Next Article in Journal
Assessing the Load, Virulence and Antibiotic-Resistant Traits of ESBL/Ampc E. coli from Broilers Raised on Conventional, Antibiotic-Free, and Organic Farms
Next Article in Special Issue
Teleost Piscidins—In Silico Perspective of Natural Peptide Antibiotics from Marine Sources
Previous Article in Journal
Study on Optimizing Novel Antimicrobial Peptides with Bifunctional Activity to Prevent and Treat Peri-Implant Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

An Overview of the Potentialities of Antimicrobial Peptides Derived from Natural Sources

by
Irene Dini
1,*,
Margherita-Gabriella De Biasi
1,* and
Andrea Mancusi
2
1
Department of Pharmacy, University of Naples Federico II, Via Domenico Montesano 49, 80131 Napoli, Italy
2
Department of Food Microbiology, Istituto Zooprofilattico Sperimentale del Mezzogiorno, Via Salute 2, 80055 Portici, Italy
*
Authors to whom correspondence should be addressed.
Antibiotics 2022, 11(11), 1483; https://doi.org/10.3390/antibiotics11111483
Submission received: 4 October 2022 / Revised: 20 October 2022 / Accepted: 21 October 2022 / Published: 26 October 2022
(This article belongs to the Special Issue Potential of Antimicrobial Peptides for an Exciting Future)

Abstract

:
Antimicrobial peptides (AMPs) are constituents of the innate immune system in every kind of living organism. They can act by disrupting the microbial membrane or without affecting membrane stability. Interest in these small peptides stems from the fear of antibiotics and the emergence of microorganisms resistant to antibiotics. Through membrane or metabolic disruption, they defend an organism against invading bacteria, viruses, protozoa, and fungi. High efficacy and specificity, low drug interaction and toxicity, thermostability, solubility in water, and biological diversity suggest their applications in food, medicine, agriculture, animal husbandry, and aquaculture. Nanocarriers can be used to protect, deliver, and improve their bioavailability effectiveness. High cost of production could limit their use. This review summarizes the natural sources, structures, modes of action, and applications of microbial peptides in the food and pharmaceutical industries. Any restrictions on AMPs’ large-scale production are also taken into consideration.

1. Introduction

Antimicrobial peptides (AMPs) are the oldest known innate immune defense molecules. They are abundant in plants, arthropods, microorganisms, and animals [1]. Eukaryotes and prokaryotes synthesize AMPs in ribosomes, fungi, and bacteria, turning them into cytosol [2]. AMPs can have broad-spectrum or specific activity against pathogenic bacteria (both Gram-positive and Gram-negative), viruses, fungi, and other parasites [3]. AMPs differ in length and composition of amino acids [4]. Defensins, puroindolines, snakins, cyclotides, glycine-rich proteins, hevein, α-hairpin, knottin, and lipid transfer proteins are some natural classes of AMPs [5]. Their activity is bound by helical structure, charge, hydrophobicity, and amphipathicity [4]. The food industry employs AMPs as biopreservants and in food packaging (alone or with other antimicrobials and essential oils) to improve product shelf-life [6]. Antimicrobial peptides are considered potential drugs for treating infections caused by microorganisms that are untreatable with antibiotics on the market today [7,8]. They can reduce the development of antimicrobial resistance, affecting multiple low-affinity targets [9]. Some AMPs are subjected to peptide engineering and mutagenesis to make compounds with improved bioactivity and reduced cytotoxicity [10,11]. This review offers an overview of structures, sources, modes of action, and applications of AMPs in the food and pharmaceutical fields.

2. Antimicrobial Peptides’ Natural Source

Antimicrobial peptides are made by lower and higher organisms responding to pathogenic challenges [12]. AMPs kill the invading pathogens and modulate the innate immune response. They are commonly classified according to their sources, amino-acid-rich species, structural characteristics, and activities [13]. In multicellular organisms and humans, they are localized into specific sites commonly exposed to microbes (i.e., mucosa epithelia and skin) [13] (Figure 1).

2.1. Viral AMPs

Some phage proteins, including lysins, depolymerases, virion-associated peptidoglycan hydrolases (VAPGHs), and holins, show antibacterial activity [14]. They are defined as “enzybiotics” to indicate their use as antibacterial materials as alternatives to standard antibiotics [15]. The two types of phage AMPs are known as phage-encoded lytic factors and phage-tail complexes [16].
Phage lysines (size range from 25 to 40 kDa) are peptidoglycan-hydrolyzing enzymes [17], which can hydrolyze the microbial cell wall, permitting bacteriophage progeny release [16]. Lysins have rapid bactericidal activity (against Gram-positive and Gram-negative bacteria) and other desirable characteristics, such as synergy with cell-wall-reducing antibiotics, anti-biofilm action, heat stability up to ~50 °C, and the possibility of lyophilization [18,19,20]. Peptidoglycan hydrolases (VAPGHs), encoded mainly by double-stranded DNA phages, have high thermal stability. They infect Gram-positive and Gram-negative bacteria. VAPGHs have a C-terminal cell-wall-binding domain, which can link them to receptors on the bacterial cell surface. They inject genetic materials into bacterial cells after partially and locally damaging bacterial cell wall peptidoglycans [21]. They can be classified into three categories: glycosidases that cut glycosidic bonds in the peptidoglycan chain, amidases that cut amide bonds (between N-acetylmuramic acid lactyl and stem peptide l-alanines), and endopeptidases that cleave peptide bonds within either the stem peptide or cross-link [22].
Phage-tail-like AMPs are high-molecular-weight cylindrical peptides with a structure like a phage tail [23,24]. They can be classified into two classes: R-type (related to Myoviridae phage tails) and F-type (related to Siphoviridae phage tails) [23].
R-type phage-tail-like bacteriocins are nonflexible and have tubes surrounded by contractile sheaths [25]. They initially make a channel in the cell membrane and successively drive their internal core into the cell. This process determines rapid cell death by decoupling cellular ion gradients [23], interfering with oxygen uptake, and affecting macromolecule synthesis [26].
F-type phage-tail-like bacteriocins are flexible and noncontractile [25]. They act similarly to R-type bacteriocins [27].

2.2. Bacterial AMPs

2.2.1. AMPs Made by Gram-Positive Bacteria

Gram-positive bacteria can produce AMPs in ribosomes (ribosomal AMPs) or enzymatically (non-ribosomal AMPs) [28,29] (Figure 2). Twenty-sixty amino acids (hydrophobic and cationic) can make up ribosomally synthesized bacterial AMPs (bacteriocins) [30].
Bacteriocins can be classified into bacteriocins produced by Gram-positive and Gram-negative bacteria [31].

2.2.2. AMPs Made by Gram-Positive Bacteria

Gram-positive organisms make bacteriocins that can be grouped into lantibiotics (class I), non-lantibiotics (class II), large-sized bacteriocins (class III), and uniquely structured bacteriocins (class IV) [32] (Figure 2).
Lantibiotics are active against primarily Gram-positive bacteria [32]. They are small peptides (<5 kDa; 19–50 amino acids) that are stable to heat, pH, and proteolysis [33]. Lantibiotics can be subdivided into subclasses Ia and Ib (Figure 2).
Subclass Ia lantibiotics form pores in bacterial membranes that determine cellular death [34].
Subclass Ib lantibiotics are inflexible peptides that decrease the activity of bacteria crucial enzymes [32].
Class II AMPs (non-lanthionine-containing bacteriocins) are small (<10 kDa) and heat-stable peptides that can form pores in the bacterial membrane. They can be grouped into four subclasses [35].
Subclass IIa consists of disulfide linear peptides with similar amino acid sequences that permeabilize the cell membrane, showing significant antilisterial activity [36].
Subclass IIb bacteriocins increase the permeability of the bacterial cell membrane to specific small molecules [37]. They contain two peptide subunits (α and β) [37].
Subclass IIc bacteriocins permeabilize the microbial membrane, dissipate the membrane potential, and cause cell death [38]. They comprise small, cyclic peptides whose C- and N-terminals are covalently linked [39].
Subclass IId comprises the remaining non-characterized bacteriocins in class II [32].
Class III bacteriocins (bacteriolysins) [35] are large (>30 kDa), heat-labile peptides [32] (Figure 2).
Class IV AMPs containing lipids or carbohydrates are susceptible to lipolytic and glycolytic enzymes [40] (Figure 2).
Non-ribosomally synthesized AMPs are made from peptide synthetases produced by Gram-positive and Gram-negative bacteria [9].
Bacteriocins can decrease food spoilage [31].

2.2.3. AMPs Made by Gram-Negative Bacteria

Gram-negative organisms make bacteriocins that can be grouped into microcins, colicins, colicin-like bacteriocins, and phage-tail-like bacteriocins [41] (Figure 3).
Microcins are made by Enterobacteriaceae. Microcins interact with some cellular targets. They can format pores that determine membrane disruption [23] or decrease the functionality of enzymes (the ATP synthase complex, DNA gyrase, RNA polymerase, and aspartyl-t RNA synthetase) [32]. They are grouped into two subclasses: subclass I (molecular weight lower than 5 kDa) and subclass II (molecular weight ranging from 5 to 10 kDa) [42,43] (Figure 3).
Colicins (MW > 10 kDa) are made mainly by Enterobacteriaceae (mainly E. coli) [44]. They can form pores in the cell wall or degrade bacteria nucleic acid structures (RNAses, DNAses, or tRNAses) [32]. Colicins can be grouped into four subclasses: colicins forming channels in the cytoplasmic membrane, colicins degrading DNA, colicins targeting rRNA or tRNA, and colicins inhibiting murein and lipopolysaccharide biosynthesis [45,46] (Figure 3).
Colicin-like bacteriocins are made by the Klebsiella genus (klebicins) and P. aeruginosa (S-type pyocins) [46]. They are similar in size, structure, and function to colicins.
Phage-tail-like bacteriocins have structures similar to phage tails. They are cylindrical peptides with high molecular weights [23]. They are grouped into the R-type and F-type subclasses [23] (Figure 3).
R-type phage-tail-like bacteriocins bind to cell surface receptors, force the internal core into the microbial cell envelope, and determine rapid cell death [23]. They also affect macromolecule synthesis and oxygen uptake [9]. F-type phage-tail-like bacteriocins have a mechanism of action similar to R-type, but do not have contractile movement [27].

2.3. Fungal AMPs

Fungi produce peptaibols and fungal defensins [47,48] (Figure 4).
The term “peptaibol” is linked to structural characteristics. It is a combination of the words “peptide,” “α-aminoisobutyrate,” and “amino alcohol” [49]. Peptaibols are mainly made by Trichoderma fungi [50]. They are short peptides (containing 5–21 amino acids) with a high proportion of non-proteinogenic amino acids (i.e., α-aminoisobutyric acid), acylated N-terminal residue, and amino alcohol (i.e., leucenol or phenylalaninol) linked to the C-terminal [51]. Their three-dimensional structures consist of α-helix and β-bend patterns [52]. They are classified based on sequence length as “long” (18−20 residues), “short” (11−16 amino acids) (Figure 4), and founded on modification types on the terminal groups, “lipo” peptaibols (i.e., N-terminal acylated by decanoic) [53]. Different mechanisms have been proposed to describe their action. Concerning large peptaibols, it was hypothesized that their helical structures oligomerize and can form ion channels in the membrane. Instead, short peptaibols can form a pore via helical bundles (within the bilayer or by a barrel-stave mechanism) and interact with diverse molecular targets [9]. Peptaibols’ modes of action that do not involve interaction with the bacterial membrane include the inhibition of cell wall synthesis, DNA, protein synthesis, and that of relevant enzymes [10].
Eukaryotes and bacteria can produce defensins. Defensins are a class of cysteine-rich AMPs with short, cationic disulfide bridges [54]. They can be grouped into two superfamilies (cis and trans) (Figure 4). Fungi can produce cis-defensins with α-helical (cysteine-stabilized) or β-sheet folds. Defensins can disrupt the microbial cytoplasmic membrane, bind the bacterial precursor lipid II of the cell wall, or prevent cell wall biosynthesis [55].

2.4. Plant AMPs

Plant AMPs are the first line of defense against infections produced by pathogenic microorganisms. They can have diverse structures and action mechanisms. Their classification is based on their tridimensional structures and amino acid sequence similarity, including thionins, hevein-like peptides, defensins, knottins, stable-like peptides, snakins, lipid transfer proteins, and cyclotides [56] (Figure 5).

2.4.1. Thionins

Thionins are classified into five types indicated by Roman numerals, have sizes ranging from 45 to 48, and are found in monocots and dicots. They include two distinct superfamilies: α/β-thionins and γ-thionins [57]. α/β thionins have similar structures (homologous amino acid sequences) [58] and are rich in arginine, cysteine, and lysine. γ-thionins are similar to defensins, so some authors classify them in this group [59]. Thionins have a broad spectrum of activities. They act against Gram-positive and Gram-negative bacteria, yeast, fungi, insect larvae, and nematodes [60,61,62] and present cytotoxic effects against mammal cells in vitro [63].

