits, S.H.J. Insight into Two ABC Transporter Families Involved in Lantibiotic Resistance. Front. Mol. Biosci. 2017, 4, 91. [Google Scholar] [CrossRef]
  508. Feng, Y.; Pan, X.; Huang, N.; Feng, Y.; Wu, Q.; Wang, B. The human beta-defensins expression in female genital tract and pregnancy-related tissues. Sichuan Da Xue Xue Bao Yi Xue Ban 2003, 34, 217–219. [Google Scholar] [PubMed]
  509. Colclough, A.L.; Alav, I.; Whittle, E.E.; Pugh, H.L.; Darby, E.M.; Legood, S.W.; McNeil, H.E.; Blair, J.M. RND Efflux Pumps in Gram-Negative Bacteria; Regulation, Structure and Role in Antibiotic Resistance. Future Microbiol. 2020, 15, 143–157. [Google Scholar] [CrossRef] [PubMed]
  510. Gaurav, A.; Bakht, P.; Saini, M.; Pandey, S.; Pathania, R. Role of Bacterial Efflux Pumps in Antibiotic Resistance, Virulence, and Strategies to Discover Novel Efflux Pump Inhibitors. Microbiology 2023, 169, 001333. [Google Scholar] [CrossRef] [PubMed]
  511. Alav, I.; Kobylka, J.; Kuth, M.S.; Pos, K.M.; Picard, M.; Blair, J.M.A.; Bavro, V.N. Structure, Assembly, and Function of Tripartite Efflux and Type 1 Secretion Systems in Gram-Negative Bacteria. Chem. Rev. 2021, 121, 5479–5596. [Google Scholar] [CrossRef] [PubMed]
  512. Beis, K.; Rebuffat, S. Multifaceted ABC Transporters Associated to Microcin and Bacteriocin Export. Res. Microbiol. 2019, 170, 399–406. [Google Scholar] [CrossRef] [PubMed]
  513. Kobayashi, N.; Nishino, K.; Yamaguchi, A. Novel Macrolide-Specific ABC-Type Efflux Transporter in Escherichia coli. J. Bacteriol. 2001, 183, 5639–5644. [Google Scholar] [CrossRef] [PubMed]
  514. Mason, K.M.; Bruggeman, M.E.; Munson, R.S.; Bakaletz, L.O. The Non-Typeable Haemophilus Influenzae Sap Transporter Provides a Mechanism of Antimicrobial Peptide Resistance and SapD-Dependent Potassium Acquisition. Mol. Microbiol. 2006, 62, 1357–1372. [Google Scholar] [CrossRef] [PubMed]
  515. Mason, K.M.; Raffel, F.K.; Ray, W.C.; Bakaletz, L.O. Heme Utilization by Nontypeable Haemophilus Influenzae Is Essential and Dependent on Sap Transporter Function. J. Bacteriol. 2011, 193, 2527–2535. [Google Scholar] [CrossRef]
  516. Shelton, C.L.; Raffel, F.K.; Beatty, W.L.; Johnson, S.M.; Mason, K.M. Sap Transporter Mediated Import and Subsequent Degradation of Antimicrobial Peptides in Haemophilus. PLoS Pathog. 2011, 7, e1002360. [Google Scholar] [CrossRef]
  517. Yang, S.-J.; Bayer, A.S.; Mishra, N.N.; Meehl, M.; Ledala, N.; Yeaman, M.R.; Xiong, Y.Q.; Cheung, A.L. The Staphylococcus Aureus Two-Component Regulatory System, GraRS, Senses and Confers Resistance to Selected Cationic Antimicrobial Peptides. Infect. Immun. 2012, 80, 74–81. [Google Scholar] [CrossRef]
  518. Kobylka, J.; Kuth, M.S.; Müller, R.T.; Geertsma, E.R.; Pos, K.M. AcrB: A Mean, Keen, Drug Efflux Machine. Ann. N. Y. Acad. Sci. 2020, 1459, 38–68. [Google Scholar] [CrossRef] [PubMed]
  519. Lyu, M.; Moseng, M.A.; Reimche, J.L.; Holley, C.L.; Dhulipala, V.; Su, C.-C.; Shafer, W.M.; Yu, E.W. Cryo-EM Structures of a Gonococcal Multidrug Efflux Pump Illuminate a Mechanism of Drug Recognition and Resistance. mBio 2020, 11, e00996-20. [Google Scholar] [CrossRef] [PubMed]
