Next Article in Journal
Effect of Fermentation by Probiotic Bacteria on the Bioaccessibility of Bioactive Compounds from the Fruit of the Juçara Palm (Euterpe edulis Martius)
Next Article in Special Issue
Advancements and Future Prospects in Hypocrellins Production and Modification for Photodynamic Therapy
Previous Article in Journal
Nutrient Consumption Patterns of Saccharomyces cerevisiae and Their Application in Fruit Wine Fermentation
Previous Article in Special Issue
Microbial Fermentation and Therapeutic Potential of p-Cymene: Insights into Biosynthesis and Antimicrobial Bioactivity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Antimicrobial Peptides: Structure, Mechanism of Action, and Molecular Optimization Strategies

1
Key Laboratory of Industrial Biotechnology, Ministry of Education, School of Biotechnology, Jiangnan University, Wuxi 214122, China
2
Yixing Institute of Food and Biotechnology Co., Ltd., Yixing 214200, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Fermentation 2024, 10(11), 540; https://doi.org/10.3390/fermentation10110540
Submission received: 26 September 2024 / Revised: 21 October 2024 / Accepted: 21 October 2024 / Published: 23 October 2024

Abstract

:
Antimicrobial peptides (AMPs) are bioactive macromolecules that exhibit antibacterial, antiviral, and immunomodulatory functions. They come from a wide range of sources and are found in all forms of life, from bacteria to plants, vertebrates, and invertebrates, and play an important role in controlling the spread of pathogens, promoting wound healing and treating tumors. Consequently, AMPs have emerged as promising alternatives to next-generation antibiotics. With advancements in systems biology and synthetic biology technologies, it has become possible to synthesize AMPs artificially. We can better understand their functional activities for further modification and development by investigating the mechanism of action underlying their antimicrobial properties. This review focuses on the structural aspects of AMPs while highlighting their significance for biological activity. Furthermore, it elucidates the membrane targeting mechanism and intracellular targets of these peptides while summarizing molecular modification approaches aimed at enhancing their antibacterial efficacy. Finally, this article outlines future challenges in the functional development of AMPs along with proposed strategies to overcome them.

1. Introduction

The easy accessibility and widespread utilization of conventional antibiotics have resulted in their excessive use, leading to the emergence of antimicrobial resistance in healthcare systems worldwide. Therefore, there is an urgent need to develop novel antibiotics or antibiotic alternatives. AMPs represent a class of oligopeptides that exhibit resistance against bacteria and viruses [1]. Typically composed of 12–50 amino acid residues [2], these peptides can be derived from various sources, including animals, plants, and other origins [3]. AMPs play a crucial role in the innate immune system as they are the primary defense against pathogenic infections [4]. Distinguishing themselves from traditional antibiotics, AMPs possess the ability to target multiple pathogens, thereby minimizing the formation of drug-resistant strains while exhibiting broad-spectrum antibacterial properties, so they can be widely used in food, medicine, and agriculture, as shown in Figure 1 [5]. Consequently, they hold significant potential as alternative therapeutic agents for combating bacterial infections in the future [6,7,8]. Numerous clinical cases have already demonstrated the efficacy of AMPs in treating pathogenic bacterial transmission, wound healing, and tumors [9]. However, natural AMPs are limited by origin, cost implications, and biosafety concerns; these limitations hinder their development for therapeutic purposes [10,11]. To overcome these constraints and enhance stability, activity levels, and targeting specificity while reducing cytotoxicity toward cells, it is essential to gain further insights into the structural characteristics and mechanisms underlying the action of AMPs. This review provides an extensive analysis of the structure and mechanism of AMPs, focusing mainly on molecular modification strategies aimed at enhancing antibacterial activity. These efforts aim to offer new prospects for future clinical development and application.

2. Structure and Characteristics of AMPs

AMPs are distinguished by their conserved physicochemical properties, despite the variation observed in their primary peptide sequences. Contrary to expectations, certain specific amino acids at crucial positions show high conservation across diverse AMPs, with lysine and arginine particularly abundant in most AMPs. The efficacy of these peptides does not depend on any single attribute alone; rather, it is a result of the precise and intricate integration of numerous characteristics.

2.1. Sequence and Size

Most AMPs contain fewer than 100 amino acids, and the composition and distribution of amino acids within these peptides vary across different species [12]. The length of AMPs significantly influences their antibacterial activity and specificity against microbial targets. Shorter membrane-active AMPs (e.g., Aurein 1.2 or Citropin 1.1) are expected to disrupt membranes through micellization, while longer membrane-active AMPs (e.g., Caerin) may adopt extended conformations to stabilize transmembrane pores. The antifungal activity of a series of synthetic peptides (KWn-NH2), consisting of varying lysine and tryptophan repeats, is dependent on peptide length. Peptide activity increases with increasing peptide length, with only the most extended KW5 peptide exhibiting cytotoxicity against human keratinocyte cell lines.

2.2. Charge

Most AMPs possess a net positive charge ranging from +1 to +9 and can form well-defined cationic domains [13]. Electrostatic forces primarily drive their initial interaction with the cell membrane. There is a strong correlation between the cationic charge and biological activity, where an increase in charge leads to enhanced activity and a higher concentration of bound peptides at the membrane interface [14]. However, there generally exists an optimal charge threshold; surpassing this threshold typically results in decreased antimicrobial peptide activity. For example, analogs of Magainin-2 demonstrate a dependence on the charge, with activity increasing as charges rise while hydrophobicity and structure remain unchanged. Moreover, exceeding the threshold point through additional cationic residues significantly reduces peptide selectivity toward microbial targets while promoting hemolysis [15]. Removing the cationic C-terminal tail from melittin analogs substantially reduces their hemolytic activity and binding affinity for cell membranes [16].

2.3. Hydrophobicity

Hydrophobicity plays a critical role in the interaction with cell membranes, as it regulates the peptide’s ability to partition into the membrane layer. AMPs typically contain approximately 50% hydrophobic residues [17]. Like charge, optimizing the percentage of hydrophobicity enhances activity against cell membranes. However, exceeding this optimal point leads to a decline in antimicrobial efficacy and an increase in cytotoxicity toward mammalian cells [18]. For instance, studies utilizing Magainin-2 analogs have demonstrated that even slight increases in hydrophobicity significantly enhance the ability to cleave zwitterion/anionic mixtures compared to purely anionic PG liposomes, highlighting the substantial impact of hydrophobicity on membrane interactions [19].

2.4. Amphipathic

The amphipathic nature of AMPs is primarily determined by two specific amino acid residues within the protein [19]. One type consists of hydrophilic residues, such as glycine, histidine, arginine, and aspartic acid. These hydrophilic residues are the main constituents of AMPs and contribute to their stability in the environment [19]. Additionally, they can bind with organic substances on cell membrane proteins to enhance adhesion to bacterial cells [20]. Another type is hydrophobic residues, such as alanine, leucine, and isoleucine. These residues are mainly located at the C-terminus of antimicrobial peptide proteins. Due to their hydrophobic properties, they interact with the hydrophobic region of bacterial cell membranes to exert their effects [21]. AMPs containing a combination of hydrophilic and hydrophobic amino acid residues play a crucial role in penetrating and sterilizing bacterial membranes. For instance, Magainin-2 possesses both residues: a hydrophilic amino acid at its N-terminus and a hydrophobic one at its C-terminus. This structural arrangement enables it to penetrate bacterial membranes through lipophilic sites effectively, disrupting them for antibacterial and bactericidal effects [22].

2.5. Secondary Structure

One of the most important features of AMPs is the diversity of their secondary structures, which are essential for their ability to interact with and disrupt microbial membranes, including α-helix, β-sheet structures, and αβ and non-αβ conformations [23]. They are not fixed and unchanging but can undergo dynamic changes when in contact with the cell membrane of microorganisms such as bacteria. This change enables AMPs to effectively interact with hydrophobic regions of the cell membrane, thereby disrupting membrane integrity [23]. The secondary structure of AMPs usually also has amphiphilicity, which means they have both hydrophilic and hydrophobic parts. This property allows them to form channels or pores on the cell membrane, leading to microbial death. In addition, the diversity and plasticity of this structure are also key factors for AMPs to resist various types of microorganisms. Table 1 shows the structures of AMPs from different sources will their mechanism of action.
α-helical AMPs have been subjected to intense investigation, owing to their prevalence, extensive occurrence across different species, and potent activity against a wide array of microbial threats, as shown in Figure 2a [24]. Nature is replete with α-helical AMPs, such as Cecropin, Pleurocidin, melittin, Magainin, and Moricin, which are sourced from a diverse range of organisms including aquatic vegetation, insects, amphibians, teleosts, and mammals [25,26]. Studies have shown that the α-helical configuration of these peptides is predominantly modulated by their engagement with the target membrane [27]. This interaction induces conformational recombination leading to the separation of hydrophilic and hydrophobic amino acid residues within the peptide, which is crucial for developing amphiphilic structures that are favorable for membrane rupture [28]. As alpha helices gradually accumulate within the lipid bilayer, their hydrophobic nuclei bind to the lipid matrix, while their hydrophilic surfaces align internally, ultimately forming membrane pores [29].
β-sheet AMPs are constructed from at least two β-strands, with numerous linear constructs adopting a β-hairpin-like conformation, as shown in Figure 2b [30]. Members of this group commonly feature well-preserved cysteine residues and disulfide bridges that are instrumental in defining their structure and function [31]. These disulfide bonds contribute to the enhanced stability of AMPs, decreasing their vulnerability to proteolytic degradation [32]. The antimicrobial efficacy of these peptides is typically ascribed to the cationic residues and hydrophobic side chains presented on the antiparallel β-sheets. Notable examples include protegrin-1 (PG-1) [33], thanatin [34], tachyplesin [35], polyphemusinI [36], and gomesin [37]. Defensins represent the predominant class within the β-sheet AMP family, further subdivided into subgroups according to the placement of the disulfide bond that forms.
The secondary structure of αβ AMPs is characterized by the presence of both α-helices and β-sheets, endowing them with a pronounced affinity for cellular membranes [23]. Plant and insect defensins, in particular, are distinguished by their antifungal capabilities, which arise from interactions with fungal membrane sphingolipids or microsomal membranes, as shown in Figure 2c [38]. For instance, the antifungal plant-derived peptide pea defensin 1 (Psd1) disrupts cyclin F in Neurospora crassa, interfering with the fungal cell cycle [39]. RsAFP2, a defensin isolated from radish, engages with glucosylceramide in yeast and fungi, initiating signaling pathways linked to reactive oxygen species in Candida albicans, which ultimately leads to cell death [40]. Moreover, the plant defensin Nad1 binds to membrane phosphatidylinositol 4,5-bisphosphate prior to interacting with intracellular targets, leading to the accumulation of reactive oxygen species [41]. These interactions highlight the intricate mechanisms by which αβ AMPs exert their antifungal effects, underscoring the role of their secondary structures in targeting and disrupting microbial membranes.
AMPs that deviate from the conventional αβ-structure classification are known as non-αβAMPs, encompassing extended or cyclic peptides. These unique molecules do not possess the typical α-helix or β-sheet conformations and are divisible into three groups: those enriched in tryptophan, proline, or glycine, as shown in Figure 2d [42]. Tryptophan-enriched peptides frequently adopt an amphipathic shape, exemplified by indolicidin, which boasts a tryptophan-centric core essential for its binding affinity [43]. Studies involving the interaction of indocyanine with dodecyl phosphocholine micelles have elucidated the engagement of aromatic rings from Trp6 and Trp9 with Pro7 and Pro10, respectively [44]. Similarly, the B2 domain of lactoferrin undergoes a transition to a non-αβ structure with an irregular backbone when engaging with SDS micelles [45]. Proline-rich Insect KingdomAMPs, consisting of 15 to 39 amino acid residues, are notable for their intracellular targeting capabilities [46]. Emerging evidence suggests that these peptides assume non-αβ conformations, thereby obstructing the ribosomal tunneling process by inhibiting the entry of aminoacyl tRNA into the Amide site [47]. On the other hand, glycine-rich peptides are prevalent in the insect kingdom and typically exhibit molecular weights between 8 and 30 kDa. A notable example is hymenoptaecin, a glycine-rich antimicrobial peptide derived from Hymenoptera with a typical αβ antimicrobial peptide structure that significantly inhibits both Gram-negative and Gram-positive bacteria [48].
Table 1. Examples of structures of AMPs and their mechanisms of action.
Table 1. Examples of structures of AMPs and their mechanisms of action.
AMPOriginLengthSecondary StructureMechanismRef.
MagaininsAmphibian27α-helical AMPsForms toroidal pores in lipid bilayers[49]
LL-37Human37α-helical AMPsPore-forming carpet model[50]
β-defensin2Human36Three antiparallel β-strands and an α-helical domainBiofilm inhibition[51]
ProtegrinsPorcine18β-sheet AmpsForms octameric transmembrane pores[52]
IndolicidinBovine13Linear peptidesLyses membranes; binds DNA[53]
DrosocinDrosophila19Linear peptidesInhibits protein translation via binding to 50S or 70S ribosomal subunit[54]
RattusinbRodent62β-sheet AMPsInhibits RNA polymerase; membranolytic[55]
NisinLactococcus lactis34α-β AMPsInhibits cell wall biosynthesis via binding to lipid II[56]
PhormicinInsect40αβ insect defensinInhibits gene expression in vivo[57]
NaD1Plant105αβ plant defensinBinds to phosphatidic acid[41]
TO17Sciaenops ocellatus17α-helical AMPsInduces DNA and RNA degradation[58]
Buforin2Sphaenorhynchus lacteus21α-helical AMPsInhibits the synthesis of DNA, RNA, and proteins[59]
MBP-1Marze kernels33α-helical AMPsForms small pores in the lipid bilayer, induces membrane permeability, and disrupts potassium balance[60]
Microcin J25Escherichia coli58Lasso peptideBinds RNA polymerase and inhibits its activity[61]
Bac7Bovine35α-helical AMPsInhibits protein synthesis[62]
ApidaecinPimpla disparis20Linear peptidesParticipates in the termination process of translation[63]
Histone H2AMammal39α-helical AMPsDisrupts bacterial proton gradient and chromosome tissue[64]
PlectasinPseudoplectania nigrella95α-β AMPsInhibits the biosynthesis of peptidoglycan, thereby suppressing cell wall synthesis[65]
HNP1Human94β-sheet AmpsInteracts with the essential precursor of cell wall synthesis lipid II to inhibit bacterial cell wall synthesis[66]
Gramicidin SBacillus pumilus10Cyclic AMPsDisrupts the integrity of the inner membrane lipid bilayer[67]