2.4.2. Hevein-like peptides

Hevein-like peptides can contain 29–45 amino acids with glycine (6), cysteine (8–10), and aromatic residues. They have a chitin-binding domain responsible for their antifungal activity [64] and 3–5 disulfide bonds that stabilize the antiparallel β-sheets and short α-helix [65]. The factors that favor chitin-binding are the three aromatic amino acids that give stability to the hydrophobic C-H group, the π electron system that determines van der Waals forces, and the hydrogen bonds between serine and N-acetylglucosamine [64]. Hevein-like peptides damage the fungal cell wall by interacting with hydrophobic residues and chitin present in the fungal cell [5]. They can constrain some enzymes’ activities by linking them with disulfide bonds [66].

2.4.3. Defensins

Defensins can comprise 45–54 amino acids and four disulfide bridges. They have an antiparallel β sheet, are enclosed by an α-helix, and are limited by intramolecular disulfide bonds [67] called cysteine-stabilized αβ (CSαβ) motifs [68]. Defensins are resistant to proteolysis and are stable to variations in temperature and pH. They prevent microbial growth, trypsin, and α-amylase activities, decrease abiotic stress, and change the redox state of ascorbic acid [56].

2.4.4. Knottins

Knottins, also called “cysteine-knot peptides”, are formed by 39 amino acids (of which six are cysteine residues), have three disulfide bonds (cysteine-knot motifs), and can be found in two conformations (cyclic and linear) [5,69]. They have high thermal stability and resistance to proteolytic action and can inhibit α-amylase, trypsin, carboxypeptidase, and cysteine protease [70,71]. They differ from protease inhibitors and defensins regarding cysteine space [5]. They are amphipathic peptides whose cationic portions can bind cell membranes, acid-sensing channels, and K+ and Na+ channels in membranes. Once they enter a cell, they attack intracellular targets (i.e., carboxypeptidases) and promote resistance [61]. Unfortunately, knottins are highly cytotoxic to human cells since their contact with membranes is not selective.

2.4.5. Stable-like Peptides

Stable-like peptides are a class of small peptides that form a helix-loop-helix structure with a typical Cys motif of XnC1X3C2XnC3X3C4Xn (-X is an amino acid residue different from cysteine). Although their amino acid sequence is highly variable, the three-dimensional structure of stable-like peptides is conserved. They can have antifungal, antibacterial, ribosome-inactivating, and trypsin inhibiting activities [72]. Their bacteriostatic effect is due to binding with DNA, which decreases RNA and protein synthesis [73]. Their activity relates to the loop region that connects the two α-helices [74].

2.4.6. Snakins

Snakins are generally small (~7 kDa), cysteine-rich, and positively charged proteins with antimicrobial, antinematode, and antifungal properties [75]. The mechanism of action is not precise. More than one hypothesis has been developed to explain it. Some authors believe they can promote immune responses by destabilizing the site of action through interaction with the negatively charged component [76,77]. Other authors hypothesized that they can act on phytohormone biosynthesis and transduction processes [78].

2.4.7. Lipid Transfer Proteins

Lipid transfer proteins (LTPs) are small, cysteine-rich proteins (containing 100 aa) having 4 to 5 helices in their structure that are stabilized by hydrogen bonds. They can transfer lipids (i.e., fatty acids, phospholipids, acyl CoA fatty acids, and sterols) between membranes. In this way, they form pores and determine cell death. They can be classified into two subfamilies, LTP1 (relative molecular weight of 9  kDa) and LTP2 (relative molecular weight of 7 kDa), or into five types (LTP1, LTP2, LTPc, LTPd, and LTPg) based on the position of the conserved intron, the space between the cysteine residues, and the identity of the amino acid sequence [69].

2.4.8. Cyclotides

Cyclotides are macrocyclic with cyclic cystine knot (CCK) structural motif peptides [79]. Disulfide bridges stabilize the head-to-tail cyclo. They can be classified into two subfamilies: Möbius and bracelets [80]. Their action depends on the cystine knot structural motif that promotes hydrophobic residue surface contact, some of which form a hydrophobic patch [81]. Cyclotides can act against bacteria, helminths, insects, and mollusks and have ecbolic anti-HIV and anticancer properties [81].

2.5. Animal AMPs

Vertebrate defensins are synthesized as “prepropeptides” and classified into α, β, and θ defensins [82]. They have short polypeptide sequences (18–45 amino acids), cationic net charges (+1 to +11), and three intramolecular disulfide bonds. In human α-defensins, the characteristic connections of disulfide bridges are Cys1–Cys6, Cys2–Cys4 and Cys3–Cys5 [83]. They are synthesized by promyelocytes and intestinal Paneth cells [84]. β-defensins differ from α-defensins in disulfide bond distributions and cysteine residues. The disulfide bridges in human β-defensins are Cys1–Cys5, Cys2–Cys4 and Cys3–Cys6 [83]. θ-defensins are cyclic octadecapeptides not expressed in humans and are active against B. anthrax, S. aureus, and C. albicans [85,86,87]. They contain a macrocyclic backbone and are structurally dissimilar to α- and β-defensins [88].
Invertebrates synthesize AMPs as components of humoral defense [89]. They are cationic peptides that can contain six or eight cysteine residues and show a cysteine-stabilized α/β motif [90]. The defensins produced by insects, arthropods, and mollusks contain six cysteines.
Eight cysteines form defensins made by mollusks and nematodes [91].
Invertebrate defensins are phylogenetically and structurally associated with vertebrate β-defensins. They have a hydrophobic domain (N-terminal) that can act against Gram-positive bacteria and a cationic domain (C-terminal containing six cysteines) that can act against Gram-negative bacteria [92].
Crustins (cationic cysteine-rich peptides that form a tightly packed structure) are found in crustaceans [93]. They have an N-terminal multidomain (rich in glycine, cysteine, and proline) and a C-terminal (with four C-terminal disulfide bridges) (Table 1) [94].
In fish, reptiles, amphibians, birds, and mammalians, AMPs ( size range of 15–200 residues) play an essential role in the immediate response to microorganisms [9]. Fish produce β-defensins, cathelicidins, hepicidins, histone-derived peptides, and piscidins [95] (Table 1).
Fish defensins are β-defensin-like proteins containing six cysteine motifs [96]. Cathelicidins are cationic proteins activated by elastase and other proteases discovered in the secretory granules of immune cells [97]. They act against Gram-positive and Gram-negative bacteria, parasites, fungi, and enveloped viruses [98,99,100,101,102]. Cathelicidins can bind and disrupt negatively charged membranes, alter RNA and DNA synthesis, damage the functions of enzymes and chaperones, and promote protein degradation [103].
Fish hepcidins are cysteine-rich peptides similar to human hepcidin with a hairpin structure linked via four disulfide bonds. They are iron-regulating antimicrobial hormones [104,105].
Hepcidins are grouped into HAMP1 and HAMP2 [95]. They act against bacteria (Gram-positive and Gram-negative) and fish pathogens and induce the internalization and degradation of ferroportin [106].
Piscidins are linear amphipathic AMPs. They have histidine residue and an α-helix that can interact with lipid bilayers [107]. They are classified into piscidins 1–7 based on their biological activity, amino acid sequence, and length [107].
Reptiles and avians produce cathelicidins and defensins (α-, β-, and θ-defensins) [108]. Cathelicidins are small-sized proteins made by macrophages and neutrophils [109].
Amphibians can produce magainin and cancrin (GSAQPYKQLHKVVNWDPYG) [13].
Mammalians make cathelicidin, defensin, platelet antimicrobial protein, dermcidin, and hepcidin AMPs [110]. Mammalian cathelicidins are cationic peptides with an amphipathic structure that assume α-helical, elongated conformations or β-hairpin forms [9].

3. Antimicrobial Peptide Structures and Activities

Most AMPs are made up of from 5 to 100 amino acids and have a positive net charge (generally lysine, arginine, and histidine amino acids; +2 to +11), with about 50% hydrophobic residues (generally aliphatic and aromatic amino acids) placed in variable sequence lengths [111,112]. AMPs can adapt to various structural changes when contacting microbe membranes [113]. Their amino acid compositions determine their charges, hydrophobic, and amphiphilic properties [114]. The number and quality of amino acids determine an AMP’s pharmacological applications. Generally, shorter AMPs are more antibacterial than long-chain linear peptides, which exhibit more hemolytic and cytotoxic activity [115]. Peptides with extremely short lengths have reduced antimicrobial potency since they have difficulty forming the amphipathic secondary structures responsible for the membrane-disruption capacity [116]. Generally, small amino acids, such as glycine, increase an AMP’s activity [117]. Glycine-rich peptides have high selectivity and antimicrobial ability (especially against Gram-negative bacteria) [118], as well as and antimycotic and anticancer activities [119,120].
Glycine-rich AMPs with net charges ranging from −1 to −2 that require cations as cofactors (i.e., Zn2+) have biocidal activity obtained by improving the eukaryotic innate immune response [121]. Proline-rich peptides can enter through membrane protein channels in the bacterial cytosol and modulate the immune system via angiogenesis or cytokine activity [122,123].
Cysteine-rich peptides can form pores in membranes [124]. The Cys residues can improve the AMP antimicrobial activity by stabilizing sheet or β-hairpin structures [125].
Aromatic-amino-acid-rich peptides cross the microbial membrane and disrupt it [126]. Trypsin-rich peptides stabilize the AMP tertiary structure since trypsin–trypsin interactions give a cross-strand contact [112].
Phenylalanine-rich peptides are highly hydrophobic molecules with intense antimicrobial activity against bacteria (Gram-positive and Gram-negative) and yeast [127]. They do not exhibit hemolytic activity [128].
Some lipopeptides (i.e., daptomycin, polymyxins B and E) and glycopeptides (i.e., teicoplanin, vancomycin, dalbavancin, telavancin, and oritavancin) are currently used for clinical purposes [129]. AMP secondary structures can be α-helices, β-sheets, non- α- or β- structures, or mixed structures [130]. Usually, amino acids with high helical propensity (i.e., alanine, arginine, leucine, lysine, etc.) synthesize novel antimicrobial peptides since α-helical structures promote interaction with membranes and determine membrane lysis [131,132,133]. Other characteristics that affect AMP activity are hydrophobicity and amphipathicity. AMPs with low hydrophobicity have antimicrobial activities since the self-association of peptides stops peptide passage through the cell wall [134]. Amphipathic AMPs have bactericidal and cytotoxic activities linked to their aptitude to form an α-helix [135]. They can interact with intracellular targets, damaging the membrane structure or making transient pores [134]. AMPs with high hydrophobicity have antimicrobial and hemolytic activities [136]. The high hydrophobicity of the α-helix improves the antimicrobial activity since the self-association of peptides stops peptide passage across the microbial cell wall [137] and enhances hemolytic activity, inducing peptides to penetrate deeper into the hydrophobic core of red blood cells [138]. In addition, amphipathic characteristics affect AMP activities. Imperfect amphiphilic peptides have more significant antimicrobial activity than perfect ones [139].

4. Antimicrobial Peptide Action

Mostly AMPs have a short half-life. They can act by disrupting the microbial membrane or without affecting membrane stability [9].

4.1. AMPs with Action on Cell Membranes

AMPs can make electrostatic interactions between their positive charges and the microbial cell surface’s negative ones, as well as hydrophobic relations between their amphipathic domain and the microbic membrane phospholipids [140]. The physical–chemical interactions and the interfacial properties determine the destabilization and permeabilization of the microbial membrane [8,141]. Both vertebrates and invertebrates produce AMPs (active in vitro at micromolar levels), which can affect the cell membrane by manipulating its components [142]. Gram-positive bacteria have a dense peptidoglycan layer, while Gram-negative ones have a fine peptidoglycan layer and an extra outer membrane [143]. Teichoic acid and lipopolysaccharides provide electronegative charges on the bacterial surface.
On the contrary, mammalian cell membranes do not have a net charge since the outer leaflet is formed by zwitterionic phospholipids (i.e., phosphatidylcholine, phosphatydylethanolamine, and sphingomyelin) [144] and the phospholipid bilayer is stabilized by cholesterol [145]. Thus, positively charged AMPs are significantly attracted by the negative charge (i.e., phospholipids, cardiolipin, phosphatidylglycerol, and phosphatidylserine) on bacterial membranes; instead, only weak hydrophobic interactions between AMPS and mammalian cell membranes can occur. Therefore, AMPs give selective antimicrobial effects without harming normal cells since the eukaryotic cell membranes have uncharged neutral residues (generally phospholipids, cholesterol, and sphingomyelins), which cannot interact with AMPs.
Highly cationic and anionic peptides have no antimicrobial activity [146,147].
Pore formation can be achieved by barrel-stave, toroidal pore [148], and carpet-like [149] mechanisms, the clustering of anionic lipids [150], aggregated channels [151], or more than one mechanism [9].
The barrel-stave model hypothesizes that AMPs place themself alongside a membrane and penetrate the lipid bilayer. The pore external face is made by aligning the hydrophobic region of AMPs with the lipid bilayer’s central lipid region. Instead, the pore interior is made by the peptide hydrophilic contribution (by a peptide–peptide interaction) [152]. Barrel-like pores can determine cytoplasmic outflow, membrane collapse, and cell death [153] (Figure 6).
The toroidal pore model hypothesizes that AMPs vertically cross a lipid membrane without peptide–peptide interactions in the lipid membrane [152]. The pores are transient and less stable than barrel-stave formations [154] (Figure 6).
The carpet or detergent-like model assumes that AMPs are adsorbed parallel to the lipid membrane until wholly covered (like a carpet), inducing membrane disruption. In this process, no peptides across the membrane, peptide–peptide interactions, or peptide structures are made [155] (Figure 6). The AMP hydrophobic regions interact with the cell membrane, and the hydrophilicity ends with an aqueous solution [156].
Anionic lipid-clustering activity is obtained by forming phase-boundary defects between lipid domains due to interaction between cationic AMPs and anionic-charged lipids [150].