  520. Rieg, S.; Huth, A.; Kalbacher, H.; Kern, W.V. Resistance against Antimicrobial Peptides Is Independent of Escherichia Coli AcrAB, Pseudomonas Aeruginosa MexAB and Staphylococcus Aureus NorA Efflux Pumps. Int. J. Antimicrob. Agents 2009, 33, 174–176. [Google Scholar] [CrossRef] [PubMed]
  521. Warner, D.M.; Levy, S.B. Different Effects of Transcriptional Regulators MarA, SoxS and Rob on Susceptibility of Escherichia Coli to Cationic Antimicrobial Peptides (CAMPs): Rob-Dependent CAMP Induction of the marRAB Operon. Microbiol. Read. Engl. 2010, 156, 570–578. [Google Scholar] [CrossRef] [PubMed]
  522. Dorschner, R.A.; Lopez-Garcia, B.; Peschel, A.; Kraus, D.; Morikawa, K.; Nizet, V.; Gallo, R.L. The Mammalian Ionic Environment Dictates Microbial Susceptibility to Antimicrobial Defense Peptides. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2006, 20, 35–42. [Google Scholar] [CrossRef] [PubMed]
  523. D’Amato, R.F.; Thornsberry, C.; Baker, C.N.; Kirven, L.A. Effect of Calcium and Magnesium Ions on the Susceptibility of Pseudomonas Species to Tetracycline, Gentamicin Polymyxin B, and Carbenicillin. Antimicrob. Agents Chemother. 1975, 7, 596–600. [Google Scholar] [CrossRef] [PubMed]
  524. Smith, B.L.; Fernando, S.; King, M.D. Escherichia Coli Resistance Mechanism AcrAB-TolC Efflux Pump Interactions with Commonly Used Antibiotics: A Molecular Dynamics Study. Sci. Rep. 2024, 14, 2742. [Google Scholar] [CrossRef] [PubMed]
  525. Trampari, E.; Prischi, F.; Vargiu, A.V.; Abi-Assaf, J.; Bavro, V.N.; Webber, M.A. Functionally Distinct Mutations within AcrB Underpin Antibiotic Resistance in Different Lifestyles. npj Antimicrob. Resist. 2023, 1, 2. [Google Scholar] [CrossRef] [PubMed]
  526. Ababou, A.; Koronakis, V. Structures of Gate Loop Variants of the AcrB Drug Efflux Pump Bound by Erythromycin Substrate. PloS One 2016, 11, e0159154. [Google Scholar] [CrossRef] [PubMed]
  527. Tamer, Y.T.; Gaszek, I.; Rodrigues, M.; Coskun, F.S.; Farid, M.; Koh, A.Y.; Russ, W.; Toprak, E. The Antibiotic Efflux Protein TolC Is a Highly Evolvable Target under Colicin E1 or TLS Phage Selection. Mol. Biol. Evol. 2021, 38, 4493–4504. [Google Scholar] [CrossRef]
  528. Saw, H.T.H.; Webber, M.A.; Mushtaq, S.; Woodford, N.; Piddock, L.J.V. Inactivation or Inhibition of AcrAB-TolC Increases Resistance of Carbapenemase-Producing Enterobacteriaceae to Carbapenems. J. Antimicrob. Chemother. 2016, 71, 1510–1519. [Google Scholar] [CrossRef] [PubMed]
  529. Pantel, L.; Guérin, F.; Serri, M.; Gravey, F.; Houard, J.; Maurent, K.; Attwood, M.; Noel, A.; MacGowan, A.; Racine, E.; et al. Exploring Cluster-Dependent Antibacterial Activities and Resistance Pathways of NOSO-502 and Colistin against Enterobacter Cloacae Complex Species. Antimicrob. Agents Chemother. 2022, 66, e00776-22. [Google Scholar] [CrossRef] [PubMed]
  530. García-Romero, I.; Srivastava, M.; Monjarás-Feria, J.; Korankye, S.O.; MacDonald, L.; Scott, N.E.; Valvano, M.A. Drug Efflux and Lipid A Modification by 4-L-Aminoarabinose Are Key Mechanisms of Polymyxin B Resistance in the Sepsis Pathogen Enterobacter Bugandensis. J. Glob. Antimicrob. Resist. 2024, 37, 108–121. [Google Scholar] [CrossRef] [PubMed]