3. Mechanism of Action of AMPs

The exploration of the mode of action of AMPs has always been rigorous, using various membrane models to reveal their effects. However, few studies have utilized membrane potential sensitive dyes and fluorescently labeled peptides to investigate the direct effects of these peptides on intact microbial cells. The results of these studies elucidate the main interaction modes between AMPs and cell membranes, which are typically classified into two categories: those involving membrane disruption and those leading to non-destructive membrane permeability [68]. The contemporary conceptual model of the mechanism of action of AMPs now includes not only the classical view of gradually forming pores but also considerations for structural reorganization, such as changes in molecular conformation or models based on interfacial activity [69]. Given that several proposed mechanisms may be applicable to different AMPs, it is crucial to understand the continuous mechanism steps leading to membrane rupture.
Research indicates that traditional antibiotics primarily target metabolic enzymes, which can selectively induce bacteria to develop resistance mechanisms. In contrast, AMPs employ distinct mechanisms of action and mainly rely on gradually forming pores on the cell membrane to eradicate microorganisms [70]. This makes it challenging for microorganisms to develop resistance against them. AMPs selectively inhibit bacterial growth and effectively kill bacteria [71,72]. Bacterial cell membranes are rich in negatively charged phospholipid head groups such as phospholipids glycerol, cardiolipin, or serine phospholipids. On the other hand, mammalian cell membranes consist primarily of zwitterionic phospholipids like cephalin, phospholecithin, and sphingomyelin. Additionally, cholesterol—an essential component in mammalian cell membranes—can bind to AMPs and diminish their activity [25,73]. Therefore, differences in cell membrane composition serve as the primary factor contributing to the selective eradication of bacteria by AMPs [74,75]. Although ion channels, transmembrane pores, and extensive membrane disruption ultimately lead to microbial cell lysis, evidence suggests that AMPs also possess additional intracellular target sites.

3.1. Cell Membrane Action Mechanism

The structure of antimicrobial peptide (AMP) molecules in solution is usually disordered. The initial binding between AMP and bacterial membrane is driven by the electrostatic affinity between the cationic component of AMP and the negatively charged lipid head group [68]. In particular, lysine and arginine residues show a clear tendency to bind to the phosphate portion within the lipid bilayer [76]. Bacterial cells typically maintain a significant electrochemical gradient, coupled with strong electrostatic interactions extending over longer distances, which facilitate the aggregation and membrane-bound accumulation of AMP, thereby improving its targeting efficiency [77]. Once close to the target membrane, most AMPs transition from a coil conformation to a helical structure [78]. Conversely, β-sheet AMPs tend to exhibit more excellent structural stability when solubilized. They maintain their structure through disulfide bonds or the cyclization of the peptide backbone [79]. For instance, the cyclic β-sheet AMP tachyplesin undergoes minimal structural changes when transitioning from an aqueous environment to a membrane analog [79]. Due to their stable structure in solution, these types of AMPs may retain their conformation during interactions with the membrane surface. These structures often display amphiphilic properties which have been found crucial for cleavage activity [80]. Moreover, throughout the binding and destruction processes, activity relies more on maintaining a balance between electrostatic interactions at early stages and hydrophobic interactions at subsequent stages.
After binding to the cell membrane surface, the antimicrobial peptide proceeds to the second stage known as the idle concentration stage. This stage is initiated by accumulating AMPs on the cell membrane surface [81]. Several factors can impede antimicrobial peptide concentration, including their concentration, their self-aggregation tendency, the composition and fluidity of phospholipid membranes, and their size. Another crucial factor that may hinder concentration is the bacterial cell membrane’s membrane potential (ΔΨ) [82]. Presumably, ΔΨ facilitates cationic peptides’ migration toward nonpolar membrane regions and effectively reduces energy barriers for pore formation. The activity of certain AMPs such as nisin is significantly influenced by ΔΨ; when N-terminal lysine is replaced with leucine in nisin, its voltage dependence diminishes [83].
Upon binding to cell membranes, AMPs undergo conformational changes. Studies have demonstrated that α-helical AMPs exhibit a random conformation in an aqueous solution environment but rapidly adopt a well-defined amphipathic α-helical conformation upon interaction with phospholipid bilayers [84]. However, some AMPs only undergo conformational transitions when bound to negatively charged bilayers. For instance, ergot PGLa displays an irregular conformation when bound to a lecithin- and sphingomyelin-composed membrane but gradually forms an α-helical conformation when interacting with a phosphatidic acid glycerol- and cephalin-composed membrane [85]. Some studies suggest that after binding to cell membranes, AMPs may progressively form complex structures; however, their potential for forming quaternary structures fundamentally depends on the composition and conformation of their monomers [86].

3.2. Membrane Permeabilization Mechanism

Many models of membrane permeabilization have been proposed. Considering the variations in microorganism ultrastructures, the same antimicrobial peptide can exert distinct antibacterial mechanisms on different microorganisms. Moreover, depending on factors such as growth status, tissue location, and the presence of other immune mechanisms, AMPs can employ diverse mechanisms to eliminate the same species.
The “barrel stave” model involves binding a limited number of cationic AMPs to the cell membrane surface, where they aggregate due to electrostatic forces and gradually form a pore or groove on its side. This pore acts as an ion channel and allows an inward rectifying current under the influence of an external electric field [68]. Once formed, the ion channel permits water from outside to enter the cell while also allowing cytoplasmic penetration from the exterior [87]. In severe cases, energy loss leads to cell membrane disintegration, resulting in cell death [88]. On the other hand, in the barrel stave model, transmembrane peptides tightly assemble with each other and create a rigid circular pore structure, as shown in Figure 3a [89]. Human picotin does not directly interact with membranes but instead forms a hexameric structure in solution that follows the barrel stave model.
The “Toroidal pore” model proposes that the extracellular antimicrobial peptide gradually adopts an α-helical conformation upon interaction with the charged and hydrophobic bacterial cell membrane. The displacement of polar heads of hydrophobic residues in the peptide bound to the cell membrane leads to the formation of a hydrophobic structure. This induces cracks in the domain, causing the bending and stretching of the cell membrane in a positive direction. A key distinction from the “barrel wall” model is that lipids and AMPs collaboratively form transmembrane channels, as shown in Figure 3c [90]. These cracks and stretches further compromise membrane integrity. When the ratio of AMPs to lipids reaches a critical value, the vertical orientation of AMPs within the cell membrane begins, accompanied by the self-assembly of each ring, ultimately forming a dynamic supramolecular peptide–lipid complex [91]. The formation and displacement of cyclic pores depend on this ratio. However, when there is an excessive concentration of AMPs, pore stability decreases due to electrostatic repulsion caused by positively charged side chains [92].
The “Carpet” model proposes that cationic AMPs arrange themselves in a parallel fashion on the cell membrane surface, gradually forming a structure resembling felt. Once reaching a critical threshold, the energy of the cell membrane deteriorates, leading to the loss of integrity and sharp bending followed by rupture, as shown in Figure 3b [84]. In this model, the hydrophobic portion of cationic AMPs does not insert into or progressively form grooves on the cell membrane. For instance, Cecropin Pl initially aligns parallel to the membrane surface. As time progresses and antibacterial agent concentration increases, a peptide layer with gradually increasing monomer concentration forms on the membrane surface until it eventually causes a rupture of the cell membrane [93].
The “Aggregate” model suggests that after insertion into the cell membrane, AMPs and lipids form micelle complexes that traverse through aggregates and dynamically create pores within the cell membrane. This mechanism also allows for the entry of AMPs into cells, as shown in Figure 3d [94]. Unlike in the wormhole model, there is no specific orientation for AMPs in this particular model [95].