4.2. AMPs with No Action on Cell Membranes

Some AMPs can kill bacteria interfering with DNA (replication, transcription, and translation), cell division, and the blocking of protein biosynthesis and folding [113,157,158]. They can also interfere with the immune system, activating white blood cells, improving angiogenesis, blocking reactive oxygen and nitrogen species [159], suppressing toll-like receptors, reducing anti-endotoxin activity, interfering with cytokine-mediated production of cytokines [160], and influencing T- and B-cell activities [161]. Moreover, AMPs can bind cell membrane receptors (alternate ligand model) or affect receptor activation (membrane disruption model) by altering a receptor’s site or releasing a membrane-bound factor (transactivation model) that binds the receptor. Finally, AMPs can interfere with lipopolysaccharides, preventing inflammation [162].

5. AMP Potential in the Food Field

5.1. AMPs in Food Preservation

AMP application in food preservation is under review since they have a broad spectrum of activity (bacteria, fungi, and protozoa), good water solubility, and are thermostable, but the high cost of large-scale production limits their use [163,164]. AMP selection for food incorporation depends on their spectrum of activity and an AMP’s specificity toward microorganisms in a food product. For example, fermenticins produced by Lactobacillus fermentum [165] and defensins, which act on lipid II and lipid A at the bacterial membrane, show broad spectra of activity, pH, and temperature stability [166]. Some AMPs prolong the shelf-life of food by acting as antimicrobials and inhibiting lipid oxidation, such as peptides from Cynoscion guatucupa protein hydrolysate obtained by enzymatic hydrolysis with Alcalase and Protamex [167]. AMPs, stable at diverse ranges of pH levels, temperatures, and proteases, have been studied because, in food technology, temperature variations are used to increase the preservation of food, and proteases can be added to foods to decrease a food’s allergy power and alter its taste [168,169,170]. Tolerance to diverse pH conditions can be obtained by changing the sequence of AMPs. For example, adding histidine at the carboxyl terminus of a piscidin-like AMP allowed a more significant antimicrobial activity against S. aureus at pH 10.5 [171]. AMP stability can be improved by modifying an AMP’s geometrical properties (i.e., the radius of gyration, lipophilicity, ovality, polar surface area, and surface area). For example, the stability of Protegrin-1 was attributed to the high number of hydrogen bonds (distances <2.5 Å) [172]. The presence of free amino, sulfur, and carbonyl functional groups affected AMP bioactivity [173]. The concurrent addition of the additives ascorbate and nitrite could increase the carbonyl compounds in proteins, altering their functionality and technological properties [174]. Similarly, sulfites used as antioxidants and antiseptics could react with the disulfide bonds of AMPs to form irreversibly bound forms of S-sulfonates [175]. Thus, AMPs can be added to low-reactive foods such as fiber-rich food (whole-grain bread, cereals, pseudocereals, legumes, nuts, fruits, and vegetables) [176] and should not be inserted into high-reactivity food, such as liquid-based food formulations [177].
Nanoparticles, nanofibers, and nanoliposomes have been examined to protect AMP antimicrobial activity [178,179,180]. For example, nisin was placed into multifunctional soy-soluble polysaccharide-based nanocarriers to enhance its stability and preserve antioxidant and antimicrobial activity [181]. In raw and pasteurized milk, nisin-loaded chitosan/alginate nanoparticles were employed to prevent the growth of S. aureus during long incubation periods [182]. The nano-encapsulation of temporin B into chitosan nanoparticles enhanced the peptide’s antibacterial activity [183].

5.2. AMPs in Food Packaging

Active packaging systems have been developed to control the release of AMPs and decrease their interactions with food components. Pentocina MS1 and MS2 from Lactobacillus pentosus MS031 isolated from Chinese Sichuan paocai were added to fresh-cut fruits in cold packaging to decrease the growth of Salmonella typhi, Listeria monocytogenes, and E. coli [184]. Partially purified Gt2 peptides active against E. coli and Salmonella typhi were put into packages to preserve tomatoes [185]. The peptide MTP1 was employed in meat and dairy product packaging [186]. Nisin, which can inhibit the growth of Listeria monocytogenes, Staphylococcus aureus, Penicillium sp., and Geotrichum sp., was used to preserve mozzarella cheese. [187]. A fish protein hydrolysate was added to preserve fish flounder fillets [188]. Nisin preventing the growth of Listeria monocytogenes was employed to preserve cold-smoked salmon [189].

6. AMP Potential in the Pharmaceutical Field

6.1. AMP Antioxidant Potential

Some AMPs can act as free-radical scavengers, reduce lipid peroxidation, have metal ion chelation activity, and impact antioxidant enzyme activity (i.e., SOD, PPO, CAT, and GSH-Px) [190]. The presence of isoleucine, leucine, and histidine amino acids [191], as well as the number of active hydrogen sites, are essential for antioxidant activity [192].

6.2. Antineoplastic Agent

Currently, cancer is a leading cause of death worldwide. AMPs have some characteristics that make them potential drugs for cancer therapy, such as high activity, specificity and affinity, small size, slight drug–drug interaction, aptitude to cross membranes, and low toxic side effects since they do not accumulate in vital organs (i.e., the liver and kidneys) [193]. Moreover, they are easily modified and synthesized [194] and are less immunogenic than recombinant antibodies [195]. Therapeutic peptides are classified into three groups: antimicrobial or pore-forming peptides (anticancer peptides, or ACPs, naturally produced by all living creatures), cell-permeable peptides, and tumor-targeting peptides [195].

6.3. AMP Potential against Respiratory Diseases

Some natural and modified AMPs appear to have potential as drugs to cure respiratory diseases and as infection markers.
Pyocins, which can inhibit the growth of P. aeruginosa, could be used to cure fibrosis patients [196].
Esc (1−21)-c, a partial D-derivative of esculentin-1 that can decrease P. aeruginosa infection and has excellent resistance to degradation due to the elastase enzyme [197], could be employed to promote bronchial epithelium repair [198].
α-and β-defensins were potential infection markers of upper respiratory tract infection [199].

6.4. AMP Potential against Hypertension

Some AMPs (SAGGYIW and APATPSFW) could inhibit angiotensin-converting I (ACE), blocking the active site via weak interactions (i.e., electrostatic interaction, hydrogen bonds, and Van Der Waals interactions) [200]. ACE is an enzyme that can convert decapeptide angiotensin I (inactive) into octapeptide angiotensin II (vasoconstrictor), which is involved in hypertension and atherosclerosis [201].

6.5. AMP Potential against Obesity

EITPEKNPQLR, CQPHPGQTC, and RKQEEDEDEEQQRE are AMPs preventing pancreatic lipase activity. Pancreatic lipase is an enzyme that can hydrolyze 50–70% of food-derived fat in human organisms. Therefore, its inhibition is helpful in obesity treatment [202].

6.6. AMP Potential against Intestine Infection and Inflammation

α-defensins and C-type lectins (AMPs) are expressed in the gastrointestinal tract to sustain intestine symbiosis and protect it from pathological bacterial translocation [203].

6.7. AMP Potential against Viral Infections

Some AMPs can act against DNA and RNA viruses [204,205]. They can act on the viral envelope or after adsorption on the viral surface [206]. AMP positively charged residues can interact electrostatically with negatively charged cell surface molecules, such as heparan sulfate (glycosaminoglycans) [207], prevent the spread across tight junctions of the virus from one cell to another cell (cell-to-cell spread), or prevent the formation of giant cells (syncytium) [13].
Lactoferrin (iron-binding glycoprotein) can act as an antiviral material by inhibiting the replication of a wide range of DNA and RNA viruses or preventing virus entry into a host cell through direct binding to virus particles or blocking cellular receptors [208].
Defensins (α- and β-) can act against human immunodeficiency virus (HIV), influenza, herpes simplex virus (HSV), and SARS-CoV [209]. It has also been hypothesized that an infusion of defensins during Cytomegalovirus infections may be helpful in the treatment of COVID-19 in pregnant women [210,211].
Frog-skin-derived peptide AR-23 and some of its derivatives can act against the viral surfaces of all enveloped viruses (i.e., coronaviruses, including SARS-CoV-2; paramyxoviruses; and herpesvirus) [212,213].

6.8. AMP Potential against Skin Infections

AMPs can be considered as a therapeutic option since they have a broad spectrum of biological activities against microbes; remain on an application site when topically administrated; and support wound healing by controlling angiogenesis, cell migration, and cytokine release chemotaxis [214]. Human keratinocytes and the granular skin layer make and store AMPs and lipids within secretory granules (lamellar bodies) [215]. The lamellar bodies make a physical barrier in superficial layers of the epidermis that can inhibit microbial growth and water loss. RNase 5 and RNase 7 are AMPs present in healthy human skin. They are active on Gram-negative and Gram-positive bacteria [216]. Other AMPs involved in skin wellbeing are psoriasin; calprotectin (iron- and zinc-binding S100 proteins) expressed by keratinocytes; β-defensins; the cathelicidin hCAP18, which must be converted to the active form; LL-37; histone 4 (active against Gram-positive bacteria) and dermcidin (active against antibacterial and antifungal mechanisms) produced by pilosebaceous follicles and eccrine glands, respectively; and α-defensins and LL-37 formed by neutrophils and natural killer cells [217]. Bee venom peptides can be helpful as a topical agent to promote skin regeneration and acne treatment. [218,219,220].

7. Conclusions

This work summarized the current knowledge regarding antimicrobial biopeptides to highlight their potential applications in the industrial field. Researchers are examining new sources of bioactive materials to use as natural preservatives in foods and to reduce the emergence of antibiotic drug resistance. AMPs seem to have good prospects as natural preservatives incorporated in food and food packaging, as well as for antioxidant, antineoplastic, antiobesity, antihypertensive, anti-inflammatory, antiviral, and dermatological agent drugs. Nanocarriers can be used to improve their bioavailability. Nevertheless, large-scale production and high cost of production could limit their use.