  531. Peschel, A.; Vincent Collins, L. Staphylococcal Resistance to Antimicrobial Peptides of Mammalian and Bacterial Origin. Peptides 2001, 22, 1651–1659. [Google Scholar] [CrossRef]
  532. Durell, S.R.; Guy, H.R. Structural Models of the KtrB, TrkH, and Trk1,2 Symporters Based on the Structure of the KcsA K(+) Channel. Biophys. J. 1999, 77, 789–807. [Google Scholar] [CrossRef] [PubMed]
  533. Schäfer, A.-B.; Wenzel, M. A How-To Guide for Mode of Action Analysis of Antimicrobial Peptides. Front. Cell. Infect. Microbiol. 2020, 10, 540898. [Google Scholar] [CrossRef] [PubMed]
  534. Yeaman, M.R.; Bayer, A.S.; Koo, S.P.; Foss, W.; Sullam, P.M. Platelet Microbicidal Proteins and Neutrophil Defensin Disrupt the Staphylococcus Aureus Cytoplasmic Membrane by Distinct Mechanisms of Action. J. Clin. Investig. 1998, 101, 178–187. [Google Scholar] [CrossRef]
  535. Turcot, I.; Ponnampalam, T.V.; Bouwman, C.W.; Martin, N.L. Isolation and Characterization of a Chromosomally Encoded Disulphide Oxidoreductase from Salmonella Enterica Serovar Typhimurium. Can. J. Microbiol. 2001, 47, 711–721. [Google Scholar] [CrossRef]
  536. Johns, B.E.; Purdy, K.J.; Tucker, N.P.; Maddocks, S.E. Phenotypic and Genotypic Characteristics of Small Colony Variants and Their Role in Chronic Infection. Microbiol. Insights 2015, 8, 15–23. [Google Scholar] [CrossRef]
  537. Vesga, O.; Groeschel, M.C.; Otten, M.F.; Brar, D.W.; Vann, J.M.; Proctor, R.A. Staphylococcus Aureus Small Colony Variants Are Induced by the Endothelial Cell Intracellular Milieu. J. Infect. Dis. 1996, 173, 739–742. [Google Scholar] [CrossRef]
  538. Proctor, R.A.; Peters, G. Small Colony Variants in Staphylococcal Infections: Diagnostic and Therapeutic Implications. Clin. Infect. Dis. 1998, 27, 419–422. [Google Scholar] [CrossRef] [PubMed]
  539. Proctor, R.A.; Balwit, J.M.; Vesga, O. Variant Subpopulations of Staphylococcus Aureus as Cause of Persistent and Recurrent Infections. Infect. Agents Dis. 1994, 3, 302–312. [Google Scholar] [PubMed]
  540. Samuelsen, Ø.; Haukland, H.H.; Kahl, B.C.; von Eiff, C.; Proctor, R.A.; Ulvatne, H.; Sandvik, K.; Vorland, L.H. Staphylococcus Aureus Small Colony Variants Are Resistant to the Antimicrobial Peptide Lactoferricin B. J. Antimicrob. Chemother. 2005, 56, 1126–1129. [Google Scholar] [CrossRef] [PubMed]
  541. Gläser, R.; Becker, K.; von Eiff, C.; Meyer-Hoffert, U.; Harder, J. Decreased Susceptibility of Staphylococcus Aureus Small-Colony Variants toward Human Antimicrobial Peptides. J. Invest. Dermatol. 2014, 134, 2347–2350. [Google Scholar] [CrossRef]
  542. de la Fuente-Núñez, C.; Korolik, V.; Bains, M.; Nguyen, U.; Breidenstein, E.B.M.; Horsman, S.; Lewenza, S.; Burrows, L.; Hancock, R.E.W. Inhibition of Bacterial Biofilm Formation and Swarming Motility by a Small Synthetic Cationic Peptide. Antimicrob. Agents Chemother. 2012, 56, 2696–2704. [Google Scholar] [CrossRef] [PubMed]
  543. Thöming, J.G.; Häussler, S. Pseudomonas Aeruginosa Is More Tolerant Under Biofilm Than Under Planktonic Growth Conditions: A Multi-Isolate Survey. Front. Cell. Infect. Microbiol. 2022, 12, 851784. [Google Scholar] [CrossRef] [PubMed]