3.3. Intracellular Mechanism of Action

Changes in membrane permeability alone cannot fully account for the potent bactericidal effects of AMPs. Inducing cell membrane permeabilization only inhibits bacterial growth, while the complete destruction of functional tissue cells significantly impacts cell death [88]. Many AMPs accumulate within tissue cells by traversing cell membranes and disrupt the normal functions of signaling cells, ultimately leading to the demise of bacteria and fungi. However, studies suggest that AMPs possess universal targets for intracellular action [96]. AMPs can bind to nucleic acids and proteins; impede replication, transcription, and translation processes; impair cellular organelles; or affect enzyme systems to disturb cell cycle progression and energy metabolism, as shown in Figure 4.
The binding mechanisms of AMPs to nucleic acids and proteins are diverse, depending on the specific AMPs and target molecules. Certain AMPs can physically interact with DNA or RNA through electrostatic interactions, hydroxyl hydrophobic interactions, etc., thereby inhibiting their replication and transcription [97]. Moreover, AMPs can also bind to protein targets such as cell membrane receptors and microbial membrane proteins, leading to structural damage and the functional impairment of the targets, ultimately compromising bacterial survival. Additionally, some AMPs can interact with nucleic acids and proteins based on their structural characteristics, including terminal charge, hydrophilicity, and hydrophobicity, to establish binding [98]. For instance, LL-37 is an antimicrobial peptide in humans with an arginine-rich region consisting of 11 amino acids at its N-terminus. This region can bind to bacterial DNA/RNA through electrostatic interaction resulting in the loss of normal replication and transcription capabilities [99]. BPI (bactericidal permeability-increasing protein), an antibacterial peptide present in bovine serum, binds to lipose sugar A on the bacterial membrane, allowing it to enter the cell, where it accelerates bacteria death with the assistance of intracellular amylase and nuclease enzymes, among others [83].
AMPs can damage bacterial mitochondria through various mechanisms. They can directly bind to ribosomes and inhibit their function: certain AMPs, upon binding to bacterial ribosomes, can hinder mRNA translation or induce premature termination, thereby impeding protein synthesis. As the concentration of AMPs increases, their affinity for bacterial ribosomes intensifies, ultimately leading to the complete disruption of ribosomal activity [100]. Additionally, AMPs can provoke an antibiotic exception response in ribosomes: specific AMPs are recognized by bacterial ribosomes, triggering an adaptive mechanism that renders them less susceptible to antibiotics [101]. For example, the antimicrobial peptide BMAP-28, an antimicrobial peptide isolated from bovine sources, is able to induce changes in mitochondrial membrane potential and release cytochrome C, thereby promoting mitochondrial outer membrane permeabilization [102].
AMPs can exert various effects on bacterial enzyme systems. Certain AMPs have the ability to directly bind to bacterial RNA polymerase, thereby inhibiting its activity and resulting in the impairment of gene transcription and protein synthesis [103]. Alternatively, AMPs can interact with extracellular adenylase in bacteria, disrupting its tertiary structure. This subsequently hampers the utilization of extracellular adenylate as an energy source, consequently affecting energy metabolism [104]. Microcin J25 is a ribosome-synthesized and post-translationally modified antimicrobial peptide that binds to the secondary channel of RNA polymerase. By obstructing the folding process of the trigger loop, it effectively prevents substrates from entering through this channel and inhibits RNA polymerase activity [105,106]. LL-37 exerts antibacterial effects on Escherichia coli by suppressing the palmitoyltransferase PagP activity responsible for repairing membrane permeability [107].
Several AMPs, including pyrrolomycin, streptomycin, and apigenin, inhibit the function of DnaK by targeting it to interfere with protein folding assisted by chaperone proteins [108]. When θ-defensins interact with bacterial membranes, they promote the secretion of cell-wall-degrading enzymes, breaking down sugar chains and peptide bonds in cell wall proteins, ultimately leading to the lysis of Staphylococcus cells [109]. Moreover, Mel4 initiates the cell death of Staphylococcus aureus by activating the release of bacterial autolysin [110]. Another type of AMP, PFR, induces necrotic cell death by inducing endoplasmic reticulum stress, increasing cytoplasmic calcium levels, and reactive oxygen species (ROS) in mitochondria [100].

3.4. Mechanism of Directed Co-Aggregationn

Directed co-aggregation is a novel antibacterial mechanism that involves the interaction between amyloidogenic AMPs (AAMPs) and bacterial target proteins, forming aggregates that cause the target proteins to lose their function. This process is also known as “protein silencing” [111]. In the directed confocal mechanism, AAMP is designed to recognize and bind to specific regions of bacterial proteins, which are typically critical areas for the formation of amyloid fibers. Through this binding, AAMP can prevent or interfere with the normal function of bacterial proteins, thereby exerting antibacterial effects. The key to this mechanism lies in the specific interaction between AAMP and bacterial proteins, which can lead to protein aggregation and prevent its participation in normal biological processes. The implementation of the directed confocal mechanism first involves selecting the target protein sequence, which typically involves analyzing the bacterial proteome to identify potential amyloid forming regions (APRs). Subsequently, AAMP is designed to bind to these APRs. This may involve the synthesis and optimization of AAMP sequences to ensure that they can effectively interact with the APRs of the target protein. Finally, the binding of AAMP to bacterial proteins leads to protein aggregation, which may be achieved through the formation of amyloid fibers or other forms of protein–protein interactions. The innovation of this mechanism lies in its provision of a different mode of action from traditional antibiotics, offering new possibilities for combating antibiotic resistance. Through the directed confocal mechanism, AAMP can specifically target bacterial proteins, reducing adverse effects on host cells and minimizing the development of drug resistance.
In summary, the mechanism of action of AMPs varies depending on their target cells’ specificity, the antimicrobial peptide concentration, and the microbial cell membranes’ biophysical properties. Moreover, when targeting the same microorganism, AMPs can employ various mechanisms to eliminate them; these include disrupting cell membrane permeability in combination with inhibiting one or multiple crucial substances’ normal functions within cells.

4. Molecular Modification Strategies for AMPs

Molecular modification technology can enhance the stability, activity, and targeting of antimicrobial peptide (AMP) engineering. Some techniques for achieving this include isomerization, biomimetic end modifications, and multimerization, as shown in Figure 5 [112,113,114]. Several factors, such as peptide length, net charge, hydrophobicity, and secondary structure influence the antibacterial activity of AMP. Alterations in peptide length can impact the antibacterial effectiveness of AMP since the peptide needs to span the lipid bilayer to stabilize pores [115]. However, changes in peptide length also result in variations in net positive charge and hydrophobicity. Increasing the positive charge of AMP can improve its binding affinity toward anionic bacterial membranes [116]. Nevertheless, highly charged peptides experience a significant reduction in biological activity under high-ionic-strength conditions. Peptide chains with hydrophobic groups have the ability to form polymers when dissolved, which allows them to insert themselves into the hydrophobic membrane core, thereby enhancing AMP stability. Additionally, hydrophobic residues improve the propensity for AMPs to adopt α-helical structures. Although amphipathicity serves as an essential structural basis for AMPs, some studies have indicated that high amphipathicity reduces their antibacterial activity while increasing hemolytic activity [117,118]. Due to the complexity associated with these factors mentioned above, no standardized solution is currently available for simultaneously optimizing various aspects of AMP engineering. However, it is possible to enhance the antibacterial activity of AMP by focusing on individual factors using different molecular engineering techniques.

4.1. Change the Number and Distribution of Net Positive Charges

The presence of basic amino acid residues, such as Arg, Lys, and His, in cationic AMPs primarily influences the electrostatic adsorption force with negatively charged phospholipids in the cell membrane [119]. Substituting Arg and Lys for amino acid residues in the α-helical region of the Tenecin1 structure can increase the overall charge and significantly enhance antibacterial activity [120]. Replacing Gln at position 39 of Rr-AFP2 with Lys can improve antibacterial activity; however, substituting Lys at position 44 with Gln decreases antibacterial activity [121]. Jiang et al. converted LV13K into a series of AMPs with net charges ranging from −5 to +10. They observed that the antibacterial activity increased over time for a range of net charges from −5 to +8 [122]. Studies have shown that AMPs containing His residues exhibit pH-dependent bactericidal activity. Under acidic conditions, protonated His residue mimics metal ion-like effects, significantly improving antibacterial activity [123]. Zhang et al. modified the net charge and distribution of positively charged residues in AR-23 antimicrobial peptide by replacing them and found that increasing the net charge affected antibacterial activity but was not correlated with the number of positively charged residues [124].

4.2. Isomerization

The conversion from L to D amino acid isomers is a strategy widely utilized to bolster the resistance of peptides against degradation by both host and microbial proteases. Enhancing the stability of peptides in serum involves several strategies, including the isomerization of AMPs, integration of non-natural amino acids, and the use of D-amino acids [125,126,127]. For example, replacing L-amino acids with D-amino acids in chicken cathelicidin-2 leads to an increase in serum stability and a decrease in cytotoxicity, while still retaining its antibacterial and LPS-neutralizing capabilities [127]. The replacement of select D-amino acid residues in the AMP W3R6 enhances its protease resistance without affecting its antibacterial efficacy. In a similar manner, a mammalian HBcARD peptide, when modified with D-amino acids, demonstrated greater stability, enhanced antibacterial activity, and reduced hemolytic effects [128]. Similarly, mammalian HBcARD peptide substituted with D-amino acids exhibited improved stability, stronger antibacterial activity, and minimal hemolytic effects [129]. Furthermore, the incorporation of disubstituted β-amino acids into the peptide sequence has been shown to not only enhance stability but also to increase lipophilicity and improve the peptide’s ability to penetrate cellular barriers [130], offering a promising route for the development of next-generation antimicrobial agents.

4.3. Change Hydrophobicity and Average Hydrophobic Moment Size

The strength of hydrophobicity is an important structural basis for determining the entry of AMPs into the phospholipid layer of the cell membrane and membrane permeability. The penetrating power is weak, but it is easy to cause the self-aggregation of AMPs, further improving the cytotoxic and hemolytic activities on the original basis [19]. Hydrophobicity exhibits a positive correlation with antibacterial activity within a certain range. Introducing hydrophobic groups can broaden the antibacterial spectrum and enhance antibacterial efficacy; however, excessive hydrophobicity may compromise cell selectivity by increasing mammalian cytotoxicity and should not exceed 50% [21]. Chen et al. demonstrated that antimicrobial peptide V13KL possesses optimal antibacterial activity with low hemolytic effects on human red blood cells. Altering its sequence to increase or decrease hydrophobicity significantly reduces its antibacterial potency, possibly due to enhanced hydrophobicity facilitating easier penetration into the hydrophobic core of red blood cell membranes [131]. The average hydrophobic moment plays a more significant role in antibacterial activity compared to helicity and hydrophobicity. The average hydrophobic moment plays a greater role in antibacterial activity than helicity and hydrophobicity. The reasonable adjustment of the average hydrophobic moment will help to further improve antibacterial activity [20]. Lee et al. substituted Gln at position 16 and Asp at position 18 in antimicrobial peptide HP with Trp, resulting in improved hydrophobicity and positively correlated enhancement in antibacterial activity [132]. Decreasing the average hydrophilic moment reduces antibacterial potency and hemolytic ability [133].