Author Contributions

Methodology, investigation, resources, writing—original draft preparation, I.D.; methodology, investigation, resources, writing—original draft preparation, M.-G.D.B.; methodology, investigation, resources, writing—original draft preparation, A.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. 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] [PubMed]
  2. Rai, M.; Pandit, R.; Gaikwad, S.; Kövics, G. Antimicrobial peptides as natural bio-preservative to enhance the shelf-life of food. J. Food Sci. Technol. 2016, 53, 3381–3394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Johnstone, K.F.; Herzberg, M.C. Antimicrobial peptides: Defending the mucosal epithelial barrier. Front. Oral Health 2022, 3, 958480. [Google Scholar] [CrossRef] [PubMed]
  4. Liu, Y.; Sameen, D.E.; Ahmed, S.; Dai, J.; Qin, W. Antimicrobial peptides and their application in food packaging. Trends Food Sci. Technol. 2021, 112, 471–483. [Google Scholar] [CrossRef]
  5. Tam, J.P.; Wang, S.; Wong, K.H.; Tan, W.L. Antimicrobial Peptides from Plants. Pharmaceuticals 2015, 8, 711–757. [Google Scholar] [CrossRef]
  6. Dini, I. Chapter 14—Use of Essential Oils in Food Packaging. In Essential Oils in Food Preservation, Flavor and Safety; Academic Press: Cambridge, MA, USA, 2016; pp. 139–147. [Google Scholar]
  7. Upton, M.; Cotter, P.; Tagg, J. Antimicrobial peptides as therapeutic agents. Int. J. Microbiol. 2012, 2012, 326503. [Google Scholar] [CrossRef]
  8. 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] [Green Version]
  9. Bin Hafeez, A.; Jiang, X.; Bergen, P.J.; Zhu, Y. Antimicrobial Peptides: An Update on Classifications and Databases. Int. J. Mol. Sci. 2021, 22, 11691. [Google Scholar] [CrossRef]
  10. Wang, C.K.; Craik, D.J. Designing macrocyclic disulfide-rich peptides for biotechnological applications perspective. Nat. Chem. Biol. 2018, 14, 417–427. [Google Scholar] [CrossRef]
  11. Henriques, S.T.; Lawrence, N.; Chaousis, S.; Ravipati, A.S.; Cheneval, O.; Benfield, A.H.; Elliott, A.G.; Kavanagh, A.M.; Cooper, M.A.; Chan, L.Y.; et al. Redesigned Spider Peptide with Improved Antimicrobial and Anticancer Properties. ACS Chem. Biol. 2017, 12, 2324–2334. [Google Scholar] [CrossRef]
  12. Borah, A.; Deb, B.; Chakraborty, S. A Crosstalk on Antimicrobial Peptides. Int. J. Pept. Res. Ther. 2020, 27, 229–244. [Google Scholar] [CrossRef]
  13. 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] [PubMed]
  14. Danis-Wlodarczyk, K.M.; Wozniak, D.J.; Abedon, S.T. Treating Bacterial Infections with Bacteriophage-Based Enzybiotics: In Vitro, In Vivo and Clinical Application. Antibiotics 2021, 10, 1497. [Google Scholar] [CrossRef] [PubMed]
  15. Nelson, D.; Loomis, L.; Fischetti, V.A. Prevention and elimination of upper respiratory colonization of mice by group A streptococci by using a bacteriophage lytic enzyme. Proc. Natl. Acad. Sci. USA 2001, 98, 4107–4112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Abdelrahman, F.; Easwaran, M.; Daramola, O.I.; Ragab, S.; Lynch, S.; Oduselu, T.J.; Khan, F.M.; Ayobami, A.; Adnan, F.; Torrents, E.; et al. Phage-Encoded Endolysins. Antibiotics 2021, 10, 124. [Google Scholar] [CrossRef] [PubMed]
  17. Yan, J.; Mao, J.; Xie, J. Bacteriophage polysaccharide depolymerases and biomedical applications. BioDrugs 2014, 28, 265–274. [Google Scholar] [CrossRef] [PubMed]
  18. Ha, E.; Son, B.; Ryu, S. Clostridium perfringens virulent bacteriophage CPS2 and its thermostable endolysin lysCPS2. Viruses 2018, 10, 251. [Google Scholar] [CrossRef] [Green Version]
  19. Plotka, M.; Kapusta, M.; Dorawa, S.; Kaczorowska, A.K.; Kaczorowski, T. Ts2631 endolysin from the extremophilic thermus scotoductus bacteriophage vB_Tsc2631 as an antimicrobial agent against gram-negative multidrug-resistant bacteria. Viruses 2019, 11, 657. [Google Scholar] [CrossRef] [Green Version]
  20. Pastagia, M.; Schuch, R.; Fischetti, V.A.; Huang, D.B. Lysins: The arrival of pathogen-directed anti-infectives. J. Med. Microbiol. 2013, 62, 1506–1516. [Google Scholar] [CrossRef]
  21. Abril, A.G.; Carrera, M.; Notario, V.; Sánchez-Pérez, Á.; Villa, T.G. The Use of Bacteriophages in Biotechnology and Recent Insights into Proteomics. Antibiotics 2022, 11, 653. [Google Scholar] [CrossRef]
  22. Latka, A.; Maciejewska, B.; Majkowska-Skrobek, G.; Briers, Y.; Drulis-Kawa, Z. Bacteriophage-encoded virion-associated enzymes to overcome the carbohydrate barriers during the infection process. Appl. Microbiol. Biotechnol. 2017, 101, 3103–3119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Scholl, D. Phage Tail-Like Bacteriocins. Annu. Rev. Virol. 2017, 4, 453–467. [Google Scholar] [CrossRef] [PubMed]
  24. Daw, M.A.; Falkiner, F.R. Bacteriocins: Nature, function and Structure. Micron 1996, 27, 467–479. [Google Scholar] [CrossRef]
  25. Chen, J.; Zhu, Y.; Yin, M.; Xu, Y.; Liang, X.; Huang, Y.P. Characterization of maltocin S16, a phage tail-like bacteriocin with antibacterial activity against Stenotrophomonas maltophilia and Escherichia coli. J. Appl. Microbiol. 2019, 127, 78–87. [Google Scholar] [CrossRef] [PubMed]
  26. Morse, S.A.; Jones, B.V.; Lysko, P.G. Pyocin inhibition of Neisseria gonorrhoea: Mechanism of action. Antimicrob. Agents Chemother. 1980, 18, 416–423. [Google Scholar] [CrossRef] [Green Version]
  27. Lee, G.; Chakraborty, U.; Gebhart, D.; Govoni, G.R.; Zhou, Z.H.; Scholl, D. F-type bacteriocins of Listeria monocytogenes: A new class of phage tail-like structures reveals broad parallel coevolution between tailed bacteriophages and high-molecular-weight bacteriocins. J. Bacteriol. 2016, 198, 2784–2793. [Google Scholar] [CrossRef] [Green Version]
  28. Tajbakhsh, M.; Karimi, A.; Fallah, F.; Akhavan, M.M. Overview of ribosomal and non-ribosomal antimicrobial peptides produced by Gram positive bacteria. Cell. Mol. Biol. 2017, 6, 20. [Google Scholar] [CrossRef] [PubMed]
  29. Diep, D.; Nes, I. Ribosomally Synthesized Antibacterial Peptides in Gram Positive Bacteria. Curr. Drug Targets 2005, 3, 107–122. [Google Scholar] [CrossRef]
  30. Yeaman, M.R.; Yount, N.Y. Mechanisms of antimicrobial peptide action and resistance. Pharmacol. Rev. 2003, 55, 27–55. [Google Scholar] [CrossRef] [Green Version]
  31. Kumariya, R.; Kumari, G.; Raiput, Y.S.; Akhtar, N.; Patel, S. Bacteriocins: Classification, synthesis, mechanism of action and resistance development in food spoilage causing bacteria. Microb. Pathog. 2019, 128, 171–177. [Google Scholar] [CrossRef]
  32. Meade, E.; Slattery, M.A.; Garvey, M. Bacteriocins, Potent Antimicrobial Peptides and the Fight against Multi Drug Resistant Species: Resistance Is Futile? Antibiotics 2020, 9, 32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Gradisteanu Pircalabioru, G.; Popa, L.I.; Marutescu, L.; Gheorghe, I.; Popa, M.; Czobor Barbu, I.; Cristescu, R.; Chifiriuc, M.-C. Bacteriocins in the Era of Antibiotic Resistance: Rising to the Challenge. Pharmaceutics 2021, 13, 196. [Google Scholar] [CrossRef] [PubMed]
  34. Bierbaum, G.; Sahl, H.-G. Lantibiotics: Mode of Action, Biosynthesis and Bioengineering. Curr. Pharm. Biotechnol. 2009, 10, 2–18. [Google Scholar] [CrossRef] [PubMed]
  35. Cotter, P.D.; Hill, C.; Ross, R.P. Food microbiology: Bacteriocins: Developing innate immunity for food. Nat. Rev. Microbiol. 2005, 3, 777–788. [Google Scholar] [CrossRef] [PubMed]
  36. Cintas, L.M.; Casaus, P.; Håvarstein, L.S.; Hernández, P.E.; Nes, I.F. Biochemical and genetic characterization of enterocin P, a novel sec-dependent bacteriocin from Enterococcus faecium P13 with a broad antimicrobial spectrum. Appl. Environ. Microbiol. 1997, 63, 4321–4330. [Google Scholar] [CrossRef] [Green Version]
  37. Nissen-Meyer, J.; Oppegård, C.; Rogne, P.; Haugen, H.S.; Kristiansen, P.E. Structure and mode-of-action of the two-peptide (class-IIb) bacteriocins. Probiotics Antimicrob. Proteins 2010, 2, 52–60. [Google Scholar] [CrossRef] [Green Version]
  38. Van Belkum, M.J.; Martin-Visscher, L.A.; Vederas, J.C. Structure and genetics of circular bacteriocins. Trends Microbiol. 2011, 19, 411–418. [Google Scholar] [CrossRef]
  39. Nissen-Meyer, J.; Rogne, P.; Oppegard, C.; Haugen, H.; Kristiansen, P. Structure-Function Relationships of the Non-Lanthionine-Containing Peptide (class II) Bacteriocins Produced by Gram-Positive Bacteria. Curr. Pharm. Biotechnol. 2009, 10, 19–37. [Google Scholar] [CrossRef]
  40. Heng, N.C.K.; Wescombe, P.A.; Burton, J.P.; Jack, R.W.; Tagg, J.R. The diversity of bacteriocins in Gram-positive bacteria. In Bacteriocins: Ecology and Evolution; Riley, M.A., Chavan, M.A., Eds.; Springer: New York, NY, USA, 2007; pp. 45–92. [Google Scholar]
  41. Simons, A.; Alhanout, K.; Duval, R.E. Bacteriocins, Antimicrobial Peptides from Bacterial Origin: Overview of Their Biology and Their Impact against Multidrug-Resistant Bacteria. Microorganisms 2020, 8, 639. [Google Scholar] [CrossRef]
  42. Duquesne, S.; Destoumieux-Garzón, D.; Peduzzi, J.; Rebuffat, S. Microcins, gene-encoded antibacterial peptides from enterobacteria. Nat. Prod. Rep. 2007, 24, 708–734. [Google Scholar] [CrossRef]
  43. Pons, A.M.; Lanneluc, I.; Cottenceau, G.; Sable, S. New developments in non-post translationally modified microcins. Biochimie 2002, 84, 531–537. [Google Scholar] [CrossRef]
  44. Cascales, E.; Buchanan, S.K.; Duche, D.; Kleanthous, C.; Lloubes, R.; Postle, K.; Riley, M.; Slatin, S.; Cavard, D. Colicin biology. Microbiol. Mol. Biol. Rev. 2007, 71, 158–229. [Google Scholar] [CrossRef] [Green Version]
  45. Gillor, O.; Kirkup, B.C.; Riley, M.A. Colicins and microcins: The next generation antimicrobials. Adv. Appl. Microbiol. 2004, 54, 129–146. [Google Scholar] [PubMed]
  46. Papadakos, G.; Wojdyla, J.A.; Kleanthous, C. Nuclease colicins and their immunity proteins. Q. Rev. Biophys. 2012, 45, 57–103. [Google Scholar] [CrossRef] [PubMed]
  47. Duclohier, H. Antimicrobial Peptides and Peptaibols, Substitutes for Conventional Antibiotics. Curr. Pharm. Des. 2010, 16, 3212–3223. [Google Scholar] [CrossRef]
  48. Wu, J.; Gao, B.; Zhu, S. The fungal defensin family enlarged. Pharmaceuticals 2014, 7, 866–880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Evans, B.S.; Robinson, S.J.; Kelleher, N.L. Surveys of non-ribosomal peptide and polyketide assembly lines in fungi and prospects for their analysis in vitro and in vivo. Fungal Genet. Biol. 2011, 48, 49–61. [Google Scholar] [CrossRef] [Green Version]
