Next Article in Journal
Toxicological Studies of 212Pb Intravenously or Intraperitoneally Injected into Mice for a Phase 1 Trial
Next Article in Special Issue
Antifungal Activity of 14-Helical β-Peptides against Planktonic Cells and Biofilms of Candida Species
Previous Article in Journal
Effects of a Formula Containing Two Types of Prebiotics, Bifidogenic Growth Stimulator and Galacto-oligosaccharide, and Fermented Milk Products on Intestinal Microbiota and Antibody Response to Influenza Vaccine in Elderly Patients: A Randomized Controlled Trial
Previous Article in Special Issue
The Fungal Defensin Family Enlarged
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Peptides and Peptidomimetics for Antimicrobial Drug Design

Department of Science, Systems and Models, Roskilde University, Universitetsvej 1, Postboks 260, Roskilde 4000, Denmark
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2015, 8(3), 366-415; https://doi.org/10.3390/ph8030366
Submission received: 26 March 2015 / Revised: 27 May 2015 / Accepted: 17 June 2015 / Published: 13 July 2015

Abstract

:
The purpose of this paper is to introduce and highlight a few classes of traditional antimicrobial peptides with a focus on structure-activity relationship studies. After first dissecting the important physiochemical properties that influence the antimicrobial and toxic properties of antimicrobial peptides, the contributions of individual amino acids with respect to the peptides antibacterial properties are presented. A brief discussion of the mechanisms of action of different antimicrobials as well as the development of bacterial resistance towards antimicrobial peptides follows. Finally, current efforts on novel design strategies and peptidomimetics are introduced to illustrate the importance of antimicrobial peptide research in the development of future antibiotics.

1. Introduction

Even though humans have adapted to live in harmony with different microorganisms throughout evolution, this balanced symbiotic relationship can sometimes shift and allow pathogenic bacteria to blossom and cause infections. In the struggle for survival, a complex mechanism involving many key components assists in the elimination of these infectious agents. Antimicrobial peptides are conserved biomolecules among all living species, including bacteria that take part in the battle against the invading pathogens. They are relatively short (<100 amino acid residues), positively charged, amphipathic (have both hydrophobic and hydrophilic domains) and exhibit diversity based on their structural properties [1]. Despite the structural difference and myriad of sequences incorporating both natural and unnatural amino acids, antimicrobial peptides exhibit broad spectrum antibacterial activities. The ability of antimicrobial peptides to kill or inhibit the growth of bacteria has attracted the attention of many research groups worldwide to study mechanism(s) of their antimicrobial action. Their amphipathic nature is believed to allow them to interact with the negatively charged structures and hydrophobic fatty acid chains found on the target microbial membranes, leading to membrane destabilization and apparently cell lysis [2]. Generally AMPs are classified in four large families based on their secondary conformations in α-helices, β-sheets, mixed structures and non- α- or β-structures (extended) [3] (Figure 1).
Figure 1. Four structural classes of antimicrobial peptides. (A) α-helical structure of human cathelicidin LL-37 (PDB code 2K6O [4]); (B) β-sheeted polyphemusin (PDB code 1RKK [5]); (C) extended indolicidin (PDB code 1G89 [6]); (D) and mixed structures like human β-defensin-2 (PDB code 1FQQ [7]).
Figure 1. Four structural classes of antimicrobial peptides. (A) α-helical structure of human cathelicidin LL-37 (PDB code 2K6O [4]); (B) β-sheeted polyphemusin (PDB code 1RKK [5]); (C) extended indolicidin (PDB code 1G89 [6]); (D) and mixed structures like human β-defensin-2 (PDB code 1FQQ [7]).
Pharmaceuticals 08 00366 g001
These structures are predominantly present upon interaction with lipid membranes. Most antimicrobial peptides from both multi- and unicellular organisms are derived from precursor sequences. Consequently, they display a number of post-translational modifications that largely modify their activity. These include: proteolytic processing, glycosylation, amidation, halogenation, phosphorylation, incorporation of unnatural d-amino acids and cyclization [8]. Some AMPs are also synthesized non-ribosomely e.g., gramicidin S and lipopeptides. Antimicrobial peptides have also been intensively modified through synthetic chemistry in order to meet the requirements of potential therapeutic drugs, thus further increasing the structural diversity. Most recently, combinations of different structures such as β-peptides, peptoids, β-peptoids, peptide-peptoid hybrids and other, have been synthesized and the antimicrobial activity of the resulting peptidomimetics has been compared to that of traditional antimicrobial peptides. This extensive research holds vast potential in dissecting the detailed mechanism by which both innate and novel antimicrobial compounds can be used to fight pathogenic infections. In addition, there is a remarkable interest for development of new antimicrobials to combat infectious diseases in parallel with the increasing evidence of their broad spectrum activity. Until now more than 2000 antimicrobial peptides (also referred to as host defense peptides) have been isolated from various cells and tissues of animals, insects, plants and bacteria (http://aps.unmc.edu/AP/main.php) [9]. Well known examples of antimicrobial peptides belong to the families of the cathelicidins and defensins (found in many insects and plants and animals, including humans), thionins (isolated from plants), cecropins (found in the hemolymph of the cecropia silk moth in the early 80s) and magainins (secreted from frog skin) [10,11]. This review will focus on these peptide groups, discuss their antibacterial mode of action in parallel with their structural characteristics, and then further elute to the potential of peptidomimetic as novel antimicrobial agents and how they can be designed based on our current knowledge on antimicrobial peptides.