  544. Vestby, L.K.; Grønseth, T.; Simm, R.; Nesse, L.L. Bacterial Biofilm and Its Role in the Pathogenesis of Disease. Antibiotics 2020, 9, 59. [Google Scholar] [CrossRef]
  545. George, J.; Halami, P.M. Presence of Extracellular DNA & Protein in Biofilm Formation by Gentamicin-Resistant Lactobacillus Plantarum. Indian J. Med. Res. 2019, 149, 257–262. [Google Scholar] [CrossRef] [PubMed]
  546. Singh, S.; Datta, S.; Narayanan, K.B.; Rajnish, K.N. Bacterial Exo-Polysaccharides in Biofilms: Role in Antimicrobial Resistance and Treatments. J. Genet. Eng. Biotechnol. 2021, 19, 140. [Google Scholar] [CrossRef]
  547. Yasir, M.; Willcox, M.D.P.; Dutta, D. Action of Antimicrobial Peptides against Bacterial Biofilms. Materials 2018, 11, 2468. [Google Scholar] [CrossRef]
  548. Wang, M.; Deng, Z.; Li, Y.; Xu, K.; Ma, Y.; Yang, S.-T.; Wang, J. Antibiofilm Property and Multiple Action of Peptide PEW300 against Pseudomonas Aeruginosa. Front. Microbiol. 2022, 13, 963292. [Google Scholar] [CrossRef] [PubMed]
  549. Friedrich, L.V.; White, R.L.; Bosso, J.A. Impact of Use of Multiple Antimicrobials on Changes in Susceptibility of Gram-Negative Aerobes. Clin. Infect. Dis. 1999, 28, 1017–1024. [Google Scholar] [CrossRef] [PubMed]
  550. Govan, J.R.; Nelson, J.W. Microbiology of Lung Infection in Cystic Fibrosis. Br. Med. Bull. 1992, 48, 912–930. [Google Scholar] [CrossRef] [PubMed]
  551. Hall-Stoodley, L.; Costerton, J.W.; Stoodley, P. Bacterial Biofilms: From the Natural Environment to Infectious Diseases. Nat. Rev. Microbiol. 2004, 2, 95–108. [Google Scholar] [CrossRef] [PubMed]
  552. Begun, J.; Gaiani, J.M.; Rohde, H.; Mack, D.; Calderwood, S.B.; Ausubel, F.M.; Sifri, C.D. Staphylococcal Biofilm Exopolysaccharide Protects against Caenorhabditis Elegans Immune Defenses. PLoS Pathog. 2007, 3, e57. [Google Scholar] [CrossRef] [PubMed]
  553. Anderson, G.G.; O’Toole, G.A. Innate and Induced Resistance Mechanisms of Bacterial Biofilms. Curr. Top. Microbiol. Immunol. 2008, 322, 85–105. [Google Scholar] [CrossRef] [PubMed]
  554. Cloete, T.E.; Westaard, D.; van Vuuren, S.J. Dynamic Response of Biofilm to Pipe Surface and Fluid Velocity. Water Sci. Technol. J. Int. Assoc. Water Pollut. Res. 2003, 47, 57–59. [Google Scholar] [CrossRef]
  555. Gillis, R.J.; White, K.G.; Choi, K.-H.; Wagner, V.E.; Schweizer, H.P.; Iglewski, B.H. Molecular Basis of Azithromycin-Resistant Pseudomonas Aeruginosa Biofilms. Antimicrob. Agents Chemother. 2005, 49, 3858–3867. [Google Scholar] [CrossRef]
  556. Kvist, M.; Hancock, V.; Klemm, P. Inactivation of Efflux Pumps Abolishes Bacterial Biofilm Formation. Appl. Environ. Microbiol. 2008, 74, 7376–7382. [Google Scholar] [CrossRef]
  557. Folkesson, A.; Haagensen, J.A.J.; Zampaloni, C.; Sternberg, C.; Molin, S. Biofilm Induced Tolerance towards Antimicrobial Peptides. PLoS ONE 2008, 3, e1891. [Google Scholar] [CrossRef]
  558. Allesen-Holm, M.; Barken, K.B.; Yang, L.; Klausen, M.; Webb, J.S.; Kjelleberg, S.; Molin, S.; Givskov, M.; Tolker-Nielsen, T. A Characterization of DNA Release in Pseudomonas Aeruginosa Cultures and Biofilms. Mol. Microbiol. 2006, 59, 1114–1128. [Google Scholar] [CrossRef] [PubMed]