  50. Tyśkiewicz, R.; Nowak, A.; Ozimek, E.; Jaroszuk-Ściseł, J. Trichoderma: The Current Status of Its Application in Agriculture for the Biocontrol of Fungal Phytopathogens and Stimulation of Plant Growth. Int. J. Mol. Sci. 2022, 23, 2329. [Google Scholar] [CrossRef]
  51. Leitgeb, B.; Szekeres, A.; Manczinger, L.; Vágvölgyi, C.; Kredics, L. The history of Alamethicin: A review of the most extensively studied peptaibol. Chem. Biodivers. 2007, 4, 1027–1051. [Google Scholar] [CrossRef]
  52. Szekeres, A.; Leiteb, B.; Kredics, L.; Antal, Z.; Hatvani, L.; Manczinger, L.; Vágvölgyi, C. Peptaibols and related peptaibiotics of Trichoderma. Acta Microbiol. Immunol. Hung. 2005, 52, 137–168. [Google Scholar] [CrossRef]
  53. Hou, X.; Sun, R.; Feng, Y.; Zhang, R.; Zhu, T.; Che, Q.; Zhang, G.; Li, D. Peptaibols: Diversity, bioactivity, and biosynthesis. Eng. Microbiol. 2022, 2, 100026. [Google Scholar] [CrossRef]
  54. Qi, S.; Gao, B.; Zhu, S. A Fungal Defensin Inhibiting Bacterial Cell-Wall Biosynthesis with Non-Hemolysis and Serum Stability. J. Fungi 2022, 8, 174. [Google Scholar] [CrossRef]
  55. 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] [Green Version]
  56. Lima, A.M.; Azevedo, M.I.; Sousa, L.M.; Oliveira, N.S.; Andrade, C.R.; Freitas, C.D.; Souza, P.F. Plant antimicrobial peptides: An overview about classification, toxicity and clinical applications. Int. J. Biol. Macromol. 2022, 214, 10–21. [Google Scholar] [CrossRef]
  57. Goyal, R.K.; Mattoo, A.K. Plant antimicrobial peptides. In Host Defense Peptides and Their Potential as Therapeutic Agents; Epand, R.M., Ed.; Springer: Cham, Switzerland, 2016; pp. 111–136. [Google Scholar]
  58. Thevissen, K.; Ferket, K.K.A.; François, I.E.J.A.; Cammue, B.P.A. Interactions of antifungal plant defensins with fungal membrane components. Peptides 2003, 24, 1705–1712. [Google Scholar] [CrossRef]
  59. Paul, M.; Chowdhury, T.; Saha, S. Antimicrobial peptide: A competent tool for plant disease control in mulberry—A review. Vegetos 2022, 1–10. [Google Scholar] [CrossRef]
  60. Vasilchenko, A.S.; Smirnov, A.N.; Zavriev, S.K.; Grishin, E.V.; Vasilchenko, A.V.; Rogozhin, E.A. Novel thionins from black seed (Nigella sativa L.) demonstrate antimicrobial activity. Int. J. Pept. Res. Ther. 2017, 23, 171–180. [Google Scholar] [CrossRef]
  61. Moore, S.J.; Leung, C.L.; Cochran, J.R. Knottins: Disulfide-bonded therapeutic and diagnostic peptides. Drug Discov. Today Technol. 2012, 9, e3–e11. [Google Scholar] [CrossRef]
  62. Taveira, G.B.; Carvalho, A.O.; Rodrigues, R.; Trindade, F.G.; Da Cunha, M.; Gomes, V.M. Thionin-like peptide from Capsicum annuum fruits: Mechanism of action and synergism with fluconazole against Candida species. BMC Microbiol. 2016, 16, 12. [Google Scholar] [CrossRef] [Green Version]
  63. Taveira, G.B.; Mello, É.O.; Carvalho, A.O.; Regente, M.; Pinedo, M.; de La Canal, L.; Rodrigues, R.; Gomes, V.M. Antimicrobial activity and mechanism of action of a thionin-like peptide from Capsicum annuum fruits and combinatorial treatment with fluconazole against Fusarium solani. Biopolymers 2017, 108, e23008. [Google Scholar] [CrossRef]
  64. Slavokhotova, A.A.; Shelenkov, A.A.; Andreev, Y.A.; Odintsova, T.I. Hevein-Like Antimicrobial Peptides of Plants. Biochemistry 2017, 82, 1659–1674. [Google Scholar] [CrossRef]
  65. Odintsova, T.; Shcherbakova, L.; Slezina, M.; Pasechnik, T.; Kartabaeva, B.; Istomina, E.; Dzhavakhiya, V. Hevein-like antimicrobial peptides WAMPs: Structure-function relationship in antifungal activity and sensitization of plant pathogenic fungi to tebuconazole by WAMP-2-derived peptides. Int. J. Mol. Sci. 2020, 21, 7912. [Google Scholar] [CrossRef]
  66. Kramer, K.J.; Klassen, L.W.; Jones, B.L.; Speirs, R.D.; Kammer, A.E. Toxicity of purothionin and its homologues to the tobacco hornworm, Manduca sexta (L.) (lepidoptera:Sphingidae). Toxicol. Appl. Pharmacol. 1979, 48, 179–183. [Google Scholar] [CrossRef]
  67. Azmi, S.; Hussain, M.K. Analysis of structures, functions, and transgenicity of phytopeptides defensin and thionin: A review. Beni Suef. Univ. J. Basic Appl. Sci. 2021, 10, 5. [Google Scholar] [CrossRef]
  68. Gao, B.; Zhu, S. A Fungal Defensin Targets the SARS−CoV−2 Spike Receptor−Binding Domain. J. Fungi 2021, 7, 553. [Google Scholar] [CrossRef]
  69. dos Santos-Silva, C.A.; Zupin, L.; Oliveira-Lima, M.; Vilela, L.M.B.; Bezerra-Neto, J.P.; Ferreira-Neto, J.R.; Ferreira, J.D.C.; de Oliveira-Silva, R.L.; de Pires, C.J.; Aburjaile, F.F.; et al. Plant Antimicrobial Peptides: State of the Art, In Silico Prediction and Perspectives in the Omics Era. Bioinform. Biol. Insights 2020, 14, 117793222095273. [Google Scholar] [CrossRef]
  70. Hellinger, R.; Gruber, C.W. Peptide-based protease inhibitors from plants. Drug Discov. Today 2019, 24, 1877–1889. [Google Scholar] [CrossRef]
  71. Molesini, B.; Treggiari, D.; Dalbeni, A.; Minuz, P.; Pandolfini, T. Plant cystine-knot peptides: Pharmacological perspectives. Br. J. Clin. Pharmacol. 2017, 83, 63–70. [Google Scholar] [CrossRef]
  72. Postic, G.; Gracy, J.; Périn, C.; Chiche, L.; Gelly, J.-C. KNOTTIN: The database of inhibitor cystine knot scaffold after 10 years, toward a systematic structure modeling. Nucleic Acids Res. 2017, 46, D454–D458. [Google Scholar] [CrossRef] [Green Version]
  73. Slavokhotova, A.A.; Rogozhin, E.A. Defense Peptides From the α-Hairpinin Family Are Components of Plant Innate Immunity. Front. Plant Sci. 2020, 11, 465. [Google Scholar] [CrossRef]
  74. Haney, E.F.; Petersen, A.P.; Lau, C.K.; Jing, W.; Storey, D.G.; Vogel, H.J. Mechanism of action of puroindoline derived tryptophan-rich antimicrobial peptides. Biochim. Biophys. Acta (BBA) Biomembr. 2013, 1828, 1802–1813. [Google Scholar] [CrossRef] [Green Version]
  75. Rogozhin, E.; Ryazantsev, D.; Smirnov, A.; Zavriev, S. Primary Structure Analysis of Antifungal Peptides from Cultivated and Wild Cereals. Plants 2018, 7, 74. [Google Scholar] [CrossRef] [Green Version]
  76. Su, T.; Han, M.; Cao, D.; Xu, M. Molecular and Biological Properties of Snakins: The Foremost Cysteine-Rich Plant Host Defense Peptides. J. Fungi 2020, 6, 220. [Google Scholar] [CrossRef]
  77. Yeung, H.; Squire, C.J.; Yosaatmadja, Y.; Panjikar, S.; López, G.; Molina, A.; Baker, E.N.; Harris, P.W.; Brimble, M.A. Protein Structures Very Important Paper Radiation Damage and Racemic Protein Crystallography Reveal the Unique Structure of the GASA/Snakin Protein Superfamily. Angew. Chem. 2016, 128, 8062–8065. [Google Scholar] [CrossRef]
  78. Rodríguez, S.; Mariana, D.; Vega, B.; Dans, P.D.; Pandolfi, V.; Benko-Iseppon, A.M.; Cecchetto, G. Antimicrobial and structural insights of a new snakin-like peptide isolated from Peltophorum dubium (Fabaceae). Amino Acids 2018, 50, 1245–1259. [Google Scholar] [CrossRef]
  79. Zhang, S.; Wang, X. One new kind of phytohormonal signaling integrator: Up-and-coming GASA family genes. Plant Signal. Behav. 2017, 12, e1226453. [Google Scholar] [CrossRef] [Green Version]
  80. de Veer, S.J.; Kan, M.-W.; Craik, D.J. Cyclotides: From Structure to Function. Chem. Rev. 2019, 119, 12375–12421. [Google Scholar] [CrossRef]
  81. Carla Barbosa da Silva Lima, S.; Maria Benko-Iseppon, A.; Pacifico Bezerra Neto, J.; Lindinalva Barbosa Amorim, L.; Ribamar Costa Ferreira Neto, J.; Crovella, S.; Pandolfi, V. Plants defense-related cyclic peptides: Diversity, structure and applications. Curr Protein Pept Sci. 2017, 18, 375–390. [Google Scholar]
  82. Huang, Y.-H.; Du, Q.; Craik, D.J. Cyclotides: Disulfide-rich peptide toxins in plants. Toxicon 2019, 172, 33–44. [Google Scholar] [CrossRef]
  83. Selsted, M.E.; Ouellette, A.J. Mammalian defensins in the antimicrobial immune response. Nat. Immunol. 2005, 6, 551–557. [Google Scholar] [CrossRef]
  84. Niyonsaba, F.; Nagaoka, I.; Ogawa, H.; Okumura, K. Multifunctional antimicrobial proteins and peptides: Natural activators of immune systems. Curr. Pharm. Des. 2009, 15, 2393–2413. [Google Scholar] [CrossRef] [PubMed]
  85. Ayabe, T.; Satchell, D.P.; Wilson, C.L.; Parks, W.C.; Selsted, M.E.; Ouellette, A.J. Secretion of microbicidal α-defensins by intestinal Paneth cells in response to bacteria. Nat. Immunol. 2000, 1, 113–118. [Google Scholar] [CrossRef] [PubMed]
  86. Basso, V.; Garcia, A.; Tran, D.Q.; Schaal, J.B.; Tran, P.; Ngole, D.; Aqeel, Y.; Tongaonkar, P.; Ouellette, A.J.; Selsteda, M.E. Fungicidal Potency and Mechanisms of –Defensins against Multidrug-Resistant Candida Species. Antimicrob. Agents Chemother. 2018, 62, e00111-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Wilmes, M.; Stockem, M.; Bierbaum, G.; Schlag, M.; Götz, F.; Tran, D.Q.; Schaal, J.B.; Ouellette, A.J.; Selsted, M.E.; Sahl, H.-G. Killing of Staphylococci by θ-Defensins Involves Membrane Impairment and Activation of Autolytic Enzymes. Antibiotics 2014, 3, 617–631. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Welkos, S.; Cote, C.K.; Hahn, U.; Shastak, O.; Jedermann, J.; Bozue, J.; Jung, G.; Ruchala, P.; Pratikhya, P.; Tang, T.; et al. Humanized theta-defensins (retrocyclins) enhance macrophage performance and protect mice from experimental anthrax infections. Antimicrob. Agents Chemother. 2011, 55, 4238–4250. [Google Scholar] [CrossRef] [Green Version]
  89. Hazlett, L.; Wu, M. Defensins in innate immunity. Cell Tissue Res. 2011, 343, 175–188. [Google Scholar] [CrossRef] [PubMed]
  90. Rowley, A.F.; Powell, A. Invertebrate immune systems specific, quasi-specific, or nonspecific? J. Immunol. 2007, 179, 7209–7214. [Google Scholar] [CrossRef] [Green Version]
  91. Froy, O. Convergent evolution of invertebrate defensins and nematode antibacterial factors. Trends Microbiol. 2005, 13, 314–319. [Google Scholar] [CrossRef]
  92. Tassanakajon, A.; Somboonwiwat, K.; Amparyup, P. Sequence diversity and evolution of antimicrobial peptides in invertebrates. Dev. Comp. Immunol. 2015, 48, 324–341. [Google Scholar] [CrossRef]
  93. Saito, T.; Kawabata, S.I.; Shigenaga, T.; Takayenoki, Y.; Cho, J.; Nakajima, H.; Hirata, M.; Iwanaga, S. A novel big defensin identified in horseshoe crab hemocytes: Isolation, amino acid sequence, and antibacterial activity. J. Biochem. 1995, 117, 1131–1137. [Google Scholar] [CrossRef]
  94. Ranganathan, S.; Simpson, K.J.; Shaw, D.C.; Nicholas, K.R. The whey acidic protein family: A new signature motif and three-dimensional structure by comparative modeling. J. Mol. Graph. Model. 1999, 17, 106–113. [Google Scholar] [CrossRef]
  95. Smith, V.J. Phylogeny of whey acidic protein (WAP) four-disulfide core proteins and their role in lower vertebrates and invertebrates. Biochem. Soc. Trans. 2011, 39, 1403–1408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Masso-Silva, J.A.; Diamond, G. Antimicrobial peptides from fish. Pharmaceuticals 2014, 7, 265–310. [Google Scholar] [CrossRef] [Green Version]