2. Antimicrobial Peptides Isolated from Mammals

In mammals, antimicrobial peptides have been isolated from different sources such as the granules of neutrophils, Peneth cells, mucosal secretions from epithelial cells and as protein degradation products [12]. Three classes of antimicrobial peptides, found in abundance in neutrophils defensins, cathelicidins and histatins [13,14,15] have been studied extensively by several groups. This section highlights some important research on defensins and cathelicidins.

2.1. Defensins

Defensins, first discovered in human neutrophils, are 18-45 amino acids long cationic molecules which are isolated from mast cells and tissues involved in host defense [16,17,18]. They have been classified as α-, β and θ-defensins based on structural differences (Figure 2). All three classes are rich in cysteine and arginine residues. The difference between the classes lies in the position of disulfide linked cysteine residues. Herein, the first class of defensins (α-defensins), have disulfide connections between cysteine residues 1-6, 2- and 3-5, whereas β-defensins disulfide bridges are located between cysteine residues 1-5, 2-4 and 3-6 (Table 1) [19]. Moreover, human α-defensins are less cationic, shorter and more hydrophobic than human β-defensins [20]. Human neutrophil peptides (HNP1-HNP4) and human defensins (HD-5 and HD-6) are six different human α-defensins found in monocytes, NK cells, B and T cells, neutrophils and Peneth cells, respectively [21,22]. Studies show that HNPs exhibit antimicrobial activity against both Gram-negative and Gram-positive bacteria [23]. Human defensin 5 exerts its antimicrobial activity against various bacteria such as Escherichia coli, Listeria monocytogenes, Salmonella typhimurium, Staphylococcus aureus and Vibrio cholerae [24]. Contrary to α-defensins, four human β-defensins (HBD 1-4) have been isolated from leukocytes and epithelial cells. The β-defensins also exhibit microbicidal activity against a panel of bacteria, with HBD-4 being the most potent against Pseudomonas aeruginosa [25]. It has been shown that epithelial human β-defensins are over-expressed in patients suffering from psoriasis, a chronic skin inflammation, with characteristic skin lesions cleared of infection. Conversely in atopic dermatitis where the expression of HBDs is suppressed, the lesions are infection-prone [26], clearly illustrating the importance of these antimicrobial peptides in host defense.