  559. Chiang, W.-C.; Nilsson, M.; Jensen, P.Ø.; Høiby, N.; Nielsen, T.E.; Givskov, M.; Tolker-Nielsen, T. Extracellular DNA Shields against Aminoglycosides in Pseudomonas Aeruginosa Biofilms. Antimicrob. Agents Chemother. 2013, 57, 2352–2361. [Google Scholar] [CrossRef]
  560. Mulcahy, H.; Charron-Mazenod, L.; Lewenza, S. Extracellular DNA Chelates Cations and Induces Antibiotic Resistance in Pseudomonas Aeruginosa Biofilms. PLoS Pathog. 2008, 4, e1000213. [Google Scholar] [CrossRef]
  561. Verstraeten, N.; Braeken, K.; Debkumari, B.; Fauvart, M.; Fransaer, J.; Vermant, J.; Michiels, J. Living on a Surface: Swarming and Biofilm Formation. Trends Microbiol. 2008, 16, 496–506. [Google Scholar] [CrossRef] [PubMed]
  562. Kim, W.; Surette, M.G. Swarming Populations of Salmonella Represent a Unique Physiological State Coupled to Multiple Mechanisms of Antibiotic Resistance. Biol. Proced. Online 2003, 5, 189–196. [Google Scholar] [CrossRef] [PubMed]
  563. Birchler, T.; Seibl, R.; Büchner, K.; Loeliger, S.; Seger, R.; Hossle, J.P.; Aguzzi, A.; Lauener, R.P. Human Toll-like Receptor 2 Mediates Induction of the Antimicrobial Peptide Human Beta-Defensin 2 in Response to Bacterial Lipoprotein. Eur. J. Immunol. 2001, 31, 3131–3137. [Google Scholar] [CrossRef]
  564. Tomasinsig, L.; Scocchi, M.; Di Loreto, C.; Artico, D.; Zanetti, M. Inducible Expression of an Antimicrobial Peptide of the Innate Immunity in Polymorphonuclear Leukocytes. J. Leukoc. Biol. 2002, 72, 1003–1010. [Google Scholar] [CrossRef]
  565. Lauth, X.; Babon, J.J.; Stannard, J.A.; Singh, S.; Nizet, V.; Carlberg, J.M.; Ostland, V.E.; Pennington, M.W.; Norton, R.S.; Westerman, M.E. Bass Hepcidin Synthesis, Solution Structure, Antimicrobial Activities and Synergism, and in Vivo Hepatic Response to Bacterial Infections. J. Biol. Chem. 2005, 280, 9272–9282. [Google Scholar] [CrossRef]
  566. Lindmark, H.; Johansson, K.C.; Stöven, S.; Hultmark, D.; Engström, Y.; Söderhäll, K. Enteric Bacteria Counteract Lipopolysaccharide Induction of Antimicrobial Peptide Genes. J. Immunol. 2001, 167, 6920–6923. [Google Scholar] [CrossRef]
  567. Sperandio, B.; Regnault, B.; Guo, J.; Zhang, Z.; Stanley, S.L.; Sansonetti, P.J.; Pédron, T. Virulent Shigella Flexneri Subverts the Host Innate Immune Response through Manipulation of Antimicrobial Peptide Gene Expression. J. Exp. Med. 2008, 205, 1121–1132. [Google Scholar] [CrossRef]
  568. Hamon, M.A.; Batsché, E.; Régnault, B.; Tham, T.N.; Seveau, S.; Muchardt, C.; Cossart, P. Histone Modifications Induced by a Family of Bacterial Toxins. Proc. Natl. Acad. Sci. USA 2007, 104, 13467–13472. [Google Scholar] [CrossRef] [PubMed]
  569. Chen, H.; Wen, Y.; Li, Z. Clear Victory for Chlamydia: The Subversion of Host Innate Immunity. Front. Microbiol. 2019, 10, 1412. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Advantages of AMPs over traditional antibiotics. This figure illustrates the differences between ATBs and AMPs regarding their target, resistance mechanisms, spectrum of activity, and cytotoxicity.