  97. Nam, B.H.; Moon, J.Y.; Kim, Y.O.; Kong, H.J.; Kim, W.J.; Lee, S.J.; Kim, K.K. Multiple β-defensin isoforms identified in early developmental stages of the teleost Paralichthys olivaceus. Fish Shellfish Immunol. 2010, 28, 267–274. [Google Scholar] [CrossRef] [PubMed]
  98. Van Harten, R.M.; Van Woudenbergh, E.; Van Dijk, A.; Haagsman, H.P. Cathelicidins: Immunomodulatory Antimicrobials. Vaccines 2018, 6, 63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Goitsuka, R.; Chen, C.-L.H.; Benyon, L.; Asano, Y.; Kitamura, D.; Cooper, M.D. Chicken cathelicidin-b1, an antimicrobial guardian at the mucosal m cell gateway. Proc. Natl. Acad. Sci. USA 2007, 104, 15063–15068. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Tossi, A.; Scocchi, M.; Zanetti, M.; Storici, P.; Gennaro, R. Pmap-37, a novel antibacterial peptide from pig myeloid cells. Cdna cloning, chemical synthesis and activity. Eur. J. Biochem. 1995, 228, 941–946. [Google Scholar] [CrossRef]
  101. Veldhuizen, E.J.A.; Scheenstra, M.R.; Tjeerdsma-van Bokhoven, J.L.M.; Coorens, M.; Schneider, V.A.F.; Bikker, F.J.; van Dijk, A.; Haagsman, H.P. Antimicrobial and immunomodulatory activity of pmap-23 derived peptides. Protein Pept. Lett. 2017, 24, 609–616. [Google Scholar] [CrossRef]
  102. Wessely-Szponder, J.; Majer-Dziedzic, B.; Smolira, A. Analysis of antimicrobial peptides from porcine neutrophils. J. Microbiol. Methods 2010, 83, 8–12. [Google Scholar] [CrossRef]
  103. Xiao, Y.; Cai, Y.; Bommineni, Y.R.; Fernando, S.C.; Prakash, O.; Gilliland, S.E.; Zhang, G. Identification and functional characterization of three chicken cathelicidins with potent antimicrobial activity. J. Biol. Chem. 2006, 281, 2858–2867. [Google Scholar] [CrossRef] [Green Version]
  104. Liu, Z.-M.; Chen, J.; Lv, Y.-P.; Hu, Z.-H.; Dai, Q.-M.; Fan, X.-L. Molecular characterization of a hepcidin homologue in starry flounder (Platichthys stellatus) and its synergistic interaction with antibiotics. Fish Shellfish Immunol. 2018, 83, 45–51. [Google Scholar] [CrossRef] [PubMed]
  105. Huang, P.H.; Chen, J.Y.; Kuo, C.M. Three different hepcidins from tilapia, Oreochromis mossambicus: Analysis of their expressions and biological functions. Mol. Immunol. 2007, 44, 1922–1934. [Google Scholar] [CrossRef] [PubMed]
  106. Hunter, H.N.; Bruce Fulton, D.; Ganz, T.; Vogel, H.J. The solution structure of human hepcidin, a peptide hormone with antimicrobial activity that is involved in iron uptake and hereditary hemochromatosis. J. Biol. Chem. 2002, 277, 37597–37603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Chaturvedi, P.; Bhat, R.A.H.; Pande, A. Antimicrobial Peptides of Fish: Innocuous Alternatives to Antibiotics. Rev. Aquac. 2020, 12, 85–106. [Google Scholar] [CrossRef]
  108. Mihailescu, M.; Sorci, M.; Seckute, J.; Silin, V.I.; Hammer, J.; Perrin, P.S., Jr.; Hernandez, J.I.; Smajic, N.; Shrestha, A.; Bogardus, K.A.; et al. structure and function in antimicrobial piscidins: Histidine position, directionality of membrane insertion, and pH-dependent permeabilization. J. Am. Chem. Soc. 2019, 141, 9837–9853. [Google Scholar] [CrossRef] [PubMed]
  109. Van Hoek, ML Antimicrobial Peptides in Reptiles. Pharmaceuticals 2014, 7, 723–753. [CrossRef] [Green Version]
  110. Cheng, Y.; Prickett, M.D.; Gutowska, W.; Kuo, R.; Belov, K.; Burt, D.W. Evolution of the avian β-defensin and cathelicidin genes. BMC Evol. Biol. 2015, 15, 188. [Google Scholar] [CrossRef] [Green Version]
  111. Ageitos, J.M.; Sánchez-Pérez, A.; Calo-Mata, P.; Villa, T.G. Antimicrobial peptides (amps): Ancient compounds that represent novel weapons in the fight against bacteria. Biochem. Pharmacol. 2017, 133, 117–138. [Google Scholar] [CrossRef]
  112. Wang, J.; Dou, X.; Song, J.; Lyu, Y.; Zhu, X.; Xu, L.; Li, W.; Shan, A. Antimicrobial peptides: Promising alternatives in the post-feeding antibiotic era. Med. Res. Rev. 2019, 39, 831–859. [Google Scholar] [CrossRef]
  113. Chen, C.H.; Lu, T.K. Development and challenges of antimicrobial peptides for therapeutic applications. Antibiotics 2020, 9, 24. [Google Scholar] [CrossRef] [Green Version]
  114. Torres, M.D.T.; Sothiselvam, S.; Lu, T.K.; de la Fuente-Nunez, C. Peptide design principles for antimicrobial applications. J. Mol. Biol. 2019, 431, 3547–3567. [Google Scholar] [CrossRef] [PubMed]
  115. Lei, J.; Sun, L.; Huang, S.; Zhu, C.; Li, P.; He, J.; Mackey, V.; Coy, D.H.; He, Q. The antimicrobial peptides and their potential clinical applications. Am. J. Transl. Res. 2019, 11, 3919–3931. [Google Scholar]
  116. Luo, Y.; McLean, D.T.F.; Linden, G.J.; McAuley, D.F.; McMullan, R.; Lundy, F.T. The naturally occurring host defense peptide, LL-37, and its truncated mimetics KE-18 and KR-12 have selected biocidal and antibiofilm activities against Candida albicans, Staphylococcus aureus, and Escherichia coli in vitro. Front. Microbiol. 2017, 8, 544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Phambu, N.; Almarwani, B.; Garcia, A.M.; Hamza, N.S.; Muhsen, A.; Baidoo, J.E.; Sunda-Meya, A. Chain length effect on the structure and stability of antimicrobial peptides of the (RW) series. Biophys. Chem. 2017, 227, 8–13. [Google Scholar] [CrossRef] [PubMed]
  118. Tripathi, A.K.; Kumari, T.; Harioudh, M.K.; Yadav, P.K.; Kathuria, M.; Shukla, P.K.; Mitra, K.; Ghosh, J.K. Identification of GXXXXG motif in Chrysophsin-1 and its implication in the design of analogs with cell-selective antimicrobial and anti-endotoxin activities. Sci. Rep. 2017, 7, 3384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Wang, J.; Chou, S.; Xu, L.; Zhu, X.; Dong, N.; Shan, A.; Chen, Z. High specific selectivity and Membrane-Active Mechanism of the synthetic centrosymmetric α-helical peptides with Gly-Gly pairs. Sci. Rep. 2015, 5, 1–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. De Cândido, E.S.; Cardoso, M.H.S.; Sousa, D.A.; Viana, J.C.; de Oliveira-Júnior, N.G.; Miranda, V.; Franco, O.L. The use of versatile plant antimicrobial peptides in agribusiness and human health. Peptides 2014, 55, 65–78. [Google Scholar] [CrossRef]
  121. Ilic, N.; Novkovic, M.; Guida, F.; Xhindoli, D.; Benincasa, M.; Tossi, A.; Juretic, D. Selective antimicrobial activity and mode of action of adepantins, glycine-rich peptide antibiotics based on anuran antimicrobial peptide sequences. Biochim. Biophys. Acta 2013, 1828, 1004–1012. [Google Scholar] [CrossRef] [Green Version]
  122. Leite, N.B.; da Costa, L.C.; Dos Santos Alvares, D.; Dos Santos Cabrera, M.P.; de Souza, B.M.; Palma, M.S.; Ruggiero Neto, J. The effect of acidic residues and amphipathicity on the lytic activities of mastoparan peptides studied by fluorescence and CD spectroscopy. Amino Acids 2011, 40, 91–100. [Google Scholar] [CrossRef]
  123. Imjongjirak, C.; Amphaiphan, P.; Charoensapsri, W.; Amparyup, P. Characterization and antimicrobial evaluation of SpPR-AMP1, a proline-rich antimicrobial peptide from the mud crab Scylla paramamosain. Dev. Comp. Immunol. 2017, 74, 209–216. [Google Scholar] [CrossRef]
  124. Li, W.; Tailhades, J.; O’Brien-Simpson, N.M.; Separovic, F.; Otvos, L., Jr.; Hossain, M.A.; Wade, J.D. Proline-rich antimicrobial peptides: Potential therapeutics against antibiotic-resistant bacteria. Amino Acids 2014, 46, 2287–2294. [Google Scholar] [CrossRef] [PubMed]
  125. Sonderegger, C.; Fizil, Á.; Burtscher, L.; Hajdu, D.; Muñoz, A.; Gáspári, Z.; Read, N.D.; Batta, G.; Marx, F. D19S mutation of the cationic, cysteine-rich protein PAF: Novel insights into its structural dynamics, thermal unfolding and antifungal function. PLoS ONE 2017, 12, e0169920. [Google Scholar]
  126. Mohanram, H.; Bhattacharjya, S. Cysteine deleted protegrin-1 (cdp-1): Antibacterial activity, outer-membrane disruption and selectivity. Biochim. Biophys. Acta 2014, 1840, 3006–3016. [Google Scholar] [CrossRef] [PubMed]
  127. Chou, S.L.; Shao, C.X.; Wang, J.J.; Shan, A.S.; Xu, L.; Dong, N.; Li, Z.Y. Short, multiple-stranded β-hairpin peptides have antimicrobial potency with high selectivity and salt resistance. Acta Biomater. 2016, 30, 78–93. [Google Scholar] [CrossRef]
  128. Lee, E.; Shin, A.; Jeong, K.W.; Jin, B.; Jnawali, H.N.; Shin, S.; Shin, S.Y.; Kim, Y. Role of phenylalanine and valine (10) residues in the antimicrobial activity and cytotoxicity of piscidin-1. PLoS ONE 2014, 9, e114453. [Google Scholar] [CrossRef] [Green Version]
  129. Tripathi, A.K.; Kumari, T.; Tandon, A.; Sayeed, M.; Afshan, T.; Kathuria, M.; Shukla, P.K.; Mitra, K.; Ghosh, J.K. Selective phenylalanine to proline substitution for improved antimicrobial and anticancer activities of peptides designed on phenylalanine heptad repeat. Acta Biomater. 2017, 57, 170–186. [Google Scholar] [CrossRef]
  130. Brogden, K.A.; Ackermann, M.; Huttner, K.M. Small, anionic, and charge-neutralizing propeptide fragments of zymogens are antimicrobial. Antimicrob. Agents Chemother. 1997, 41, 1615–1617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Mojsoska, B.; Jenssen, H. Peptides and Peptidomimetics for Antimicrobial Drug Design. Pharmaceuticals 2015, 8, 366–415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Epand, R.M.; Vogel, H.J. Diversity of antimicrobial peptides and their mechanisms of action. Biochim. Biophys. Acta 1999, 1462, 11–28. [Google Scholar] [CrossRef] [Green Version]
  133. Padmanabhan, S.; York, E.J.; Stewart, J.M.; Baldwin, R.L. Helix propensities of basic amino acids increase with the length of the side-chain. J. Mol. Biol. 1996, 257, 726–734. [Google Scholar] [CrossRef] [Green Version]
  134. Pace, C.N.; Scholtz, J.M. A helix propensity scale based on experimental studies of peptides and proteins. Biophys. J. 1998, 75, 422–427. [Google Scholar] [CrossRef] [Green Version]
  135. Schmidtchen, A.; Pasupuleti, M.; Malmsten, M. Effect of hydrophobic modifications in antimicrobial peptides. Adv. Colloid Interface Sci. 2014, 205, 265–274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. 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] [PubMed]
  137. 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] [PubMed] [Green Version]
  138. Hädicke, A.; Blume, A. Binding of cationic peptides (KX) 4K to DPPG bilayers. Increasing the hydrophobicity of the uncharged amino acid X drives formation of membrane bound β-sheets: A DSC and FT-IR study. Biochim. Biophys. Acta (BBA) Biomembr. 2016, 1858, 1196–1206. [Google Scholar] [CrossRef]