4.4. Cyclization

The cyclization of peptides has become an efficient strategy to enhance the stability and biological activity of AMPs [134]. These cyclic peptides exhibit significant affinity for bacterial membranes, attributed to the formation of β-folding structures at the membrane interface [135]. For example, Dathe and his colleagues developed a series of concise cyclic hexapeptides that showed enhanced antibacterial activity against Bacillus subtilis and Escherichia coli, surpassing their linear versions of activity [136]. To further demonstrate the benefits of cyclization, a study showed that cyclic AMPs analogs equipped with flexible linkers exhibited superior activity against Staphylococcus aureus and Pseudomonas aeruginosa compared to linear precursors [137]. In addition, the formation of intramolecular disulfide bonds between cysteine residues represents an alternative cyclization method to enhance protein hydrolysis resistance, thereby expanding the therapeutic potential of these peptides.

4.5. Change the Position of Active Amino Acid Residues and the Length of the Peptide Chain

The majority of AMPs start with a Gly amino acid residue. AMPs that are rich in Arg and Val exhibit strong antibacterial activity due to their robust electrostatic adsorption onto the lipid layer of cell membranes. Among them, an abundance of Arg facilitates membrane perforation, while an abundance of Val promotes the formation of β-sheet spatial conformation and enhances antibacterial activity. Additionally, Pro and Gly abundances play a crucial role in stabilizing the structure and antibacterial activity [46,138]. Li et al. systematically removed each C-terminal and N-terminal amino acid from antimicrobial peptide P1 individually and discovered that Pro is the primary hindrance to antibacterial activity [139]. On the other hand, Mishra et al. investigated Try-rich AMPs and observed that positional changes can impact antibacterial activity; however, quantitative changes have minimal effect on such activity [140]. Furthermore, substituting Lys for His residue within the colistin sequence diminishes its antibacterial efficacy [141].

4.6. Construction of Hybrid AMPs

Protein fusion refers to the modification method of using genetic engineering technology to fuse protein or peptide molecules with some functional proteins to produce new molecules, which is an effective means of prolonging the half-life of peptide drugs. By integrating specific residues from two to three peptides with distinct modes of action into a singular sequence, it is possible to engineer novel hybrid AMPs [142,143]. Li et al. utilized this method to generate broad-spectrum antibacterial fusion peptides (BPI21/LL-37) by fusing BPI21 and LL-37, which were then used to modify human umbilical cord mesenchymal stem cells (hUC MSCs). Research has shown that the BPI21/LL-37 fusion peptide not only has an extended half-life, but also exhibits broader antibacterial activity. In addition, this modification enhances the antibacterial and endotoxin neutralizing abilities of hUC MSCs without compromising their fundamental functions in stem cell-based tissue repair, regeneration, and anti-inflammatory responses, which is crucial for the prevention and treatment of sepsis [144].
Furthering the innovation in this field, Alzoubi’s research team designed H4, a hybrid peptide that merges alpha-helical segments from BMAP-27 and OP-145, showcasing broad-spectrum antibacterial activity and reduced toxicity [143]. Additionally, a “triple hybrid” peptide composition involving sericin-A, melittin, and LL-37 has demonstrated a significant boost in antibacterial efficacy against both Gram-positive and Gram-negative bacteria, while concurrently diminishing hemolytic activity [145]. The development of antibiotic peptide conjugates (APCs) represents another frontier in antimicrobial therapy, wherein antibiotics are linked to peptides via specific connectors. These APCs leverage the synergistic antibacterial effects of both components, while simultaneously overcoming limitations associated with standalone antibiotics or peptides. Such limitations include challenges with cell permeability, serum stability, cytotoxicity, hemolysis, and resilience under high salt concentrations [146]. This approach holds the promise of providing a novel strategy for addressing the complexities of bacterial infections and the urgent need for new antimicrobial agents.

5. Summary and Prospect

AMPs are a class of natural antimicrobial agents with broad-spectrum bioactivity and potent bactericidal effects. They play a significant role in the endogenous antimicrobial systems of organisms and are gradually emerging as potential alternatives to traditional antibiotics in the medical field. Although their clinical application is currently limited by challenges in biosynthesis and uncertainties regarding human safety, the rapid advancement of genetic and biotechnological techniques is opening up vast prospects for the application of AMPs.
Firstly, the progress of modern biotechnology has made the large-scale production of AMPs possible. The development of gene sequencing technologies allows us to isolate and screen antimicrobial peptide genes from a variety of organisms, which can then be introduced into expression hosts such as bacteria, yeast, and plants through genetic engineering techniques, enabling the mass production of AMPs. This method has already been successfully applied in the production of numerous AMPs. In this process, fermentation technology plays a crucial role. Due to the high production costs of AMPs, traditional chemical synthesis methods struggle to meet the demands of large-scale production, making microbial fermentation an ideal solution. In recent years, significant progress has been made in the application of fermentation technology in the production of AMPs. Researchers have selected microbial strains with high antimicrobial peptide production capabilities and optimized fermentation conditions, such as temperature, pH, and dissolved oxygen, to enhance the yield of AMPs. Additionally, genetic engineering techniques for genetically modifying microbial strains to improve their antimicrobial peptide synthesis capacity are also an important means of increasing fermentation efficiency. During the fermentation process, both solid-state and liquid fermentation methods are widely used, with solid-state fermentation typically employing natural materials such as grains and soybean meal as carriers, which not only yield higher antimicrobial peptide production but also lower production costs. Beyond optimizing fermentation techniques, the extraction and purification of AMPs are also critical steps. Researchers have developed a series of efficient extraction and purification methods, such as centrifugation, filtration, gel filtration chromatography, and high-performance liquid chromatography, to ensure the purity and activity of AMPs. Fermentation technology has significantly improved the production efficiency of AMPs. For example, nisin, an antimicrobial peptide produced by lactic acid bacteria fermentation, has been widely used in food preservation [147]. Furthermore, AMPs produced through fermentation technology, such as plant-derived AMPs, animal-derived antimicrobial polypeptides, and fungal-derived AMPs, show broad application prospects in the fields of pharmaceuticals, cosmetics, and feed additives.
Secondly, the specific mechanisms of AMPs can be harnessed to develop safer antimicrobial drugs. Compared with traditional antibiotics, AMPs have different killing mechanisms, making them less prone to resistance, more adaptable, and less toxic to host cells. Therefore, they are considered a safer and more reliable class of antimicrobial agents.
Lastly, modifying the structure of AMPs can yield new functionalities. Many AMPs have a wide range of applications in the field of life sciences. For instance, tachyplesin, an antimicrobial peptide, has been used in permanent pacemakers to prevent device infections [148]. Additionally, combining AMPs with other biological properties can lead to a range of new antimicrobial products, such as antimicrobial sutures and bandages.
Overall, AMPs hold tremendous potential in the fields of medicine and biology. With improvements in production methods and further research progress, it is anticipated that the importance and effectiveness of AMPs as disinfectant tools will be enhanced, which is a development that will contribute to creating healthier, safer, and more satisfying lives for individuals.

Author Contributions

X.M., Q.W., K.R., T.X., Z.Z., M.X., X.Z., and Z.R. conducted the literature review and drafted the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 32171471, No. 32071470), the Project Funded by the Priority Academic Program Development of Jiangsu Higher Education Institutions, Top-notch Academic Programs Project of Jiangsu Higher Education Institutions.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