Figure 1. Advantages of AMPs over traditional antibiotics. This figure illustrates the differences between ATBs and AMPs regarding their target, resistance mechanisms, spectrum of activity, and cytotoxicity.
Microorganisms 12 01259 g001
Figure 2. Classification of antimicrobial peptides. AMPs are classified according to seven criteria. The classification of AMPs based on their three-dimensional structure encompasses four categories: alpha (α), beta (β), alpha–beta (α-β), and non-α-β. These categories are defined by the presence of helical structures, β strands, a combination of helical and β structures, or the absence of both α and β structures, respectively. Regarding covalent bond classification, peptides are divided into four groups based on their polypeptide chain-bonding patterns: Class O peptides adopt circular structures due to a bond between the N-terminal and C-terminal backbone atoms; class P peptides take on a shape resembling the letter “P”, formed by a bond between the side chain of one amino acid and the backbone of another; class S peptides exhibit bonds between different side chains; and class L includes all linear peptides [11].
Figure 2. Classification of antimicrobial peptides. AMPs are classified according to seven criteria. The classification of AMPs based on their three-dimensional structure encompasses four categories: alpha (α), beta (β), alpha–beta (α-β), and non-α-β. These categories are defined by the presence of helical structures, β strands, a combination of helical and β structures, or the absence of both α and β structures, respectively. Regarding covalent bond classification, peptides are divided into four groups based on their polypeptide chain-bonding patterns: Class O peptides adopt circular structures due to a bond between the N-terminal and C-terminal backbone atoms; class P peptides take on a shape resembling the letter “P”, formed by a bond between the side chain of one amino acid and the backbone of another; class S peptides exhibit bonds between different side chains; and class L includes all linear peptides [11].
Microorganisms 12 01259 g002
Figure 3. Modifications of LPS for enhanced resistance against AMPs. The position of a substituent or branch is identified by the number of the carbon atom it is bonded to.
Figure 3. Modifications of LPS for enhanced resistance against AMPs. The position of a substituent or branch is identified by the number of the carbon atom it is bonded to.
Microorganisms 12 01259 g003
Figure 4. Modifications of teichoic acids, lipoteichoic acids, and peptidoglycans associated with AMP resistance. Enzymes involved in the modification reaction are mentioned between parentheses.
Figure 4. Modifications of teichoic acids, lipoteichoic acids, and peptidoglycans associated with AMP resistance. Enzymes involved in the modification reaction are mentioned between parentheses.
Microorganisms 12 01259 g004
Figure 5. Bacterial mechanisms of AMP resistance not associated with modifications of cell wall structures. (A): Active efflux. (B): Secreted proteinases degrading AMPs. (C): Surface proteins trapping/inhibiting AMPs. (D): Secreted proteins blocking/inhibiting AMPs. (E): Pili-mediated blocking of AMPs. (F): Capsule-mediated protection. (G): Modifications of cytoplasmic membrane phospholipids by the incorporation of cationic residues. (H): Transcriptional downregulation of AMP production by mucosal host cells.
Figure 5. Bacterial mechanisms of AMP resistance not associated with modifications of cell wall structures. (A): Active efflux. (B): Secreted proteinases degrading AMPs. (C): Surface proteins trapping/inhibiting AMPs. (D): Secreted proteins blocking/inhibiting AMPs. (E): Pili-mediated blocking of AMPs. (F): Capsule-mediated protection. (G): Modifications of cytoplasmic membrane phospholipids by the incorporation of cationic residues. (H): Transcriptional downregulation of AMP production by mucosal host cells.
Microorganisms 12 01259 g005
Table 1. Selected antimicrobial peptides with bacterial membrane depolarization/disruption actions under clinical trials or approved for clinical application.
Table 1. Selected antimicrobial peptides with bacterial membrane depolarization/disruption actions under clinical trials or approved for clinical application.
AMPsStructure/ChargeSourceAmino Acid Sequence aTarget Bacteria/IndicationsClinical Phase bAdministrationRefs.