  139. Wood, S.J.; Park, Y.A.; Kanneganti, N.P.; Mukkisa, H.R.; Crisman, L.L.; Davis, S.E. Modified cysteine-deleted tachyplesin (CDT) analogs as linear antimicrobial peptides: Influence of chain length, positive charge, and hydrophobicity on antimicrobial and hemolytic activity. Int. J. Pept. Res. Ther. 2014, 20, 519–530. [Google Scholar] [CrossRef]
  140. Wang, J.; Chou, S.; Yang, Z.; Yang, Y.; Wang, Z.; Song, J.; Dou, X.; Shan, A. Combating drug-resistant fungi with novel imperfectly amphipathic palindromic peptides. J. Med. Chem. 2018, 61, 3889–3907. [Google Scholar] [CrossRef]
  141. 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] [Green Version]
  142. Jakel, C.E.; Meschenmoser, K.; Kim, Y.; Weiher, H.; Schmidt-Wolf, I.G. Efficacy of a proapoptotic peptide towards cancer cells. In Vivo 2012, 26, 419–426. [Google Scholar]
  143. Shai, Y. Mode of action of membrane active antimicrobial peptides. Biopolymers 2002, 66, 236–248. [Google Scholar] [CrossRef]
  144. Lee, A.C.-L.; Harris, J.L.; Khanna, K.K.; Hong, J.-H. A Comprehensive Review on Current Advances in Peptide Drug Development and Design. Int. J. Mol. Sci. 2019, 20, 2383. [Google Scholar] [CrossRef] [Green Version]
  145. Ebenhan, T.; Gheysens, O.; Kruger, H.G.; Zeevaart, J.R.; Sathekge, M.M. Antimicrobial peptides: Their role as infection-selective tracers for molecular imaging. BioMed. Res. Int. 2014, 2014, 867381. [Google Scholar] [CrossRef] [Green Version]
  146. Fanelli, F.; Cozzi, G.; Raiola, A.; Dini, I.; Mulè, G.; Logrieco, A.F.; Ritieni, A. Raisins and currants as conventional nutraceuticals in Italian market: Natural occurrence of Ochratoxin A. J. Food Sci. 2017, 82, 2306–2312. [Google Scholar] [CrossRef] [PubMed]
  147. Zasloff, M. Antimicrobial peptides of multicellular organisms. Nature 2002, 415, 389–395. [Google Scholar] [CrossRef] [PubMed]
  148. McPhee, J.B.; Hancock, R.E. Function and therapeutic potential of host defence peptides. J. Pept. Sci 2005, 11, 677–687. [Google Scholar] [CrossRef]
  149. Bahar, A.A.; Ren, D. Antimicrobial peptides. Pharmaceuticals 2013, 6, 1543–1575. [Google Scholar] [CrossRef] [Green Version]
  150. Shai, Y.; Oren, Z. From “carpet” mechanism to de-novo designed diastereomeric cell-selective antimicrobial peptides. Peptides 2001, 22, 1629–1641. [Google Scholar] [CrossRef]
  151. Cruciani, R.A.; Barker, J.L.; Durell, S.R.; Raghunathan, G.; Robert Guy, H.; Zasloff, M.; Stanley, E.F. Magainin 2, a natural antibiotic from frog skin, forms ion channels in lipid bilayer membranes. Eur. J. Pharmacol. Mol. Pharmacol. 1992, 226, 287–296. [Google Scholar] [CrossRef]
  152. Sengupta, D.; Leontiadou, H.; Mark, A.E.; Marrink, S.J. Toroidal pores formed by antimicrobial peptides show significant disorder. Biochim. Biophys. Acta Biomembr. 2008, 1778, 2308–2317. [Google Scholar] [CrossRef] [Green Version]
  153. 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]
  154. Rapaport, D.; Shai, Y. Interaction of fluorescently labeled pardaxin and its analogues with lipid bilayers. J. Biol. Chem. 1991, 266, 23769–23775. [Google Scholar] [CrossRef]
  155. Yang, L.; Harroun, T.A.; Weiss, T.M.; Ding, L.; Huang, H.W. Barrel-stave model or toroidal model? A case study on melittin pores. Biophys. J. 2001, 81, 1475–1485. [Google Scholar] [CrossRef] [Green Version]
  156. Leontiadou, H.; Mark, A.E.; Marrink, S.J. Antimicrobial peptides in action. J. Am. Chem. Soc. 2006, 128, 12156–12161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. 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]
  158. Zhang, L.; Rozek, A.; Hancock, R.E.W. Interaction of Cationic Antimicrobial Peptides with Model Membranes. J. Biol. Chem. 2001, 276, 35714–35722. [Google Scholar] [CrossRef] [Green Version]
  159. Park, C.B.; Kim, M.S.; Kim, S.C. A novel antimicrobial peptide from Bufo bufo gargarizans. Biochem. Biophys. Res. Commun. 1996, 218, 408–413. [Google Scholar] [CrossRef]
  160. Park, C.B.; Kim, H.S.; Kim, S.C. Mechanism of action of the antimicrobial peptide buforin II: Buforin II kills microorganisms by penetrating the cell membrane and inhibiting cellular functions. Biochem. Biophys. Res. Commun. 1998, 244, 253–257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Hancock, R.E.W.; Nijnik, A.; Philpott, D.J. Modulating immunity as a therapy for bacterial infections. Nat. Rev. Microbiol. 2012, 10, 243–254. [Google Scholar] [CrossRef] [PubMed]
  162. Yeung, A.T.; Gellatly, S.L.; Hancock, R.E. Multifunctional cationic host defence peptides and their clinical applications. Cell Mol Life Sci 2011, 68, 2161–2176. [Google Scholar] [CrossRef]
  163. Hilchie, A.L.; Wuerth, K.; Hancock, R.E.W. Immune modulation by multifaceted cationic host defense (antimicrobial) peptides. Nat. Chem. Biol. 2013, 9, 761–768. [Google Scholar] [CrossRef]
  164. Lai, Y.; Gallo, R.L. AMPed up immunity: How antimicrobial peptides have multiple roles in immune defense. Trends Immunol. 2009, 30, 131–141. [Google Scholar] [CrossRef] [Green Version]
  165. Jiao, K.; Gao, J.; Zhou, T.; Yu, J.; Song, H.; Wei, Y.; Gao, X. Isolation and purification of a novel antimicrobial peptide from Porphyra yezoensis. J. Food Biochem. 2019, 43, e12864. [Google Scholar] [CrossRef] [PubMed]
  166. Keymanesh, K.; Soltani, S.; Sardari, S. Application of antimicrobial peptides in agriculture and food industry. World J. Microbiol. Biotechnol. 2009, 25, 933–944. [Google Scholar] [CrossRef]
  167. Naghmouchi, K.; Belguesmia, Y.; Bendali, F.; Spano, G.; Seal, B.S.; Drider, D. Lactobacillus fermentum: A bacterial species with potential for food preservation and biomedical applications. Crit. Rev. Food Sci. Nutr. 2019, 1, 3387–3399. [Google Scholar] [CrossRef] [PubMed]
  168. Schmitt, P.; Rosa, R.D.; Destoumieux-Garzon, D. An intimate link between antimicrobial peptide sequence diversity and binding to essential components of bacterial membranes. Biochim. Biophys. Acta Biomembr. 2016, 1858, 958–970. [Google Scholar] [CrossRef]
  169. Lima, K.O.; da Costa de Quadros, C.; Rocha, M.d.; Jocelino Gomes de Lacerda, J.T.; Juliano, M.A.; Dias, M.; Mendes, M.A.; Prentice, C. Bioactivity and bioaccessibility of protein hydrolyzates from industrial byproducts of Stripped weakfish (Cynoscion guatucupa). LWT 2019, 111, 408–413. [Google Scholar] [CrossRef]
  170. Liu, Y.; Du, Q.; Ma, C.; Xi, X.; Wang, L.; Zhou, M.; Burrows, J.F.; Chen, T.; Wang, H. Structure–activity relationship of an antimicrobial peptide, Phylloseptin-PHa: Balance of hydrophobicity and charge determines the selectivity of bioactivities. Drug Des. Devel. Ther. 2019, 13, 447–458. [Google Scholar] [CrossRef] [Green Version]
  171. Gddoa Al-sahlany, S.T.; Altemimi, A.B.; Abd Al Manhel, A.J.; Niamah, A.K.; Lakhssasi, N.; Ibrahim, S.A. Purification et bioactive peptide with antimicrobial properties produced by Saccharomyces cerevisiae. Foods 2020, 9, 324. [Google Scholar] [CrossRef] [Green Version]
  172. Tavano, O.L. Protein hydrolysis using proteases: An important tool for food biotechnology. J. Mol. Catal. B Enzym. 2013, 90, 1–11. [Google Scholar] [CrossRef]
  173. Mao, Y.; Niu, S.; Xu, X.; Wang, J.; Su, Y.; Wu, Y.; Zhong, S. The effect of an adding histidine on biological activity and stability of pc-pis from Pseudosciaena crocea. PLoS ONE 2013, 8, e83268. [Google Scholar] [CrossRef] [Green Version]
  174. Shruti, S.R.; Rajasekaran, R. Identification of protegrin-1 as a stable and nontoxic scaffold among protegrin family—A computational approach. J. Biomol. Struct. Dyn. 2019, 37, 2430–2439. [Google Scholar] [CrossRef]
  175. Van Lancker, F.; Adams, A.; De Kimpe, N. Chemical modifications of peptides and their impact on food properties. Chem. Rev. 2011, 111, 7876–7903. [Google Scholar] [CrossRef] [PubMed]
  176. Berardo, A.; De Maere, H.; Stavropoulou, D.A.; Rysman, T.; Leroy, F.; De Smet, S. Effect of sodium ascorbate and sodium nitrite on protein and lipid oxidation in dry fermented sausages. Meat Sci. 2016, 121, 359–364. [Google Scholar] [CrossRef] [PubMed]
  177. Peña-Egido, M.J.; García-Alonso, B.; García-Moreno, C. S-sulfonate contents in raw and cooked meat products. JAFC 2005, 53, 4198–4201. [Google Scholar] [CrossRef] [PubMed]
  178. Sun, X.; Acquah, C.; Aluko, R.E.; Udenigwe, C.C. Considering food matrix and gastrointestinal effects in enhancing bioactive peptide absorption and bioavailability. J. Funct. Foods 2020, 64, 103680. [Google Scholar] [CrossRef]
  179. Sarabandi, K.; Gharehbeglou, P.; Jafari, S.M. Spray-drying encapsulation of protein hydrolysates and bioactive peptides: Opportunities and challenges. Dry. Technol. 2020, 38, 577–595. [Google Scholar] [CrossRef]
  180. Yekta, M.M.; Rezaei, M.; Nouri, L.; Azizi, M.H.; Jabbari, M.; Eş, I.; Khaneghah, A.M. Antimicrobial and antioxidant properties of burgers with quinoa peptide-loaded nanoliposomes. J. Food Saf. 2020, 40, e12753. [Google Scholar] [CrossRef]
  181. Soto, K.M.; Hernandez-Iturriaga, M.; Loarca-Pina, G.; Luna-Barcenas, G.; Gomez-Aldapa, C.A.; Mendoza, S. Stable nisin food-grade electrospun fibers. J. Food Sci. Technol. 2016, 53, 3787–3794. [Google Scholar] [CrossRef] [Green Version]
  182. Wu, X.; Wei, P.H.; Zhu, X.; Wirth, M.J.; Bhunia, A.; Narsimhan, G. Effect of immobilization on the antimicrobial activity of a cysteine-terminated antimicrobial Peptide Cecropin P1 tethered to silica nanoparticle against E. coli O157:H7 EDL. Colloids Surf. B Biointerfaces 2017, 156, 305–312. [Google Scholar] [CrossRef]
  183. Yi, L.; Qi, T.; Ma, J.; Zeng, K. Genome and metabolites analysis reveal insights into control of foodborne pathogens in fresh-cut fruits by Lactobacillus pentosus MS031 isolated from Chinese sichuan paocai. Postharvest Biol. Technol. 2020, 164, 111150. [Google Scholar] [CrossRef]
  184. Tenea, G.N.; Pozo, T.D. Antimicrobial peptides from Lactobacillus plantarum UTNGt2 prevent harmful bacteria growth on fresh tomatoes. J. Microbiol. Biotechnol. 2019, 29, 1553–1560. [Google Scholar] [CrossRef]
  185. Gogliettino, M.; Balestrieri, M.; Ambrosio, R.L.; Anastasio, A.; Smaldone, G.; Proroga, Y.T.R.; Moretta, R.; Rea, I.; De Stefano, L.; Agrillo, B.; et al. Extending the Shelf-Life of Meat and Dairy Products via PET-Modified Packaging Activated with the Antimicrobial Peptide MTP1. Front. Microbiol. 2020, 10, 2963. [Google Scholar] [CrossRef] [PubMed]
  186. Dos Santos Pires, A.C.; De Ferreira Soares, N.F.; De Andrade, N.J.; Mendes Da Silva, L.H.; Peruch Camilloto, G.; Campos Bernardes, P. Development and evaluation of active packaging for sliced mozzarella preservation. Packag. Technol. Sci. 2008, 21, 375–383. [Google Scholar] [CrossRef]