All authors were employed by the company Yixing Institute of Food and Biotechnology Co., Ltd. The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Talapko, J.; Skrlec, I. The Principles, Mechanisms, and Benefits of Unconventional Agents in the Treatment of Biofilm Infection. Pharmaceuticals 2020, 13, 299. [Google Scholar] [CrossRef] [PubMed]
  2. Nawrot, R.; Barylski, J.; Nowicki, G.; Broniarczyk, J.; Buchwald, W. Gozdzicka-Jozefiak, Plant antimicrobial peptides. Folia Microbiol. 2014, 59, 181–196. [Google Scholar] [CrossRef] [PubMed]
  3. Zasloff, M. Antimicrobial peptides of multicellular organisms. Nature 2002, 415, 389–395. [Google Scholar] [CrossRef] [PubMed]
  4. Mendez-Samperio, P. Recent advances in the field of antimicrobial peptides in inflammatory diseases. Adv. Biomed. Res. 2013, 2, 50. [Google Scholar] [CrossRef]
  5. Bhattacharjya, S.; Mohid, S.A.; Bhunia, A. Atomic-Resolution Structures and Mode of Action of Clinically Relevant Antimicrobial Peptides. Int. J. Mol. Sci. 2022, 23, 4558. [Google Scholar] [CrossRef]
  6. Zaman, S.B.; Hussain, M.A.; Nye, R.; Mehta, V.; Mamun, K.T.; Hossain, N. A Review on Antibiotic Resistance: Alarm Bells are Ringing. Cureus 2017, 9, e1403. [Google Scholar] [CrossRef]
  7. Burnham, J.P. Climate change and antibiotic resistance: A deadly combination. Ther. Adv. Infect. Dis. 2021, 8, 7. [Google Scholar] [CrossRef]
  8. Giacomini, E.; Perrone, V.; Alessandrini, D.; Paoli, D.; Nappi, C.; Esposti, L.D. Evidence of Antibiotic Resistance from Population-Based Studies: A Narrative Review. Infect. Drug Resistance 2021, 14, 849–858. [Google Scholar] [CrossRef]
  9. Gan, B.H.; Gaynord, J.; Rowe, S.M.; Deingruber, T.; Spring, D.R. The multifaceted nature of antimicrobial peptides: Current synthetic chemistry approaches and future directions. Chem. Soc. Rev. 2021, 50, 7820–7880. [Google Scholar] [CrossRef]
  10. Larsson, D.G.J.; Flach, C.F. Antibiotic resistance in the environment. Nat. Rev. Microbiol. 2022, 20, 257–269. [Google Scholar] [CrossRef]
  11. Murray, C.J.L.; Ikuta, K.S.; Sharara, F.; Swetschinski, L.; Aguilar, G.R.; Gray, A.; Han, C.; Bisignano, C.; Rao, P.; Wool, E.; et al. Antimicrobial Resistance, Global burden of bacterial antimicrobial resistance in 2019: A systematic analysis. Lancet 2022, 399, 629–655. [Google Scholar] [CrossRef] [PubMed]
  12. Huan, Y.C.; Kong, Q.; Mou, H.J.; Yi, H.X. Antimicrobial Peptides: Classification, Design, Application and Research Progress in Multiple Fields. Front. Microbiol. 2020, 11, 21. [Google Scholar] [CrossRef] [PubMed]
  13. Bonucci, A.; Balducci, E.; Martinelli, M.; Pogni, R. Human neutrophil peptide 1 variants bearing arginine modified cationic side chains: Effects on membrane partitioning. Biophys. Chem. 2014, 190, 32–40. [Google Scholar] [CrossRef] [PubMed]
  14. Fillion, M.; Valois-Paillard, G.; Lorin, A.; Noel, M.; Voyer, N.; Auger, M. Membrane Interactions of Synthetic Peptides with Antimicrobial Potential: Effect of Electrostatic Interactions and Amphiphilicity. Probiotics Antimicrob. Proteins 2015, 7, 66–74. [Google Scholar] [CrossRef]
  15. Juba, M.L.; Porter, D.K.; Williams, E.H.; Rodriguez, C.A.; Barksdale, S.M.; Bishop, B.M. Helical cationic antimicrobial peptide length and its impact on membrane disruption. Biochim. Biophys. Acta—Biomembr. 2015, 1848, 1081–1091. [Google Scholar] [CrossRef]
  16. Hall, K.; Lee, T.H.; Aguilar, M.I. The role of electrostatic interactions in the membrane binding of melittin. J. Mol. Recognit. 2011, 24, 108–118. [Google Scholar] [CrossRef]
  17. Pasupuleti, M.; Schmidtchen, A.; Malmsten, M. Antimicrobial peptides: Key components of the innate immune system. Crit. Rev. Biotechnol. 2012, 32, 143–171. [Google Scholar] [CrossRef]
  18. Chou, H.T.; Kuo, T.Y.; Chiang, J.C.; Pei, M.J.; Yang, W.T.; Yu, H.C.; Lin, S.B.; Chen, W.J. Design and synthesis of cationic antimicrobial peptides with improved activity and selectivity against Vibrio spp. Int. J. Antimicrob. Agents 2008, 32, 130–138. [Google Scholar] [CrossRef]
  19. Hollmann, A.; Martínez, M.; Noguera, M.E.; Augusto, M.T.; Disalvo, A.; Santos, N.C.; Semorile, L.; Maffía, P.C. Role of amphipathicity and hydrophobicity in the balance between hemolysis and peptide-membrane interactions of three related antimicrobial peptides. Colloids Surf. B Biointerfaces 2016, 141, 528–536. [Google Scholar] [CrossRef]
  20. Cherry, M.A.; Higgins, S.K.; Melroy, H.; Lee, H.S.; Pokorny, A. Peptides with the same composition, hydrophobicity, and hydrophobic moment bind to phospholipid bilayers with different affinities. J. Phys. Chem. B 2014, 118, 12462–12470. [Google Scholar] [CrossRef]
  21. Jiang, Z.; Kullberg, B.J.; van der Lee, H.; Vasil, A.I.; Hale, J.D.; Mant, C.T.; Hancock, R.E.; Vasil, M.L.; Netea, M.G.; Hodges, R.S. Effects of hydrophobicity on the antifungal activity of α-helical antimicrobial peptides. Chem. Biol. Drug Des. 2008, 72, 483–495. [Google Scholar] [CrossRef] [PubMed]
  22. Eisenberg, D.; Weiss, R.M.; Terwilliger, T.C. The hydrophobic moment detects periodicity in protein hydrophobicity. Proc. Natl. Acad. Sci. USA 1984, 81, 140–144. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, G. Improved methods for classification, prediction, and design of antimicrobial peptides. Methods Mol. Biol. 2015, 1268, 43–66. [Google Scholar]
  24. Nar, H.; Schmid, A.; Puder, C.; Potterat, O. High-Resolution Crystal Structure of a Lasso Peptide. ChemMedChem 2010, 5, 1689–1692. [Google Scholar] [CrossRef] [PubMed]
  25. Hollmann, A.; Martinez, M.; Maturana, P.; Semorile, L.C.; Maffia, P.C. Antimicrobial Peptides: Interaction with Model and Biological Membranes and Synergism with Chemical Antibiotics. Front. Chem. 2018, 6, 204. [Google Scholar] [CrossRef] [PubMed]
  26. Yi, H.Y.; Chowdhury, M.; Huang, Y.D.; Yu, X.Q. Insect antimicrobial peptides and their applications. Appl. Microbiol. Biotechnol. 2014, 98, 5807–5822. [Google Scholar] [CrossRef]
  27. Lei, J.; Sun, L.C.; Huang, S.Y.; Zhu, C.H.; Li, P.; He, J.; Mackey, V.; Coy, D.H.; He, Q.Y. The antimicrobial peptides and their potential clinical applications. Am. J. Transl. Res. 2019, 11, 3919–3931. [Google Scholar]
  28. Uggerhoj, L.E.; Poulsen, T.J.; Munk, J.K.; Fredborg, M.; Sondergaard, T.E.; Frimodt-Moller, N.; Hansen, P.R.; Wimmer, R. Rational Design of Alpha-Helical Antimicrobial Peptides: Do’s and Don’ts. Chembiochem 2015, 16, 242–253. [Google Scholar] [CrossRef]
  29. Oren, Z.; Shai, Y. Mode of action of linear amphipathic α-helical antimicrobial peptides. Biopolymers 1998, 47, 451–463. [Google Scholar] [CrossRef]
  30. Bulet, P.; Stocklin, R.; Menin, L. Anti-microbial peptides: From invertebrates to vertebrates. Immunol. Rev. 2004, 198, 169–184. [Google Scholar] [CrossRef]
  31. Mohanram, H.; Bhattacharjya, S. Cysteine deleted protegrin-1 (CDP-1): Anti-bacterial activity, outer-membrane disruption and selectivity. Biochim. Biophys. Acta—Gen. Subj. 2014, 1840, 3006–3016. [Google Scholar] [CrossRef] [PubMed]
  32. Dhople, V.; Krukemeyer, A.; Ramamoorthy, A. The human beta-defensin-3, an antibacterial peptide with multiple biological functions. Biochim. Biophys. Acta—Biomembr. 2006, 1758, 1499–1512. [Google Scholar] [CrossRef] [PubMed]
  33. Rodziewicz-Motowidlo, S.; Mickiewicz, B.; Greber, K.; Sikorska, E.; Szultka, L.; Kamysz, E.; Kamysz, W. Antimicrobial and conformational studies of the active and inactive analogues of the protegrin-1 peptide. FEBS J. 2010, 277, 1010–1022. [Google Scholar] [CrossRef] [PubMed]
  34. Mandard, N.; Sodano, P.; Labbe, H.; Bonmatin, J.M.; Bulet, P.; Hetru, C.; Ptak, M.; Vovelle, F. Solution structure of thanatin, a potent bactericidal and fungicidal insect peptide, determined from proton two-dimensional nuclear magnetic resonance data. Eur. J. Biochem. 1998, 256, 404–410. [Google Scholar] [CrossRef] [PubMed]
  35. Laederach, A.; Andreotti, A.H.; Fulton, D.B. Solution and micelle-bound structures of tachyplesin I and its active aromatic linear derivatives. Biochemistry 2002, 41, 12359–12368. [Google Scholar] [CrossRef]
  36. Powers, J.P.S.; Rozek, A.; Hancock, R.E.W. Structure-activity relationships for the β-hairpin cationic antimicrobial peptide polyphemusin I. Biochim. Biophys. Acta—Proteins Proteom. 2004, 1698, 239–250. [Google Scholar] [CrossRef]
  37. Mandard, N.; Bulet, P.; Caille, A.; Daffre, S.; Vovelle, F. The solution structure of gomesin, an antimicrobial cysteine-rich peptide from the spider. Eur. J. Biochem. 2002, 269, 1190–1198. [Google Scholar] [CrossRef]
  38. Vriens, K.; Cammue, B.P.A.; Thevissen, K. Antifungal Plant Defensins: Mechanisms of Action and Production. Molecules 2014, 19, 12280–12303. [Google Scholar] [CrossRef]
  39. Lobo, D.S.; Pereira, I.B.; Fragel-Madeira, L.; Medeiros, L.N.; Cabral, L.M.; Faria, J.; Bellio, M.; Campos, R.C.; Linden, R.; Kurtenbach, E. Antifungal Pisum sativum defensin 1 interacts with Neurospora crassa cyclin F related to the cell cycle. Biochemistry 2007, 46, 987–996. [Google Scholar] [CrossRef]
  40. Aerts, A.M.; Francois, I.; Meert, E.M.K.; Li, Q.T.; Cammue, B.P.A.; Thevissen, K. The antifungal activity of RsAFP2, a plant defensin from Raphanus sativus, involves the induction of reactive oxygen species in Candida albicans. J. Mol. Microbiol. Biotechnol. 2007, 13, 243–247. [Google Scholar] [CrossRef]
  41. Hayes, B.M.E.; Bleackley, M.R.; Wiltshire, J.L.; Anderson, M.A.; Traven, A.; van der Weerden, N.L. Identification and Mechanism of Action of the Plant Defensin NaD1 as a New Member of the Antifungal Drug Arsenal against Candida albicans. Antimicrob. Agents Chemother. 2013, 57, 3667–3675. [Google Scholar] [CrossRef] [PubMed]
  42. Nguyen, L.T.; Schibli, D.J.; Vogel, H.J. Structural studies and model membrane interactions of two peptides derived from bovine lactoferricin. J. Pept. Sci. 2005, 11, 379–389. [Google Scholar] [CrossRef] [PubMed]
  43. Sitaram, N.; Nagaraj, R. Interaction of antimicrobial peptides with biological and model membranes: Structural and charge requirements for activity. Biochim. Biophys. Acta—Biomembr. 1999, 1462, 29–54. [Google Scholar] [CrossRef] [PubMed]
  44. Rozek, A.; Friedrich, C.L.; Hancock, R.E.W. Structure of the bovine antimicrobial peptide indolicidin bound to dodecylphosphocholine and sodium dodecyl sulfate micelles. Biochemistry 2000, 39, 15765–15774. [Google Scholar] [CrossRef] [PubMed]
  45. Krizsan, A.; Volke, D.; Weinert, S.; Strater, N.; Knappe, D.; Hoffmann, R. Insect-Derived Proline-Rich Antimicrobial Peptides Kill Bacteria by Inhibiting Bacterial Protein Translation at the 70S Ribosome. Angew. Chem.-Int. Ed. 2014, 53, 12236–12239. [Google Scholar] [CrossRef]
  46. WLi, Y.; Tailhades, J.; O’Brien-Simpson, N.M.; Separovic, F.; Otvos, L.; Hossain, M.A.; Wade, J.D. Proline-rich antimicrobial peptides: Potential therapeutics against antibiotic-resistant bacteria. Amino Acids 2014, 46, 2287–2294. [Google Scholar]
  47. Mardirossian, M.; Grzela, R.; Giglione, C.; Meinnel, T.; Gennaro, R.; Mergaert, P.; Scocchi, M. The Host Antimicrobial Peptide Bac7(1-35) Binds to Bacterial Ribosomal Proteins and Inhibits Protein Synthesis. Chem. Biol. 2014, 21, 1639–1647. [Google Scholar] [CrossRef]
  48. Casteels, P.; Ampe, C.; Jacobs, F.; Tempst, P. Functional and chemical characterization of Hymenoptaecin, an antibacterial polypeptide that is infection-inducible in the honeybee (Apis mellifera). J. Biol. Chem. 1993, 268, 7044–7054. [Google Scholar] [CrossRef]
  49. Glattard, E.; Salnikov, E.S.; Aisenbrey, C.; Bechinger, B. Investigations of the synergistic enhancement of antimicrobial activity in mixtures of magainin 2 and PGLa. Biophys. Chem. 2016, 210, 35–44. [Google Scholar] [CrossRef]
  50. Bucki, R.; Leszczynska, K.; Namiot, A.; Sokolowski, W. Cathelicidin LL-37: A Multitask Antimicrobial Peptide. Arch. Immunol. Ther. Exp. 2010, 58, 15–25. [Google Scholar] [CrossRef]
  51. Prasad, S.V.; Fiedoruk, K.; Daniluk, T.; Piktel, E.; Bucki, R. Expression and Function of Host Defense Peptides at Inflammation Sites. Int. J. Mol. Sci. 2020, 21, 104. [Google Scholar] [CrossRef] [PubMed]
  52. Gour, S.; Kumar, V.; Singh, A.; Gadhave, K.; Goyal, P.; Pandey, J.; Giri, R.; Yadav, J.K. Mammalian antimicrobial peptide protegrin-4 self assembles and forms amyloid-like aggregates: Assessment of its functional relevance. J. Pept. Sci. 2019, 25, e3151. [Google Scholar] [CrossRef] [PubMed]