2.2. Cathelicidins

The cathelicidins are another class of antimicrobial peptides which have been identified in many vertebrates such as fish, bird, cow, pig, rabbit, sheep, mouse, monkey, horse and human [27,28,29]. They are primarily produced in epithelial cells, neutrophils and macrophages. Indolicidin and human cathelicidin LL-37 (Table 1) are two of the most studied members of this class of peptides. LL-37 is the only member of this group of peptides found in human. The importance of these peptides is elucidated in studies where increased susceptibility to infections is observed in patients deficient in neutrophil production of LL-37, the major source of human antimicrobial peptides [30].
Figure 2. Three structural classes of defensines. The PDB codes are all indicated, in addition to the primary sequence with the disulfide bridges indicated with numbers in subscript (A) Human alpha-defensin-6 (PDB code 1ZMQ) AFTC1HC2RRSC3YSTEYSYGTC2TVMGINHRFC3C1L [31]; (B) Human beta-defensin-1 (PDB code 1IJV) DHYNC1VSSGGQC2LYSAC3PIFTKIQGTC2YRGKAKC1C3K [32] and (C) rhesus theta defensin-1 (PDB code 1HVZ) GFC1RC2LC3RRGVC3RC2IC1TR [33].
Figure 2. Three structural classes of defensines. The PDB codes are all indicated, in addition to the primary sequence with the disulfide bridges indicated with numbers in subscript (A) Human alpha-defensin-6 (PDB code 1ZMQ) AFTC1HC2RRSC3YSTEYSYGTC2TVMGINHRFC3C1L [31]; (B) Human beta-defensin-1 (PDB code 1IJV) DHYNC1VSSGGQC2LYSAC3PIFTKIQGTC2YRGKAKC1C3K [32] and (C) rhesus theta defensin-1 (PDB code 1HVZ) GFC1RC2LC3RRGVC3RC2IC1TR [33].
Pharmaceuticals 08 00366 g002
Indolicidin is a small antimicrobial peptide with 13 amino acids isolated from bovine neutrophils (Figure 1 and Table 1). It is rich in tryptophan (39%) and arginine (23%) residues, and it is amidated at the C-terminal arginine [34,35]. The antimicrobial activity towards both Gram-negative and Gram-positive bacteria, fungi and protozoa is most likely not a result of any confined secondary structure, as nuclear magnetic resonance spectroscopy studies failed to report defined secondary or amphipathic structure which is characteristic for majority of antimicrobial peptides [6,36]. However the hydrophobic and positively charged domains of indolicidin are crucial for its interactions with bacterial pathogens. In terms of mechanistic studies, it has been shown that indolicidin failed to cause bacterial lysis even at concentrations up to 4 times the MIC concentration, thus excluding traditional membrane depolarization and pore models. In addition, indolicidin is capable of binding to DNA and inhibiting DNA synthesis which results in cell filamentation [36,37,38]. Recent studies have also demonstrated that the inhibition of DNA replication and transcription is due to the peptides central PWWP motif, which is able to wrap around and stabilize DNA structures [39]. In order to improve the potency of indolicidin, the effect of various structural modifications has been tested. These include tryptophan replacement with non-natural amino acids [40], expression of mutant structures such as CP10A (ILAWKWAWWAWRR-NH2) and CP11 (ILKKWPWWPWRRK-NH2) with increased potency against Gram-positive bacteria [41] and enhanced cationicity which improved the hemolytic profile and the activity against Gram-negative bacteria [42]. Additionally, studies have shown that cyclization of indolicidin improved the peptide proteolytic stability without significantly affecting the peptide antibacterial mode of action and activity [43].
Table 1. Overview of some natural antimicrobial peptides.
Table 1. Overview of some natural antimicrobial peptides.
NameAmino Acid Sequence aOriginReference
α-defensin (HNP-2)C1YC2RIPAC3IAGERRYGTC2IYQGRLWAFC3C1Human[44]
β-defensin (BD2)GIGDPVTC1LKSGAIC2HPVFC3PRRYKQIGTC2GLPGTKC1C3KKPHuman[20]
LL-37LLGDFFRKSKEKIGKEFKIVQRIKDFLRNLVPRTESHuman[45]
ProtegrinRGGRLC1YC2RRRFC2VC1VGRPig[46,47]
IndolicidinILPWKWPWWPWRR-NH2Cattle[34,35]
Magainin 2GIGKFLHSAKKFGKAFVGEIMNSAfrican clawed frog[48]
Cecropine AKWKLFKKIEKVGQNIRDGIIKAGPAVAVVGQATQIAK-NH2Hyalophora cecropia[49]
MellitinGIGAVLKVLTTGLPALISWIKRKRQQHoney bee[50]
Magainin IIGIGKFLHSAKKFGKAFVGEIMNSAfrican clawed frog[10,11]
PolyphemusinRRWC1FRVC2YRGFC2YRKC1RHorseshoe crab[5,51]
Gramicidin Scyclo-(Val-Orn-Leu-D-Phe-Pro)2Bacillus brevis[52]
Nisin A bI-DHB-A1I-DHA-LA1-ABA2-PGA2K-ABA3-GALMGA3NMK-ABA4-A-ABA5-A4HA5SIHV-DHA-KLactococcus lactis[53]
a connected cysteines forming disulfide bridges are indicated with numbers in subscript; b nisin contains five lanthionine/β-methyllanthinonine rings between alanine and/or aminobutyric acid (ABA) residues, it also contains several dehydroalanine (DHA), dehyrdobutyrine (DHB) residues. The ring pattern is indicated with numbers in subscript.
Human cathelicidin LL-37 is an extensively studied member of the cathelicidin family of antimicrobial peptides, and is also the only cathelicidin found in humans. It is most abundant in the granules of neutrophils (cells of the immune system). Upon infection and inflammation, LL-37 is released in high concentrations at neutrophil accumulation sites. Human cathelicidin LL-37 is also produced by lymphocytes and macrophages as well as different epithelial cells, and hence detected in plasma, sweat and other body fluids [54]. The family of cathelicidins are classified based on their highly conserved cathelicidin domain, which is flanked by an N-terminal signal peptide and a C-terminal segment corresponding to the active antimicrobial peptide. Thus, LL-37 is the predominant fragment from proteolytic cleavage of the C-terminal of the human host defense precursor protein (hCAP-18) (Figure 1 and Table 1) [45]. The gene encoding hCAP-18 contains three vitamin D receptor elements in its promoter region and is under regulation by various signaling pathways where multiple receptors are involved [55,56]. Ligand binding to the vitamin D receptor, triggers complexation with vitamin D receptor elements in the promoter region, initiating transcription of mRNA that translates into hCAP-18 precursor protein [57,58]. LL-37 is up-regulated under inflammatory conditions and is also specifically up-regulated in response to compounds like butyrate and vitamin D3 [59]. Bacterial components are another type of LL-37/hCAP-18 gene expression inducers. For example, it has been shown that expression of LL-37/hCAP-18 significantly increased in Helicobacter pylori-infected patients in contrast to individuals with non-Helicobacter pylori induced inflammation [60]. Once expressed, LL-37 exerts various biological activities; i.e., direct bactericidal activity against Gram-positive and Gram-negative bacteria observed in vitro [61] and modulation of inflammation and immune response in host cells against various infections [62,63]. LL-37 promotes apoptosis of infected epithelial cells, promoting clearance of respiratory pathogens [64]. Production of cytokines like interleukin-6 and interleukin-10 is also stimulated by LL-37 in a synergistic manner, thus enhancing the immune response through a complex mechanism involving multiple pathways [65]. Other studies have also shown that LL-37 enhances production of interleukin-8 and induces the expression of α-defensins, another class of cationic host defense peptides [66]. Enhanced clearance of Pseudomonas aeruginosa, an opportunistic pathogen, of immuno-compromised individuals has been demonstrated by LL-37 expression in murine lungs [67]. Besides stimulating production of immune signalling molecules, LL-37 can also function as a chemo-attractant, recruiting mast cells involved in wound healing and defence against pathogens, thus promoting the release of pro-inflammatory mediators [68]. In parallel to up-regulating several cytokines LL-37 can also inhibit interferon-gamma expression, a crucial cytokine for the innate and adaptive immunity against viral and bacterial infections [69]. Even though the fundamental mechanisms of LL-37 contribution to pathogen elimination have been extensively reported, the specific mechanisms by which LL-37 modulates the innate immune response remain poorly characterized.