  187. Da Rocha, M.; Alemán, A.; Romani, V.P.; López-Caballero, M.E.; Gómez-Guillén, M.C.; Montero, P.; Prentice, C. Effects of agar films incorporated with fish protein hydrolysate or clove essential oil on flounder (Paralichthys orbignyanus) fillets shelf-life. Food Hydrocoll. 2018, 81, 351–363. [Google Scholar] [CrossRef]
  188. Neetoo, H. Use of nisin-coated plastic films to control Listeria monocytogenes on vacuum-packaged cold-smoked salmon. Int. J. Food Microbiol. 2008, 122, 8–15. [Google Scholar] [CrossRef]
  189. Luo, L.; Wu, Y.; Liu, C.; Zou, Y.; Huang, L.; Liang, Y.; Ren, J.; Liu, Y.; Lin, Q. Elaboration and characterization of curcumin-loaded soy soluble polysaccharide (SSPS)-based nanocarriers mediated by antimicrobial peptide nisin. Food Chem. 2021, 336, 127669. [Google Scholar] [CrossRef] [PubMed]
  190. Zohri, M.; Shafiee, M.; Ismaeil, A. A comparative study between the antibacterial effect of nisin and nisin-loaded chitosan/alginate nanoparticles on the growth of Staphylococcus aureus in raw and pasteurized milk samples. Probiotics Antimicrob. Proteins 2010, 2, 258–266. [Google Scholar] [CrossRef]
  191. Piras, A.M.; Maisetta, G.; Sandreschi, S.; Gazzarri, M.; Bartoli, C.; Grassi, L.; Esin, S.; Chiellini, F.; Batoni, G. Chitosan nanoparticles loaded with the antimicrobial peptide temporin B exert a long-term antibacterial activity in vitro against clinical isolates of Staphylococcus epidermidis. Front. Microbiol. 2015, 6, 372. [Google Scholar] [CrossRef] [Green Version]
  192. Jiang, Y.; Sun, J.; Yin, Z.; Li, H. Evaluation of antioxidant peptides generated from Jiuzao (residue after Baijiu distillation) protein hydrolysates and their effect of enhancing healthy value of Chinese Baijiu. Soc. Chem. Ind. 2019, 100, 59–73. [Google Scholar] [CrossRef]
  193. Najafian, L.; Babji, A.S. Purification and Identification of Antioxidant Peptides from Fermented Fish Sauce (Budu). J. Aquat. Food Prod. Technol. 2018, 8850, 14–24. [Google Scholar] [CrossRef]
  194. Wu, D.; Sun, N.; Ding, J.; Zhu, B.; Lin, S. Evaluation and structure-activity relationship analysis of antioxidant shrimp peptides. Food Funct. 2019, 10, 5605–5615. [Google Scholar] [CrossRef]
  195. Marqus, S.; Pirogova, E.; Piva, T.J. Evaluation of the use of therapeutic peptides for cancer treatment. J. Biomed. Sci. 2017, 24, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Boohaker, R.J.; Lee, M.W.; Vishnubhotla, P.; Perez, J.M.; Khaled, A.R. The use of therapeutic peptides to target and to kill cancer cells. Curr. Med. Chem. 2012, 19, 3794–3804. [Google Scholar] [CrossRef] [PubMed]
  197. McGregor, D.P. Discovering and improving novel peptide therapeutics. Curr. Opin. Pharmacol. 2008, 8, 616–619. [Google Scholar] [CrossRef] [PubMed]
  198. 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]
  199. Di Grazia, A.; Cappiello, F.; Cohen, H.; Casciaro, B.; Luca, V.; Pini, A.; Di, Y.P.; Shai, Y.; Mangoni, M.L. d-Amino acids incorporation in the frog skin-derived peptide esculentin-1a(1–21)NH2 is beneficial for its multiple functions. Amino Acids 2015, 47, 2505–2519. [Google Scholar] [CrossRef]
  200. Cappiello, F.; Di Grazia, A.; Segev-Zarko, L.A.; Scali, S.; Ferrera, L.; Galietta, L.; Pini, A.; Shai, Y.; Di, Y.P.; Mangoni, M.L. Esculentin-1a-derived peptides promote clearance of Pseudomonas aeruginosa internalized in bronchial cells of cystic fibrosis patients and lung cell migration: Biochemical properties and a plausible mode of action. Antimicrob. Agents Chemother. 2016, 60, 7252–7262. [Google Scholar] [CrossRef] [Green Version]
  201. Pero, R.; Brancaccio, M.; Mennitti, C.; Gentile, L.; Franco, A.; Laneri, S.; De Biasi, M.G.; Pagliuca, C.; Colicchio, R.; Salvatore, P.; et al. HNP-1 and HBD-1 as Biomarkers for the Immune Systems of Elite Basketball Athletes. Antibiotics 2020, 9, 306. [Google Scholar] [CrossRef]
  202. Zhang, P.; Chang, C.; Liu, H.; Li, B.; Yan, Q.; Jiang, Z. Identification of novel angiotensin I-converting enzyme (ACE) inhibitory peptides from wheat gluten hydrolysate by the protease of Pseudomonas aeruginosa. J. Funct. Foods 2020, 65, 103751. [Google Scholar] [CrossRef]
  203. Wu, C.; Mohammadmoradi, S.; Chen, J.Z.; Sawada, H.; Daugherty, A.; Lu, H.S. Renin-Angiotensin System and Cardiovascular Functions. Arterosclerosis. Thromb. Vasc. Biol. 2018, 38, 108–116. [Google Scholar] [CrossRef] [Green Version]
  204. Martinez-Villaluenga, C.; Rupasinghe, S.G.; Schuler, M.A.; Gonzalez de Mejia, E. Peptides from purified soybean beta-conglycinin inhibit fatty acid synthase by interaction with the thioesterase catalytic domain. FEBS J. 2010, 277, 1481–1493. [Google Scholar] [CrossRef]
  205. Hendrikx, T.; Schnabl, B. Antimicrobial proteins: Intestinal guards to protect against liver disease. J. Gastroenterol. 2019, 54, 209–217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. McCann, K.B.; Lee, A.; Wan, J.; Roginski, H.; Coventry, M.J. The effect of bovine lactoferrin and lactoferricin B on the ability of feline calicivirus (a norovirus surrogate) and poliovirus to infect cell cultures. J. Appl. Microbiol. 2003, 95, 1026–1033. [Google Scholar] [CrossRef] [PubMed]
  207. Pietrantoni, A.; Ammendolia, M.G.; Tinari, A.; Siciliano, R.; Valenti, P.; Superti, F. Bovine lactoferrin peptidic fragments involved in inhibition of Echovirus 6 in vitro infection. Antivir. Res. 2006, 69, 98–106. [Google Scholar] [CrossRef] [PubMed]
  208. Belaid, A.; Aouni, M.; Khelifa, R.; Trabelsi, A.; Jemmali, M.; Hani, K. In vitro antiviral activity of dermaseptins against herpes simplex virus type 1. J. Med. Virol. 2002, 66, 229–234. [Google Scholar] [CrossRef]
  209. Mettenleiter, T.C. Brief overview on cellular virus receptors. Virus Res. 2002, 82, 3–8. [Google Scholar] [CrossRef]
  210. Elnagdy, S.; AlKhazindar, M. The Potential of Antimicrobial Peptides as an Antiviral Therapy against COVID-19. ACS Pharm. Transl. Sci. 2020, 3, 780–782. [Google Scholar] [CrossRef]
  211. Kudryashova, E.; Zani, A.; Vilmen, G.; Sharma, A.; Lu, W.; Yount, J.S.; Kudryashov, D.S. SARS-CoV-2 incativation by human defensin HNP1 and retrocyclin RC-101. bioRxiv 2021. [Google Scholar] [CrossRef]
  212. Brancaccio, M.; Mennitti, C.; Calvanese, M.; Gentile, A.; Musto, R.; Gaudiello, G.; Scamardella, G.; Terracciano, D.; Frisso, G.; Pero, R.; et al. Diagnostic and Therapeutic Potential for HNP-1, HBD-1 and HBD-4 in Pregnant Women with COVID-19. Int. J. Mol. Sci. 2022, 23, 3450. [Google Scholar] [CrossRef]
  213. Laneri, S.; Brancaccio, M.; Mennitti, C.; De Biasi, M.G.; Pero, M.E.; Pisanelli, G.; Scudiero, O.; Pero, R. Antimicrobial Peptides and Physical Activity: A Great Hope against COVID 19. Microorganisms 2021, 9, 1415. [Google Scholar] [CrossRef]
  214. Chianese, A.; Zannella, C.; Monti, A.; De Filippis, A.; Doti, N.; Franci, G.; Galdiero, M. The Broad-Spectrum Antiviral Potential of the Amphibian Peptide AR-23. Int. J. Mol. Sci. 2022, 23, 883. [Google Scholar] [CrossRef]
  215. Zannella, C.; Chianese, A.; Palomba, L.; Marcocci, M.E.; Bellavita, R.; Merlino, F.; Grieco, P.; Folliero, V.; De Filippis, A.; Mangoni, M.; et al. Broad-Spectrum Antiviral Activity of the Amphibian Antimicrobial Peptide Temporin L and Its Analogs. Int. J. Mol. Sci. 2022, 23, 2060. [Google Scholar] [CrossRef] [PubMed]
  216. Ramos, R.; Silva, J.P.; Rodrigues, A.C.; Costa, R.; Guardao, L.; Schmitt, F.; Soares, R.; Vilanova, M.; Domingues, L.; Gama, M. Wound healing activity of the human antimicrobial peptide ll37. Peptides 2011, 32, 1469–1476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Simanski, M.; Gläser, R.; Harder, J. Human skin engages different epidermal layers to provide distinct innate defense mechanisms. Exper. Dermat. 2014, 23, 230–231. [Google Scholar] [CrossRef]
  218. Schröder, J.M.; Harder, J. Antimicrobial skin peptides and proteins. Cell. Mol. Life Sci. 2006, 63, 469–486. [Google Scholar] [CrossRef] [PubMed]
  219. Mangoni, M.L.; McDermott, A.M.; Zasloff, M. Antimicrobial peptides and wound healing: Biological and therapeutic considerations. Exp. Dermatol. 2016, 25, 167–173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. El-Seedi, H.; Abd El-Wahed, A.; Yosri, N.; Musharraf, S.G.; Chen, L.; Moustafa, M.; Zou, X.; Al-Mousawi, S.; Guo, Z.; Khatib, A.; et al. Antimicrobial Properties of Apis mellifera’s Bee Venom. Toxins 2020, 12, 451. [Google Scholar] [CrossRef]
Figure 1. Several ways to classify antimicrobial peptides (AMPs).
Figure 1. Several ways to classify antimicrobial peptides (AMPs).
Antibiotics 11 01483 g001
Figure 2. Classification of the AMPs produced by Gram-positive bacteria.
Figure 2. Classification of the AMPs produced by Gram-positive bacteria.
Antibiotics 11 01483 g002
Figure 3. Classification of the AMPs produced by Gram-negative bacteria.
Figure 3. Classification of the AMPs produced by Gram-negative bacteria.
Antibiotics 11 01483 g003
Figure 4. Classification of the AMPs produced by fungi.
Figure 4. Classification of the AMPs produced by fungi.
Antibiotics 11 01483 g004
Figure 5. Classification of the AMPs produced by plants.
Figure 5. Classification of the AMPs produced by plants.
Antibiotics 11 01483 g005
Figure 6. The AMP mechanisms of action on cell membranes.
Figure 6. The AMP mechanisms of action on cell membranes.
Antibiotics 11 01483 g006
Table 1. Animal AMPs.
Table 1. Animal AMPs.
AnimalsAMPs
Mammalianscathelicidins
defensins (α-, β-, and θ-defensins; θ-defensins are not expressed in adult humans)
platelet antimicrobial proteins
dermicidins
hepcidins
Reptilesdefensins (α, β-, and θ-defensins)
cathelicidins
Fishβ-defensins
cathelicidins
hepicidins (HAMP1 and HAMP2)
histone-derived peptides
piscidins (piscidins 1–7)
Amphibiansmagainins
cancrins
Crustaceanscrustins
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dini, I.; De Biasi, M.-G.; Mancusi, A. An Overview of the Potentialities of Antimicrobial Peptides Derived from Natural Sources. Antibiotics 2022, 11, 1483. https://doi.org/10.3390/antibiotics11111483

AMA Style

Dini I, De Biasi M-G, Mancusi A. An Overview of the Potentialities of Antimicrobial Peptides Derived from Natural Sources. Antibiotics. 2022; 11(11):1483. https://doi.org/10.3390/antibiotics11111483

Chicago/Turabian Style

Dini, Irene, Margherita-Gabriella De Biasi, and Andrea Mancusi. 2022. "An Overview of the Potentialities of Antimicrobial Peptides Derived from Natural Sources" Antibiotics 11, no. 11: 1483. https://doi.org/10.3390/antibiotics11111483

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