  53. Smirnova, M.P.; Kolodkin, N.I.; Kolobov, A.A.; Afonin, V.G.; Afonina, I.V.; Stefanenko, L.I.; Shpen, V.M.; Shamova, O.V. Indolicidin analogs with broad-spectrum antimicrobial activity and low hemolytic activity. Peptides 2020, 132, 170356. [Google Scholar] [CrossRef]
  54. Cardoso, M.H.; Meneguetti, B.T.; Costa, B.O.; Buccini, D.F.; Oshiro, K.G.N.; Preza, S.L.E.; Carvalho, C.M.E.; Migliolo, L.; Franco, O.L. Non-Lytic Antibacterial Peptides That Translocate Through Bacterial Membranes to Act on Intracellular Targets. Int. J. Mol. Sci. 2019, 20, 4877. [Google Scholar] [CrossRef]
  55. Ji, S.; Yun, H.; Park, G.; Min, H.J.; Lee, C.W. Expression and characterization of recombinant rattusin, an α-defensin-related peptide with a homodimeric scaffold formed by intermolecular disulfide exchanges. Protein Expr. Purif. 2018, 147, 17–21. [Google Scholar] [CrossRef]
  56. Li, W.F.; Ma, G.X.; Zhou, X.X. Apidaecin-type peptides: Biodiversity, structure-function relationships and mode of action. Peptides 2006, 27, 2350–2359. [Google Scholar] [CrossRef]
  57. Wang, B.; Yao, Y.; Wei, P.; Song, C.; Wan, S.; Yang, S.; Zhu, G.M.; Liu, H.M. Housefly Phormicin inhibits Staphylococcus aureus and MRSA by disrupting biofilm formation and altering gene expression in vitro and in vivo. Int. J. Biol. Macromol. 2021, 167, 1424–1434. [Google Scholar] [CrossRef]
  58. He, S.-W.; Wang, G.-H.; Yue, B.; Zhou, S.; Zhang, M. TO17: A teleost antimicrobial peptide that induces degradation of bacterial nucleic acids and inhibits bacterial infection in red drum, Sciaenops ocellatus. Fish Shellfish Immunol. 2018, 72, 639–645. [Google Scholar] [CrossRef]
  59. Muñoz-Camargo, C.; Salazar, V.A.; Barrero-Guevara, L.; Camargo, S.; Mosquera, A.; Groot, H.; Boix, E. Unveiling the Multifaceted Mechanisms of Antibacterial Activity of Buforin II and Frenatin 2.3S Peptides from Skin Micro-Organs of the Orinoco Lime Treefrog (Sphaenorhynchus lacteus). Int. J. Mol. Sci. 2018, 19, 2170. [Google Scholar] [CrossRef]
  60. Gigon, L.; Müller, P.; Haenni, B.; Iacovache, I.; Barbo, M.; Gosheva, G.; Yousefi, S.; Soragni, A.; von Ballmoos, C.; Zuber, B.; et al. Membrane damage by MBP-1 is mediated by pore formation and amplified by mtDNA. Cell Rep. 2024, 43, 114084. [Google Scholar] [CrossRef]
  61. Baquero, F.; Beis, K.; Craik, D.J.; Li, Y.; Link, A.J.; Rebuffat, S.; Salomón, R.; Severinov, K.; Zirah, S.; Hegemann, J.D. The pearl jubilee of microcin J25: Thirty years of research on an exceptional lasso peptide. Nat. Prod. Rep. 2024, 41, 469–511. [Google Scholar] [CrossRef] [PubMed]
  62. Guida, F.; Benincasa, M.; Zahariev, S.; Scocchi, M.; Berti, F.; Gennaro, R.; Tossi, A. Effect of Size and N-Terminal Residue Characteristics on Bacterial Cell Penetration and Antibacterial Activity of the Proline-Rich Peptide Bac7. J. Med. Chem. 2015, 58, 1195–1204. [Google Scholar] [CrossRef] [PubMed]
  63. Casteels, P.; Romagnolo, J.; Castle, M.; Casteels-Josson, K.; Erdjument-Bromage, H.; Tempst, P. Biodiversity of apidaecin-type peptide antibiotics. Prospects of manipulating the antibacterial spectrum and combating acquired resistance. J. Biol. Chem. 1994, 269, 26107–26115. [Google Scholar] [CrossRef] [PubMed]
  64. Doolin, T.; Amir, H.M.; Duong, L.; Rosenzweig, R.; Urban, L.A.; Bosch, M.; Pol, A.; Gross, S.P.; Siryaporn, A. Mammalian histones facilitate antimicrobial synergy by disrupting the bacterial proton gradient and chromosome organization. Nat. Commun. 2020, 11, 3888. [Google Scholar] [CrossRef]
  65. Schneider, T.; Kruse, T.; Wimmer, R.; Wiedemann, I.; Sass, V.; Pag, U.; Jansen, A.; Nielsen, A.K.; Mygind, P.H.; Raventós, D.S.; et al. Plectasin, a fungal defensin, targets the bacterial cell wall precursor Lipid II. Science 2010, 328, 1168–1172. [Google Scholar] [CrossRef]
  66. de Leeuw, E.; Li, C.; Zeng, P.; Li, C.; Buin, M.D.-D.; Lu, W.Y.; Breukink, E.; Lu, W. Functional interaction of human neutrophil peptide-1 with the cell wall precursor lipid II. FEBS Lett. 2010, 584, 1543–1548. [Google Scholar] [CrossRef]
  67. Pavithrra, G.; Rajasekaran, R. Gramicidin Peptide to Combat Antibiotic Resistance: A Review. Int. J. Pept. Res. Ther. 2020, 26, 191–199. [Google Scholar] [CrossRef]
  68. Melo, M.N.; Ferre, R.; Castanho, M. OPINION Antimicrobial peptides: Linking partition, activity and high membrane-bound concentrations. Nat. Rev. Microbiol. 2009, 7, 245–250. [Google Scholar] [CrossRef]
  69. 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. Military Med. Res. 2021, 8, 25. [Google Scholar] [CrossRef]
  70. 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]
  71. Klubthawee, N.; Adisakwattana, P.; Hanpithakpong, W.; Somsri, S.; Aunpad, R. A novel, rationally designed, hybrid antimicrobial peptide, inspired by cathelicidin and aurein, exhibits membrane-active mechanisms against Pseudomonas aeruginosa. Sci. Rep. 2020, 10, 17. [Google Scholar] [CrossRef] [PubMed]
  72. Le, C.F.; Fang, C.M.; Sekaran, S.D. Intracellular Targeting Mechanisms by Antimicrobial Peptides. Antimicrob. Agents Chemother. 2017, 61, 16. [Google Scholar] [CrossRef] [PubMed]
  73. Juhasz, T.; Queme-Pena, M.; Kovago, B.; Mihaly, J.; Ricci, M.; Horvati, K.; Bosze, S.; Zsila, F. Beke-Somfai, Interplay between membrane active host defense peptides and heme modulates their assemblies and in vitro activity. Sci. Rep. 2021, 11, 15. [Google Scholar] [CrossRef]
  74. Baharin, N.H.Z.; Mokhtar, N.F.K.; Desa, M.N.M.; Gopalsamy, B.; Zaki, N.N.M.; Yuswan, M.H.; Muthanna, A.; Dzaraly, N.D.; Abbasiliasi, S.; Hashim, A.M.; et al. The characteristics and roles of antimicrobial peptides as potential treatment for antibiotic-resistant pathogens: A review. PeerJ 2021, 9, 25. [Google Scholar]
  75. Mink, C.; Strandberg, E.; Wadhwani, P.; Melo, M.N.; Reichert, J.; Wacker, I.; Castanho, M.; Ulrich, A.S. Overlapping Properties of the Short Membrane-Active Peptide BP100 with (i) Polycationic TAT and (ii) α-helical Magainin Family Peptides. Front. Cell. Infect. Microbiol. 2021, 11, 14. [Google Scholar] [CrossRef]
  76. Mavri, J.; Vogel, H.J. Ion pair formation of phosphorylated amino acids and lysine and arginine side. Proteins Struct. Funct. Bioinform. 1996, 24, 495–501. [Google Scholar] [CrossRef]
  77. Yeaman, M.R.; Yount, N.Y. Mechanisms of antimicrobial peptide action and resistance. Pharmacol. Rev. 2003, 55, 27–55. [Google Scholar] [CrossRef]
  78. Dathe, M.; Wieprecht, T. Structural features of helical antimicrobial peptides: Their potential to modulate activity on model membranes and biological cells. Biochim. Biophys. Acta—Biomembr. 1999, 1462, 71–87. [Google Scholar] [CrossRef]
  79. Oishi, O.; Yamashita, S.; Nishimoto, E.; Lee, S.; Sugihara, G.; Ohno, M. Conformations and orientations of aromatic amino acid residues of tachyplesin I in phospholipid membranes. Biochemistry 1997, 36, 4352–4359. [Google Scholar] [CrossRef]
  80. Blondelle, S.E.; Houghten, R.A. Design of model amphipathic peptides having potent antimicrobial activities. Biochemistry 1992, 31, 12688–12694. [Google Scholar] [CrossRef]
  81. Hall, K.; Lee, T.H.; Daly, N.L.; Craik, D.J.; Aguilar, M.I. Gly(6) of kalata B1 is critical for the selective binding to phosphatidylethanolamine membranes. Biochim. Biophys. Acta—Biomembr. 2012, 1818, 2354–2361. [Google Scholar] [CrossRef] [PubMed]
  82. Holthuis, J.C.M.; Menon, A.K. Lipid landscapes and pipelines in membrane homeostasis. Nature 2014, 510, 48–57. [Google Scholar] [CrossRef] [PubMed]
  83. Ho, S.; Pothoulakis, C.; Koon, H.W. Antimicrobial Peptides and Colitis. Curr. Pharm. Des. 2013, 19, 40–47. [Google Scholar] [PubMed]
  84. Lee, T.H.; Hall, K.N.; Aguilar, M.I. Antimicrobial Peptide Structure and Mechanism of Action: A Focus on the Role of Membrane Structure. Curr. Top. Med. Chem. 2016, 16, 25–39. [Google Scholar] [CrossRef]
  85. Navarre, W.W.; Schneewind, O. Surface proteins of gram-positive bacteria and mechanisms of their targeting to the cell wall envelope. Microbiol. Mol. Biol. Rev. 1999, 63, 174–229. [Google Scholar] [CrossRef]
  86. Silva, J.P.; Appelberg, R.; Gama, F.M. Antimicrobial peptides as novel anti-tuberculosis therapeutics. Biotechnol. Adv. 2016, 34, 924–940. [Google Scholar] [CrossRef]
  87. Tornesello, A.L.; Borrelli, A.; Buonaguro, L.; Buonaguro, F.M.; Tornesello, M.L. Antimicrobial Peptides as Anticancer Agents: Functional Properties and Biological Activities. Molecules 2020, 25, 2850. [Google Scholar] [CrossRef]
  88. Brogden, K.A. Antimicrobial peptides: Pore formers or metabolic inhibitors in bacteria? Nat. Rev. Microbiol. 2005, 3, 238–250. [Google Scholar] [CrossRef]
  89. Sur, S.; Romo, T.D.; Grossfield, A. Selectivity and Mechanism of Fengycin, an Antimicrobial Lipopeptide, from Molecular Dynamics. J. Phys. Chem. B 2018, 122, 2219–2226. [Google Scholar] [CrossRef]
  90. Ciumac, D.; Gong, H.N.; Hu, X.Z.; Lu, J.R. Membrane targeting cationic antimicrobial peptides. J. Colloid Interface Sci. 2019, 537, 163–185. [Google Scholar] [CrossRef]
  91. 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] [PubMed]
  92. Das, P.; Sercu, T.; Wadhawan, K.; Padhi, I.; Gehrmann, S.; Cipcigan, F.; Chenthamarakshan, V.; Strobelt, H.; Santos, C.D.; Chen, P.Y.; et al. Accelerated antimicrobial discovery via deep generative models and molecular dynamics simulations. Nat. Biomed. Eng. 2021, 5, 613–623. [Google Scholar] [CrossRef] [PubMed]
  93. Moore, A.J.; Beazley, W.D.; Bibby, M.C.; Devine, D.A. Antimicrobial activity of cecropins. J. Antimicrob. Chemother. 1996, 37, 1077–1089. [Google Scholar] [CrossRef] [PubMed]
  94. Hale, J.D.; Hancock, R.E. Alternative mechanisms of action of cationic antimicrobial peptides on bacteria. Expert Rev. Anti-Infect. Ther. 2007, 5, 951–959. [Google Scholar] [CrossRef]
  95. Hancock, R.E.W.; Patrzykat, A. Clinical development of cationic antimicrobial peptides: From natural to novel antibiotics. Curr. Drug Targets. Infect. Disord. 2002, 2, 79–83. [Google Scholar] [CrossRef]
  96. Ulvatne, H.; Samuelsen, O.; Haukland, H.H.; Kramer, M.; Vorland, L.H. Lactoferricin B inhibits bacterial macromolecular synthesis in Escherichia coli and Bacillus subtilis. FEMS Microbiol. Lett. 2004, 237, 377–384. [Google Scholar] [CrossRef]
  97. Wu, C.Y.; Biswas, S.; De Gonzalo, C.V.G.; van der Donk, W.A. Investigations into the Mechanism of Action of Sublancin. ACS Infect. Dis. 2019, 5, 454–459. [Google Scholar] [CrossRef]
  98. Kang, H.K.; Seo, C.H.; Luchian, T.; Park, Y. Pse-T2, an Antimicrobial Peptide with High-Level, Broad-Spectrum Antimicrobial Potency and Skin Biocompatibility against Multidrug-Resistant Pseudomonas aeruginosa Infection. Antimicrob. Agents Chemother. 2018, 62, 10–1128. [Google Scholar] [CrossRef]
  99. Aidoukovitch, A.; Dahl, S.; Falt, F.; Nebel, D.; Svensson, D.; Tufvesson, E.; Nilsson, B.O. Antimicrobial peptide LL-37 and its pro-form, hCAP18, in desquamated epithelial cells of human whole saliva. Eur. J. Oral Sci. 2020, 128, 1–6. [Google Scholar] [CrossRef]
  100. Lv, Y.D.; Shao, G.; Zhang, Q.Y.; Wang, X.; Meng, Y.M.; Wang, L.F.; Huang, F.Y.; Yang, T.X.; Jin, Y.T.; Fu, C.Y. The antimicrobial peptide PFR induces necroptosis mediated by ER stress and elevated cytoplasmic calcium and mitochondrial ROS levels: Cooperation with Ara-C to act against acute myeloid leukemia. Signal Transduct. Target. Ther. 2019, 4, 38. [Google Scholar] [CrossRef]
  101. Kurpe, S.R.; Grishin, S.Y.; Surin, A.K.; Selivanova, O.M.; Fadeev, R.S.; Dzhus, U.F.; Gorbunova, E.Y.; Mustaeva, L.G.; Azev, V.N.; Galzitskaya, O.V. Antimicrobial and Amyloidogenic Activity of Peptides Synthesized on the Basis of the Ribosomal S1 Protein from Thermus Thermophilus. Int. J. Mol. Sci. 2020, 21, 6382. [Google Scholar] [CrossRef] [PubMed]
  102. Guo, Y.; Xun, M.; Han, J. A bovine myeloid antimicrobial peptide (BMAP-28) and its analogs kill pan-drug-resistant Acinetobacter baumannii by interacting with outer membrane protein A (OmpA). Medicine 2018, 97, e12832. [Google Scholar] [CrossRef] [PubMed]
  103. Bechinger, B.; Gorr, S.U. Antimicrobial Peptides: Mechanisms of Action and Resistance. J. Dent. Res. 2017, 96, 254–260. [Google Scholar] [CrossRef]
  104. Tang, S.S.; Prodhan, Z.H.; Biswas, S.K.; Le, C.F.; Sekaran, S.D. Antimicrobial peptides from different plant sources: Isolation, characterisation, and purification. Phytochemistry 2018, 154, 94–105. [Google Scholar] [CrossRef]