3. Insect Antimicrobial Peptides

Insects represent by far the most diverse group of animals on Earth. Lacking an adaptive immune system, they fight against various pathogens by quick cellular and humoral responses. Thus, antimicrobial peptides play an important role in the insect’s defense system as they are secreted in the hemolymph as a result of the humoral defense mechanism upon microbial infections. Antimicrobial peptides isolated from different insect species are categorized as i.e., cecropins, attacins, lysozymes, defensins and dipteracins. Furthermore, in Drosophila two new groups, drosomycin and the metchikowins, have been identified and characterized [70].

3.1. Cecropins

Cecropins are the most abundant class of linear antimicrobial peptides in insects. They are devoid of cysteine residues and adopt most often an α-helical conformation in membrane mimetic environments. To exemplify, cecropin A (KWKLFKKIEKVGQNIRDGIIKAGPAVAVVGQATQIAK-NH2) is one of the first cecropins for which a solution confirmation for adaptation of a stable α-helix in hydrophobic environments has been demonstrated using circular dichroism spectroscopy (Table 1) [49]. In general, the structural fingerprint of cecropins is the presence of tryptophan residues at position 1 and 2 in addition to the amidated C-terminus [71]. However, a few exceptional structures including non-amidated C-terminus and peptides without N-terminal tryptophans have also been reported, with broad spectrum antibacterial activity comparable to that of cecropin A [72].

3.2. Melittin

Melittin is the most characterized α-helical, cationic and hemolytic antimicrobial peptide in the literature, which constitutes 50% of the dry weight of bee venom [73]. It is a 26-amino acids amphiphilic peptide (GIGAVLKVLTTGLPALISWIKRKRQQ), characterized by a hydrophilic C-terminal and predominantly hydrophobic N-terminal region, and serves as a model peptide in many membrane-peptide interaction studies (Table 1) [50]. Moreover, structure-activity relationship studies where different segments of melittin have been exposed to substitution or deletion, have demonstrated the key elements required for the high hemolytic and antimicrobial activity of this peptide. For example, replacement of Pro14 with Ala yields a 2-fold higher hemolytic peptide [50], deletion of the 6 residues from the C-terminal end results in a non-hemolytic peptide fragment [74]. Also, deletion of specific residues such as Leu6, Leu9, Leu16, Iso17 and Trp19 causes decrease in the hemolytic activity and deletion of Ala4 and Lys7 reduces the antibacterial activity relative to the native melittin [75,76]. In regards to the mechanism of action, melittin is established as a pore forming antimicrobial peptide, though there is no universal agreement of the type of pore it forms (for a review see [73]). In an attempt to design small analogs of naturally occurring antimicrobial peptides, researchers have designed hybrid peptides consisting of elements from cecropin A and melittin, which have demonstrated very potent antibacterial activity and no hemolytic activities [77]. One of the hybrid peptides, cecropin A-melittin (CAM), has been further modified by replacement of four specific residues with Trp5,11,21,25 in order to improve the antibacterial activity and improve the proteolytic stability. Thus, the new hybrid, CAM-W exhibits 3–12 times higher antibacterial activity, and strong antifungal activity while retaining a moderate cytotoxicity (IC50 > 300 mg/L) [78]. In summary, native or modified insect antimicrobial peptides hold potential as alternatives to the conventional antibiotics against various bacterial and fungal pathogens.

4. Plant Antimicrobial Peptides

Plants have also evolved a plethora of defense molecules to fight pathogenic infections. Antimicrobial peptides isolated from plants are positively charged and display the typical structural motifs such as rtα-helices, β-sheets and loops. They are divided into five classes namely; lipid transfer proteins [79], thionins, defensins [80], chitin-binding proteins [81] and cyclotides [82]. Thionins and defensins represent the first antimicrobial peptides isolated from plants, thus their activities will be discussed in the following section.