Polymyxin BCyclic/+5Bacillus polymyxa6-mo-DabTDabDab[γDablLDabDabT] (Polymyxin B1)MDR G-infectionsIVTopical, oral, IV,
ophthalmic, aerosolized
[101,102]
ColistinCyclic/+5Bacillus colistinus6-mh-DabTDab[γDablLDabDabT] (Colistin A)MDR G-infectionsIVTopical, oral, IV[103,104,105]
NisinCyclic/+4Lactococcus lactisITSISLCTPGCKTGALMGCNMKTATCHCSIHVSK (Nisin A)Broad spectrum, food preservativeHigh clinical
potential
Undefined; oral or
IP in animal models
[106,107,108]
DaptomycinCyclic/-3Streptomyces roseosporusWNDTGKDADGSEYG+ skin infections, endocarditis, & bacteremiaIVIV[109,110]
Gramicidin SCyclic/+2Brevibacillus brevisVKLFPVKLFPBroad spectrum; wound infections,
conjunctivitis, genital ulcers
In marketOphthalmic and topical preparations[111,112]
LTX-109 (Lytixar)Cyclic/+3SyntheticR-Tbt-R-NH-EtPhG+ skin infections, anti-MRSA and VRSAII/IIITopical or nasal[113,114]
Murepavadin (POL708)Cyclic/+5SyntheticAla-Ser-D-Pro-Pro-Thr-Trp-Ile-Dab-Orn-D-Dab-Dab-Trp-Dab-DabPseudomonas in cystic fibrosisIIIIV and eFlow®
nebulizer system
[115,116]
LL-37 (hCAP18)α-helical/+6HumanLLGDFFRKSKEKIGKEFKRIVQRIKDFLRNLVPRTESHard-to-heal venous leg ulcersIIWound bed preparations[117,118,119]
Pexiganan (MSI-78)α-helical/+9SyntheticGIGKFLKKAKKFGKAFVKILKKInfected diabetic foot ulcersIIITopical[119,120,121]
Melimine Random coil c/+15 SyntheticTLISWIKNKRKQRPRVSRRRRRRGGRRRRContact lens colonizers, anti-biofilmII/IIIOcular[122]
Omiganan
(CLS001 or MBI-226)
Linear c/+5SyntheticILRWPWWPWRRKSkin and catheter infections, antisepsisIIITopical[123]
hLF1-11Linear c/+4SyntheticGRRRSVQWCAVMDR A. baumannii MRSA, Lm, E. coli, and KpI/IIIV[124]
Brilacidin
(PMX-30063)
Linear/+4 SyntheticNot available since it is a non-peptide arylamide oligomerStaphylococcus spp. skin infectionsIITopical and mouth rinse[125]
Abbreviations: hLF1-11: Human lactoferrin 1–11; IV: Intravenous, IP: Intraperitoneal; G-: Gram-negative; G+: Gram-positive; Kp: Klebsiella pneumoniae; Lm: Listeria monocytogenes; MRSA: Methicillin-resistant S. aureus; MDR: multidrug resistant; VRSA: Vancomycin-resistant S. aureus. a Available at http://dramp.cpu-bioinfor.org (accessed on 14 May 2024); b Available at https://www.clinicaltrials.gov/ (accessed on 14 May 2024); c May increase their α-helical content/change conformation in bacterial membrane-mimetic environments. The underlined AMPs are FDA-approved.
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

Tajer, L.; Paillart, J.-C.; Dib, H.; Sabatier, J.-M.; Fajloun, Z.; Abi Khattar, Z. Molecular Mechanisms of Bacterial Resistance to Antimicrobial Peptides in the Modern Era: An Updated Review. Microorganisms 2024, 12, 1259. https://doi.org/10.3390/microorganisms12071259

AMA Style

Tajer L, Paillart J-C, Dib H, Sabatier J-M, Fajloun Z, Abi Khattar Z. Molecular Mechanisms of Bacterial Resistance to Antimicrobial Peptides in the Modern Era: An Updated Review. Microorganisms. 2024; 12(7):1259. https://doi.org/10.3390/microorganisms12071259

Chicago/Turabian Style

Tajer, Layla, Jean-Christophe Paillart, Hanna Dib, Jean-Marc Sabatier, Ziad Fajloun, and Ziad Abi Khattar. 2024. "Molecular Mechanisms of Bacterial Resistance to Antimicrobial Peptides in the Modern Era: An Updated Review" Microorganisms 12, no. 7: 1259. https://doi.org/10.3390/microorganisms12071259

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