  105. Zhu, C.; Tan, H.L.; Cheng, T.; Shen, H.; Shao, J.J.; Guo, Y.Y.; Shi, S.F.; Zhang, X.L. Human β-defensin 3 inhibits antibiotic-resistant Staphylococcus biofilm formation. J. Surg. Res. 2013, 183, 204–213. [Google Scholar] [CrossRef]
  106. Braffman, N.R.; Piscotta, F.J.; Hauver, J.; Campbell, E.A.; Link, A.J.; Darst, S.A. Structural mechanism of transcription inhibition by lasso peptides microcin J25 and capistruin. Proc. Natl. Acad. Sci. USA 2019, 116, 1273–1278. [Google Scholar] [CrossRef]
  107. Yang, H.; Fu, J.Y.; Zhao, Y.Y.; Shi, H.P.; Hu, H.; Wang, H.L. Escherichia coli PagP Enzyme-Based De Novo Design and In Vitro Activity of Antibacterial Peptide LL-37. Med. Sci. Monit. 2017, 23, 2558–2564. [Google Scholar] [CrossRef]
  108. Kragol, G.; Lovas, S.; Varadi, G.; Condie, B.A.; Hoffmann, R.; Otvos, L. The antibacterial peptide pyrrhocoricin inhibits the ATPase actions of DnaK and prevents chaperone-assisted protein folding. Biochemistry 2001, 40, 3016–3026. [Google Scholar] [CrossRef]
  109. Wilmes, M.; Stockem, M.; Bierbaum, G.; Schlag, M.; Gotz, F.; Tran, D.Q.; Schaal, J.B.; Ouellette, A.J.; Selsted, M.E.; Sahl, H.-G. Killing of staphylococci by theta-defensins involves membrane impairment and activation of autolytic enzymes. Antibiotics 2014, 3, 617–631. [Google Scholar] [CrossRef]
  110. Yasir, M.; Dutta, D.; Willcox, M.D.P. Mode of action of the antimicrobial peptide Mel4 is independent of Staphylococcus aureus cell membrane permeability. PLoS ONE 2019, 14, e0215703. [Google Scholar] [CrossRef]
  111. Galzitskaya, O.V.; Kurpe, S.R.; Panfilov, A.V.; Glyakina, A.V.; Grishin, S.Y.; Kochetov, A.P.; Deryusheva, E.I.; Machulin, A.V.; Kravchenko, S.V.; Domnin, P.A.; et al. Amyloidogenic Peptides: New Class of Antimicrobial Peptides with the Novel Mechanism of Activity. Int. J. Mol. Sci. 2022, 23, 5463. [Google Scholar] [CrossRef] [PubMed]
  112. Li, W.Y.; Separovic, F.; O’Brien-Simpson, N.M.; Wade, J.D. Chemically modified and conjugated antimicrobial peptides against superbugs. Chem. Soc. Rev. 2021, 50, 4932–4973. [Google Scholar] [CrossRef] [PubMed]
  113. Malins, L.R. Peptide modification and cyclization via transition-metal catalysis. Curr. Opin. Chem. Biol. 2018, 46, 25–32. [Google Scholar] [CrossRef] [PubMed]
  114. Lei, M.; Jayaraman, A.; Van Deventer, J.A.; Lee, K. Engineering Selectively Targeting Antimicrobial Peptides. Annu. Rev. Biomed. Eng. 2021, 23, 339–357. [Google Scholar] [CrossRef] [PubMed]
  115. Schmidtchen, A.; Pasupuleti, M.; Malmsten, M. Effect of hydrophobic modifications in antimicrobial peptides. Adv. Colloid Interface Sci. 2014, 205, 265–274. [Google Scholar] [CrossRef]
  116. Porter, S.L.; Coulter, S.M.; Pentlavalli, S.; Thompson, T.P. Laverty, Self-assembling diphenylalanine peptide nanotubes selectively eradicate bacterial biofilm infection. Acta Biomater. 2018, 77, 96–105. [Google Scholar] [CrossRef]
  117. Pedron, C.N.; de Oliveira, C.S.; da Silva, A.F.; Andrade, G.P.; da Silva Pinhal, M.A.; Cerchiaro, G.; da Silva Junior, P.I.; da Silva, F.D.; Torresd, M.D.T.; Oliveira, V.X. The effect of lysine substitutions in the biological activities of the scorpion venom peptide VmCT1. Eur. J. Pharm. Sci. 2019, 136, 104952. [Google Scholar] [CrossRef]
  118. Torres, M.D.T.; Pedron, C.N.; Higashikuni, Y.; Kramer, R.M.; Cardoso, M.H.; Oshiro, K.G.N.; Franco, O.L.; Junior, P.I.S.; Silva, F.D.; Junior, V.X.O.; et al. Structure-function-guided exploration of the antimicrobial peptide polybia-CP identifies activity determinants and generates synthetic therapeutic candidates. Commun. Biol. 2018, 1, 221. [Google Scholar] [CrossRef]
  119. Gagnon, M.-C.; Strandberg, E.; Grau-Campistany, A.; Wadhwani, P.; Reichert, J.; Bürck, J.; Rabanal, F.; Auger, M.; Paquin, J.-F.; Ulrich, A.S. Influence of the Length and Charge on the Activity of α-Helical Amphipathic Antimicrobial Peptides. Biochemistry 2017, 56, 1680–1695. [Google Scholar] [CrossRef]
  120. HAhn, S.; Cho, W.; Kang, S.H.; Ko, S.S.; Park, M.S.; Cho, H.; Lee, K.H. Design and synthesis of novel antimicrobial peptides on the basis of α helical domain of Tenecin 1, an insect defensin protein, and structure-activity relationship study. Peptides 2006, 27, 640–648. [Google Scholar]
  121. Li, S.; Wang, Y.; Xue, Z.; Jia, Y.; Li, R.; He, C.; Chen, H. The structure-mechanism relationship and mode of actions of antimicrobial peptides: A review. Trends Food Sci. Technol. 2021, 109, 103–115. [Google Scholar] [CrossRef]
  122. Jiang, Z.; Vasil, A.I.; Hale, J.; Hancock, R.E.; Vasil, M.L.; Hodges, R.S. Effects of net charge and the number of positively charged residues on the biological activity of amphipathic α-helical cationic antimicrobial peptides. Adv. Exp. Med. Biol. 2009, 611, 561–562. [Google Scholar] [PubMed]
  123. Kacprzyk, L.; Rydengård, V.; Mörgelin, M.; Davoudi, M.; Pasupuleti, M.; Malmsten, M.; Schmidtchen, A. Antimicrobial activity of histidine-rich peptides is dependent on acidic conditions. Biochim. Biophys. Acta 2007, 1768, 2667–2680. [Google Scholar] [CrossRef] [PubMed]
  124. Zhang, S.K.; Song, J.W.; Gong, F.; Li, S.B.; Chang, H.Y.; Xie, H.M.; Gao, H.W.; Tan, Y.X.; Ji, S.P. Design of an α-helical antimicrobial peptide with improved cell-selective and potent anti-biofilm activity. Sci. Rep. 2016, 6, 27394. [Google Scholar] [CrossRef] [PubMed]
  125. Franco, O.L. Peptide promiscuity: An evolutionary concept for plant defense. FEBS Lett. 2011, 585, 995–1000. [Google Scholar] [CrossRef]
  126. Molhoek, E.M.; van Dijk, A.; Veldhuizen, E.J.A.; Haagsman, H.P.; Bikker, F.J. Improved proteolytic stability of chicken cathelicidin-2 derived peptides by D-amino acid substitutions and cyclization. Peptides 2011, 32, 875–880. [Google Scholar] [CrossRef]
  127. Hilchie, A.L.; Haney, E.F.; Pinto, D.M.; Hancock, R.E.W.; Hoskin, D.W. Enhanced killing of breast cancer cells by a D-amino acid analog of the winter flounder-derived pleurocidin NRC-03. Exp. Mol. Pathol. 2015, 99, 426–434. [Google Scholar] [CrossRef]
  128. Li, Y.; Liu, T.; Liu, Y.; Tan, Z.; Ju, Y.; Yang, Y.; Dong, W.B. Antimicrobial activity, membrane interaction and stability of the D-amino acid substituted analogs of antimicrobial peptide W3R6. J. Photochem. Photobiol. B—Biol. 2019, 200, 111645. [Google Scholar] [CrossRef]
  129. Chen, H.L.; Su, P.Y.; Shih, C.H. Improvement of in vivo antimicrobial activity of HBcARD peptides by D-arginine replacement. Appl. Microbiol. Biotechnol. 2016, 100, 9125–9132. [Google Scholar] [CrossRef]
  130. Zhao, L.L.; Cao, Z.X.; Bian, Y.Q.; Hu, G.D.; Wang, J.H.; Zhou, Y.Q. Molecular Dynamics Simulations of Human Antimicrobial Peptide LL-37 in Model POPC and POPG Lipid Bilayers. Int. J. Mol. Sci. 2018, 19, 1186. [Google Scholar] [CrossRef]
  131. Chen, Y.X.; Guarnieri, M.T.; Vasil, A.I.; Vasil, M.L.; Mant, C.T.; Hodges, R.S. Role of peptide hydrophobicity in the mechanism of action of α-helical antimicrobial peptides. Antimicrob. Agents Chemother. 2007, 51, 1398–1406. [Google Scholar] [CrossRef] [PubMed]
  132. Lee, K.H.; Lee, D.G.; Park, Y.; Kang, D.I.; Shin, S.Y.; Hahm, K.S.; Kim, Y. Interactions between the plasma membrane and the antimicrobial peptide HP (2-20) and its analogues derived from Helicobacter pylori. Biochem. J. 2006, 394, 105–114. [Google Scholar] [CrossRef] [PubMed]
  133. Pathak, N.; Salasauvert, R.; Ruche, G.; Janna, M.H.; McCarthy, D.; Harrison, R.G. Comparison of the effects of hydrophobicity, amphiphilicity, and α-helicity on the activities of antimicrobial peptides. Proteins 1995, 22, 182–186. [Google Scholar] [CrossRef] [PubMed]
  134. Gunasekera, S.; Muhammad, T.; Stromstedt, A.A.; Rosengren, K.J.; Goransson, U. Backbone Cyclization and Dimerization of LL-37-Derived Peptides Enhance Antimicrobial Activity and Proteolytic Stability. Front. Microbiol. 2020, 11, 168. [Google Scholar] [CrossRef]
  135. Ting, D.S.J.; Beuerman, R.W.; Dua, H.S.; Lakshminarayanan, R.; Mohammed, I. Strategies in Translating the Therapeutic Potentials of Host Defense Peptides. Front. Immunol. 2020, 11, 983. [Google Scholar] [CrossRef]
  136. Dathe, M.; Nikolenko, H.; Klose, J.; Bienert, M. Cyclization increases the antimicrobial activity and selectivity of arginine- and tryptophan-containing hexapeptides. Biochemistry 2004, 43, 9140–9150. [Google Scholar] [CrossRef]
  137. Thomsen, T.T.; Mendel, H.C.; Al-Mansour, W.; Oddo, A.; Lobner-Olesen, A.; Hansen, P.R. Analogues of a Cyclic Antimicrobial Peptide with a Flexible Linker Show Promising Activity against Pseudomonas aeruginosa and Staphylococcus aureus. Antibiotics 2020, 9, 366. [Google Scholar] [CrossRef]
  138. Edwards-Gayle, C.J.C.; Barrett, G.; Roy, S.; Castelletto, V.; Seitsonen, J.; Ruokolainen, J.; Hamley, I.W. Selective Antibacterial Activity and Lipid Membrane Interactions of Arginine-Rich Amphiphilic Peptides. ACS Appl. Bio Mater. 2020, 3, 1165–1175. [Google Scholar] [CrossRef]
  139. Lee, D.G.; Hahm, K.S.; Shin, S.Y. Structure and fungicidal activity of a synthetic antimicrobial peptide, P18, and its truncated peptides. Biotechnol. Lett. 2004, 26, 337–341. [Google Scholar] [CrossRef]
  140. Mishra, A.K.; Choi, J.; Moon, E.; Baek, K.H. Tryptophan-Rich and Proline-Rich Antimicrobial Peptides. Molecules 2018, 23, 815. [Google Scholar] [CrossRef]
  141. Cantisani, M.; Finamore, E.; Mignogna, E.; Falanga, A.; Nicoletti, G.F.; Pedone, C.; Morelli, G.; Leone, M.; Galdiero, M.; Galdiero, S. Structural insights into and activity analysis of the antimicrobial peptide myxinidin. Antimicrob. Agents Chemother. 2014, 58, 5280–5290. [Google Scholar] [CrossRef] [PubMed]
  142. Almaaytah, A.; Qaoud, M.T.; Abualhaijaa, A.; Al-Balas, Q.; Alzoubi, K.H. Hybridization and antibiotic synergism as a tool for reducing the cytotoxicity of antimicrobial peptides. Infect. Drug Resist. 2018, 11, 835–847. [Google Scholar] [CrossRef] [PubMed]
  143. Wang, C.; Yang, C.; Chen, Y.-C.; Ma, L.; Huang, K. Rational Design of Hybrid Peptides: A Novel Drug Design Approach. Curr. Med. Sci. 2019, 39, 349–355. [Google Scholar] [CrossRef]
  144. Li, Z.; Song, Y.; Yuan, P.; Guo, W.; Hu, X.; Xing, W.; Ao, L.; Tan, Y.; Wu, X.; Ao, X.; et al. Antibacterial Fusion Protein BPI21/LL-37 Modification Enhances the Therapeutic Efficacy of hUC-MSCs in Sepsis. Mol. Ther. 2020, 28, 1806–1817. [Google Scholar] [CrossRef]
  145. Fernandez, D.I.; Lee, T.H.; Sani, M.A.; Aguilar, M.I.; Separovic, F. Proline Facilitates Membrane Insertion of the Antimicrobial Peptide Maculatin 1.1 via Surface Indentation and Subsequent Lipid Disordering. Biophys. J. 2013, 104, 1495–1507. [Google Scholar] [CrossRef]
  146. David, A.A.; Park, S.E.; Parang, K.; Tiwari, R.K. Antibiotics-Peptide Conjugates Against Multidrug-Resistant Bacterial Pathogens. Curr. Top. Med. Chem. 2018, 18, 1926–1936. [Google Scholar] [CrossRef]
  147. Małaczewska, J.; Kaczorek-Łukowska, E. Nisin—A lantibiotic with immunomodulatory properties: A review. Peptides 2021, 137, 170479. [Google Scholar] [CrossRef]
  148. Riool, M.; de Breij, A.; Drijfhout, J.W.; Nibbering, P.H.; Zaat, S.A.J. Antimicrobial Peptides in Biomedical Device Manufacturing. Front. Chem. 2017, 5, 63. [Google Scholar] [CrossRef]
Figure 1. The source and application of AMPs.
Figure 1. The source and application of AMPs.
Fermentation 10 00540 g001
Figure 2. Secondary structures of several representative AMPs. (a) α-helix antimicrobial peptide; (b) β-sheet AMPs; (c) αβ antibacterial peptides; (d) non-αβ antibacterial peptides.
Figure 2. Secondary structures of several representative AMPs. (a) α-helix antimicrobial peptide; (b) β-sheet AMPs; (c) αβ antibacterial peptides; (d) non-αβ antibacterial peptides.
Fermentation 10 00540 g002
Figure 3. Membrane permeabilization mechanism of AMPs. (a) Barrel stave model; (b) Carpet model; (c) Toroidal pore model; (d) Aggregate model.
Figure 3. Membrane permeabilization mechanism of AMPs. (a) Barrel stave model; (b) Carpet model; (c) Toroidal pore model; (d) Aggregate model.
Fermentation 10 00540 g003
Figure 4. The intracellular mechanisms of action of AMPs.
Figure 4. The intracellular mechanisms of action of AMPs.
Fermentation 10 00540 g004
Figure 5. Molecular modification strategies for antibacterial activity of AMPs.
Figure 5. Molecular modification strategies for antibacterial activity of AMPs.
Fermentation 10 00540 g005
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