Thionins and Defensins

Thionins consist of up to 48 amino acids, predominantly arginine, lysine and cystein residues and contain a few conserved disulfide linkages (Figure 3). As a part of the plant defense system, thionins act on various pathogens and exert toxic effects on bacteria, fungi, yeast and mammalian cells [83]. A subgroup of the thionins family is the β-purothionins, which interact strongly with lipid bilayer, then insert into the hydrophobic core, resulting in cell lysis [84]. Viscotoxins, another class of plant thionins, were isolated from the leaves and stems of the European mistletoe (Viscum album) in 1973 [85]. Studies have demonstrated their high conformational stability and their ability to disrupt bacterial membranes [86]. γ-Thionins, now known as plant defensins, are another superfamily of plant antimicrobial peptides.
Figure 3. Three structural classes of thionins. (A) Members from all three classes have been superimposed to show their similarities. The same structures are illustrated separately in panels B-D. The PDB codes are all indicated, in addition to the primary sequence with the disulfide bridges indicated with subscript numbers; (B) β-purothionin (PDB code 1BHP) KSC1C2KSTLGRNC3YNLC4RARGAQKLC4ANVC3RC2KLTSGLSC1PKDFPK [87]; (C) γ 1-H thinonin (PDB code 1GPT) RIC1RRRSAGFKGPC2VSNKNC3AQVC4MQEGWGGG NC2DGPLRRC3KC4MRRC1 [88]; and (D) hellethionins (PDB code 1NBL) KSC1C2RNTLARNC3YNAC4RFTGGSQPTC4GILC3DC2IHVTTTTC1PSSHPS [89].
Figure 3. Three structural classes of thionins. (A) Members from all three classes have been superimposed to show their similarities. The same structures are illustrated separately in panels B-D. The PDB codes are all indicated, in addition to the primary sequence with the disulfide bridges indicated with subscript numbers; (B) β-purothionin (PDB code 1BHP) KSC1C2KSTLGRNC3YNLC4RARGAQKLC4ANVC3RC2KLTSGLSC1PKDFPK [87]; (C) γ 1-H thinonin (PDB code 1GPT) RIC1RRRSAGFKGPC2VSNKNC3AQVC4MQEGWGGG NC2DGPLRRC3KC4MRRC1 [88]; and (D) hellethionins (PDB code 1NBL) KSC1C2RNTLARNC3YNAC4RFTGGSQPTC4GILC3DC2IHVTTTTC1PSSHPS [89].
Pharmaceuticals 08 00366 g003

5. Antimicrobial Peptides Produced by Bacteria

Antimicrobial peptides produced by bacteria, originally named colicins, are referred to nowadays as bacteriocins. Bacteriocins produced by the host bacteria are able to selectively act against a broad spectrum of bacterial species without harming the producer. This family of antimicrobial peptides finds expansive application in the food industry to protect and prevent food contamination, particularly as many bacteriocins are produced by food-grade lactic acid bacteria. Additionally, bacteriocins attract special attention for their clinical implications for treatments of methicillin-resistant S. aureus, enterococci (including VRE), streptococci, Clostridium botulinum and Propionibacterium acnes. As a result of extensive work over the past decades there have recently been calls for the classification of bacteriocins to be revised and the current grouping suggested by Cotter et al. includes Class I (lanthionine-containing), Class II (non-lanthionine containing) and bacteriolysins (non-bacteriocin lytic proteins) [90].

5.1. Nisin

Nisin is a classical example of a Class I bacteriocin that has been successfully approved as an antimicrobial with commercial relevance in over 50 countries worldwide since its discovery back in 1927 [91]. Isolated from Lactoccocus lactis it was not until 1971 that its complex structure, consisting of 34 amino acids including some unusual lanthionine residues, four β-methyllanthionines, didehydroalanine and didehydroaminobutyric acid was fully described (Figure 4 and Table 1) [53]. Its activity against Gram-positive pathogenic bacteria is attributed to its interaction with the bacterial membrane resulting in disruption of the membrane integrity. This occurs via pore formation and so far two mechanisms are proposed. The first one explains the low-affinity pore formation of nisin alone, whereas the second one accounts of nisin and lipid II interaction thus ascribing to the pore formation by this complex (Lipid II-dependent pathway) [92,93,94,95]. Several structure-activity relationship (SAR) studies have been reported where the importance of different segments, especially those containing lanthionine rings, have been highlighted to be essential for the observed antimicrobial activity [96].
Figure 4. Covalent structures of two highly related bacteriosins. (A) Nisin A from Lactoccocus lactis and (B) Mutacin 1140 from Streptococcus mutans.
Figure 4. Covalent structures of two highly related bacteriosins. (A) Nisin A from Lactoccocus lactis and (B) Mutacin 1140 from Streptococcus mutans.
Pharmaceuticals 08 00366 g004

5.2. Mutacin 1140

Mutacin 1140 [97] is another member of the lanthionine-containing bacteriocins which has been extensively studied. It is a 22 amino acid long peptide produced by Streptococcus mutans [98] with a mode of action similar to that of nisin, which relates to the ability to inhibit peptidoglycan synthesis by binding to lipid II (Figure 4) [95]. Mutacin 1440 has a broad spectrum of activity against clinically important bacteria with time of kill profiles consistent with that of vancomycin, an antibiotic approved for treatment of severe infections caused by Gram-positive bacteria. This peptide is currently under investigation for its pharmacological relevance and has entered preclinical studies [99].
Lysostaphin is a bacteriolysin that acts by hydrolyzing the cell wall of the pathogenic bacteria, thus it is currently being investigated for its potential in food and nutritional microbiology. Even though bacteriolysins are structurally much larger than the traditional bacteriocins and antimicrobial peptides, their application as a peptide antibiotics hold vast potential. Lysostaphin (27 kDa) for example, is able to cleave the pentaglycine cross-links in the cell wall of Staphylococci and thus lyse the pathogenic bacteria in all metabolic states. The therapeutic efficacy of this bacteriolysin against clinical methicillin-resistant S. aureus MRSA isolate has been reported in animal models and compared to that of vancomycin. The recombinant lysostaphin, displayed better in vitro and in vivo antibacterial activity against methicillin-resistant S. aureus compared to vancomycin, in addition to having a low toxicity profile, therefore it is a potential lead candidate for further structural optimization studies [100].

6. Structural Properties of Antimicrobial Peptides

The structural properties of any given biomolecule play a major role in the interpretation of its biological activity. Many studies have been conducted to dissect the important elements that define peptides as antimicrobial “weapons”. Broad diversity of antimicrobial peptide sequences exists in nature, which contributes to the overall structural diversity, yet there are evolutionarily traits that have been conserved to ensure their activity on various types of bacteria with different membrane composition and different targets. The secondary structure, cationicity, hydrophobicity and amphipathicity are the most important key elements that allow characterization of antimicrobial peptides and therefore are briefly discussed in the following section.

6.1. Secondary Structure

Based on their secondary structures, antimicrobial peptides are classified in four groups; α-helices β-sheets, mixed structures and non- α- or β- structures (extended) (Figure 1). Circular dichroism spectroscopy, X-ray crystallography and nuclear magnetic resonance spectroscopy have been used intensively for the structure determination of antimicrobial peptides. For example, the first X-ray crystallographic structure of native human α-defensin, the human neutrophil peptide 3, appeared in 1991, followed by a nuclear magnetic resonance structure of human neutrophil peptide 1 [101,102,103]. Such structural information is found to be useful when investigating the importance of the secondary structure in deciphering the antimicrobial activity of antimicrobial peptides. To exemplify this, studies have demonstrated that bacterial killing by human neutrophil peptide 1 is structure-independent [104]. The most common motif found in many proteins and peptides with biological activities is the amphipathic α-helix, however, many antimicrobial peptides exist as extended or unstructured conformers and only adapt α-helical conformations upon interaction with phospholipid membranes [105,106]. Circular dichroism analysis has shown that in the presence of unilamellar phospholipid vesicles with varied content of zwitterionic and negatively charged phospholipids, several antimicrobial peptides adopt well defined α-helical and/or β-sheet like structures in contrast to the buffer environment [107].
The presence of α-helical structures in antimicrobial peptides is generally believed to promote interaction with membranes and assist membrane lysis [108] and therefore specific amino acids such as alanine, leucine, arginine, lysine, etc., that have high helical propensity (they occur more frequently in α-helices) have been actively included in the design of potential novel antimicrobial peptides [109,110]. In this context, Deslouches et al. showed that in general, the peptides from their de novo library, with more than 80% helical content, exhibited maximal antimicrobial potency against the tested bacterial strains [111]. On the other hand, Javadpour et al. also designed a library of highly α-helical peptides containing lysine and/or leucine/alanine/glycine residues in sequences that extended up to 21 residues. In this library the propensities of α-helical conformation was found to be proportional to the toxicity of the peptides against mammalian model cells. However, no definite conclusion could be obtained for a correlation of these peptides’ conformations and their observed antibacterial activity [112].

6.2. Conserved Salt Bridges

Salt bridges can be found in both α-helical and β-sheet peptides, and contribute greatly to the overall stability of the secondary structure of antimicrobial peptides. However, studies on α-defensin has demonstrated that salt bridges in human defensin 5 do not contribute to the antimicrobial activity, but rather affect the correct disulfide pairing of the pro-α-defensins, which are inactive defensins subjected to further processing. Furthermore, the salt bridges can also indirectly increase proteolytic stability, as seen for human defensin 5 (Figure 5), where modifications that disrupted the salt bridges also accelerated degradation by trypsin [113]. Similar salt bridges has also been observed in other antimicrobial peptide structures, e.g., in human lactoferricin 1-49 (Figure 5) [114].
Figure 5. Salt bridges in antimicrobial peptides. The overall secondary structure is depicted, in addition to side-chains and back-bone elements that are involved in stabilizing the structure; hydroben bonding (green dotted lines) or salt bridges (grey dotted lines). The PDB codes are all indicated, in addition to the primary sequence with the disulfide bridges indicated with numbers in subscript. (A) On human defensin 5 (PDB code 2LXZ) ATC1YC2RTGRC3ATRESLSGVC2EISGRLYRLC3C1R [115], it is apparent how Arg6 and Arg28 from β-strand 1 and 3, clam around and stabilize the random coil backbone through formation of one hydrogen bond and one salt bridge; (B) Human lactoferricin 1-49 (PDB code 1Z6V) GRRRRSVQWC1AVSQPEATKC2FQWQRNMRKVRGPPVSC2IKRDSPIQC1IQA [114], contains a very loose structure with a disrupted or segmented helical elements. The composition of the helical element is pending on how the structure is captured, and as depicted in this structure, it is apparent how a salt bridge between Lys19 and Gln24 are involved in stabilizing the random coil bridging the two largest helical segments.
Figure 5. Salt bridges in antimicrobial peptides. The overall secondary structure is depicted, in addition to side-chains and back-bone elements that are involved in stabilizing the structure; hydroben bonding (green dotted lines) or salt bridges (grey dotted lines). The PDB codes are all indicated, in addition to the primary sequence with the disulfide bridges indicated with numbers in subscript. (A) On human defensin 5 (PDB code 2LXZ) ATC1YC2RTGRC3ATRESLSGVC2EISGRLYRLC3C1R [115], it is apparent how Arg6 and Arg28 from β-strand 1 and 3, clam around and stabilize the random coil backbone through formation of one hydrogen bond and one salt bridge; (B) Human lactoferricin 1-49 (PDB code 1Z6V) GRRRRSVQWC1AVSQPEATKC2FQWQRNMRKVRGPPVSC2IKRDSPIQC1IQA [114], contains a very loose structure with a disrupted or segmented helical elements. The composition of the helical element is pending on how the structure is captured, and as depicted in this structure, it is apparent how a salt bridge between Lys19 and Gln24 are involved in stabilizing the random coil bridging the two largest helical segments.
Pharmaceuticals 08 00366 g005