Ma, X.; Wang, Q.; Ren, K.; Xu, T.; Zhang, Z.; Xu, M.; Rao, Z.; Zhang, X. A Review of Antimicrobial Peptides: Structure, Mechanism of Action, and Molecular Optimization Strategies. Fermentation 2024, 10, 540. https://doi.org/10.3390/fermentation10110540

AMA Style

Ma X, Wang Q, Ren K, Xu T, Zhang Z, Xu M, Rao Z, Zhang X. A Review of Antimicrobial Peptides: Structure, Mechanism of Action, and Molecular Optimization Strategies. Fermentation. 2024; 10(11):540. https://doi.org/10.3390/fermentation10110540

Chicago/Turabian Style

Ma, Xu, Qiang Wang, Kexin Ren, Tongtong Xu, Zigang Zhang, Meijuan Xu, Zhiming Rao, and Xian Zhang. 2024. "A Review of Antimicrobial Peptides: Structure, Mechanism of Action, and Molecular Optimization Strategies" Fermentation 10, no. 11: 540. https://doi.org/10.3390/fermentation10110540

APA Style

Ma, X., Wang, Q., Ren, K., Xu, T., Zhang, Z., Xu, M., Rao, Z., & Zhang, X. (2024). A Review of Antimicrobial Peptides: Structure, Mechanism of Action, and Molecular Optimization Strategies. Fermentation, 10(11), 540. https://doi.org/10.3390/fermentation10110540

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