6.3. Cationicity

One common property of the majority of the antimicrobial peptides is the cationic nature represented by the number of positively charged residues (lysine, arginine and histidine) within their structure, which ranges between +1 and +7 charges [3]. A number of studies have correlated this property with the antimicrobial activity observed in antimicrobial peptides. The attributed importance lies mainly in the interaction between the positive charge in the peptides and the negatively charged bacterial membrane surfaces via electrostatic interactions. For instance, the antimicrobial activity of most of the members of the defensin family appears to be related to their cationicity. To exemplify, human defensin 5 interacts with the bacterial surface via its arginine residues and thus exerts its antimicrobial activity. Replacement of arginine residues at position 9 and 28 with alanine or lysine residues reduces the antibacterial killing as well as the host cell interaction, the latter which is found to be receptor mediated [116]. The overall positive charge may not be the main determinant in the observed antibacterial activity as in case of human α-defensins that generally show to be more effective against Gram-positive bacteria than human β-defensins, even though human β-defensins are more cationic [117,118]. Herein, human neutrophil peptide-1 (ACYCRIPACIAGERRYGTCIYQGRL WAFCC, net charge +3) appeared to be more effective against S. aureus than human β-defensin-3 (GIINTLQKYYCRVRGGRCAVLSCLPKEEQIGKCSTRGRKCCRRKK net charge +11) [23,119,120]. In addition, cationicity meets a limit, beyond which increasing the charge no longer results in increased antibacterial activity. That being said, Dathe et al. has demonstrated that increasing the charge from +3 to +5 in magainin-2 analogues, also increased the peptides antibacterial activity against both Gram-positive and Gram-negative bacteria, but further increase to +6 or +7 led to a loss of the antibacterial activity and increased hemolytic propensity [121]. In addition to the net charge of antimicrobial peptides, it has further been illustrated that the position of the charged residues is also an important factor that determines the peptides overall antibacterial activity, e.g., changing the position of a few amino acid residues within the native structure of the linear bactenecin, Bac2A (RLARIVVIRVAR-NH2), resulted in a scrambled sequence that showed increased antibacterial activity [122]. Besides the importance of cationicity in mediating initial interaction with target membranes, in Gram-negative bacteria, the net positive charge is important for the so-called self-promoted uptake of antimicrobial peptides. Herein cationic antimicrobial peptides interact with the outer membrane surface where there are divalent cations such as Mg2+ or Ca2+ cross bridging LPS molecules. Displacement of these cations causes destabilization of the outer membrane and therefore allowing uptake of molecules [123,124]. Moreover, bacteria carry more negative transmembrane potential when compared with that of a normal mammalian cell membrane and this will facilitate insertion of charged antimicrobial peptides into the membranes [125].

6.4. Hydrophobicity

Hydrophobicity is certainly an inevitable structural feature that decides the overall activity of a given antimicrobial peptide and therefore it is continuously characterized in the literature as a key functional property. It affects the potential of interaction between antimicrobial peptides and different membrane compositions and furthermore directs the degree of peptide partitioning into the lipid bilayer. Increased hydrophobicity is well correlated with loss of antibacterial specificity, resulting in high toxicity towards mammalian cells. To elaborate, Yin et al. showed that substituting four alanine residues with four hydrophobic leucine residues induced higher hemolytic activity of their membrane active model peptide [126]. Magainin-2 analogues with varying hydrophobicity have also been used to demonstrate that minor changes in the hydrophobicity may drastically influence and increase its membrane binding and permeabilization activity [127]. The effect of hydrophobicity on the antibacterial and hemolytic activities has been further demonstrated via peptide series constructed of repeats of lysine and tryptophan (LysTrp)n residues. Addition of a number of repeats of these two residues resulted in parallel increase of both hydrophobicity, as estimated by the retention times on HPLC, and antimicrobial activity. However, when five repeated units were present in the structure (LysTrp)5, the increase caused unfavorable change in the hemolytic profile (increased toxicity) and decrease in the antimicrobial activity due to self-aggregation [128]. In conclusion, the hydrophobicity of a specific sequence in the development of a novel antimicrobial peptide should be challenged but not exaggerated.

6.5. Amphipathicity

Amphipathicity in antimicrobial peptide structures reflect the abundance and polarization of the hydrophobic and hydrophilic domains. Most of the cationic antimicrobial peptides display a net positive charge (+1 to +7) [3] and consist of about 50% hydrophobic residues which contribute to the recognition and interference with the cytoplasmic membrane barrier or self-promoting uptake across the cellular membranes. The cationic charges and the hydrophobic groups segregate into amphiphilic structures. One quantitative measure of amphipathicity is the hydrophobic moment, that applies for peptides in α-helical confirmation and is used as a descriptor in dissecting the role of amphipathicity for peptide antimicrobial activity [129]. The importance of the amphiphilicity in determination of the antimicrobial activities of these peptides is controversial because different research groups report on favorable and unfavorable contributions such as increase in antimicrobial and increase in hemolytic activities, respectively [127,130,131,132,133,134,135]. For example, amphipathicity has been reported as a major structural determinant for the biological activity of a small library of arginine and tryptophan rich linear and cyclic hexapeptides [133]. Here, the structural and conformational constrains of the peptides together with the ideal positioning of the hydrophobic clusters appeared to determine the antimicrobial activity and selectivity of the peptides [133]. In another study of magainin-2 and its analogues it was demonstrated that the antimicrobial activity was governed explicitly by the peptide amphipathicity and not by hydrophobicity or α-helical content [134]. Similarly, a design of cecropin A and melittin hybrid structures has demonstrated increased amphiphilicity and helicity correlating with high antibacterial activity and low toxicity against mammalian cells [135]. Contradicting these results are numerous other peptide studies, demonstrating that high amphipathicity if measured as hydrophobic moment, increases membrane disruption resulting in increase in both the antibacterial and hemolytic activity [127,130,131]. Amphipathicity in antimicrobial peptides that exist in β-sheet conformations is characterized by the number of β-strands organized by two distinct polar and non-polar domains. β-strands in antimicrobial peptides are usually stabilized via disulfide bridges or head-to-tail cyclization which provides high conformational rigidity in aqueous solution. The polar and non-polar domains in β-strands allow antimicrobial peptides to successfully interact with target membranes and once associated with the membrane, the amphipathic nature enables membrane disruption via formation of transmembrane channels [105]. Alteration in the non-polar domain of gramicidin S derivatives, decreased hydrophobicity and the overall amphipathicity, thus not surprisingly decreased the peptides hemolytic activity [136].

6.6. Cyclic Antimicrobial Peptides

Besides adopting linear structures, antimicrobial peptides can be naturally found in cyclic conformations. Antimicrobial peptides are constrained in this conformation either by disulfide cross linkages (tachyplesins (KWC1FRVC2YRGIC2YRRC1R-NH2) [137], protegrins (RGGRLC1YC2RRRFC2VC1VGR-NH2) [138], and polyphemeusins (Figure 1)) or backbone cyclization (gramicidin S, tyrocidines (Figure 6) and θ-defensins (Figure 2)). Cyclic antimicrobial peptides have demonstrated strong antimicrobial activities against different pathogenic bacteria, however with poor selectivity. Therefore, numerous structure-activity relationship studies have been done for dissecting the important elements that contribute to the observed activities with the intention to improve their therapeutic profiles [136]. Scheinpflug et al. demonstrated improved antimicrobial activity against Gram-negative and Gram-positive bacteria and low hemolytic activity of a small cyclic hexapeptides (c-RRRWFW), when compared to that of the linear analog sequence. Regarding the mechanism of action, this cyclic peptide did not act on the bacterial membrane, rather it has been suggested that it travers the cell wall with the help of direct interaction to lipopolysaccharide [139] and then further translocates to the cytoplasm [140]. Other studies have used the native structure of the gramicidin S, as a model peptide for synthesis of novel cyclic antimicrobial peptides [141]. In this context, the effect of the ring size (4-14 residues) in gramicidin S on the antimicrobial and hemolytic activity had been analyzed, in addition to the size effect on the secondary structure and the peptides lipid binding potential. The ring structures containing 6, 10 and 14 residues folded in β-sheet structures, whereas disordered structures had been observed for rings formed by 8 or 12 residues. The membrane disruption had been observed only in the peptide analogues with 10 or more residues in the ring structure, whereas those with less appeared to be completely inactive without any hemolytic activities. Out of this study only one cyclic analog showed improved antibacterial specificity, indicating that varying native cyclic peptide structures can contribute to the development of clinically useful cyclic peptide antibiotics [142]. The 14-mer analogue, GS14 forms a highly amphipathic β-sheet structure and has served as a model peptide for further modifications and structure activity studies. GS14 exhibits very high hemolytic and limited antimicrobial activity so any change in the structure that will disturb the amphiphilic β-sheet structure gives a direct relation of the amphipathicity and the biological properties of the cyclic peptide. By systematic synthesis of GS14 variants, Kondejewski et al. demonstrated that the hemolytic activity exhibited by GS14 could be easily reduced by decreasing the amphipathicity (reduction of directed hydrophobicity) or by reduction of the overall peptide hydrophobicity [136]. Other attempts to identify structural features and improved antimicrobial activity of cyclic peptides have been communicated in two consecutive studies [133,143]. Here, the cyclic peptides with the highest antimicrobial activity were those with three aromatic residues positioned adjacent to each other. In summary, since amphipathicity is of a great importance when tuning the antimicrobial activities of peptides, cyclic peptides permit induced amphipathicity and greater enzymatic stability [43,144] and should therefore be considered a valuable path forward.
Figure 6. Cyclic peptides and lipopeptides. Covalent structures of some conventional and lead drug molecules. (A) Gramicidin S; (B) Tyrocidines; (C) Polymyxin B; (D) Daptomycin and (E) cyclo-[D-Ala-(12-guanidinododecanoyl)Thr-D-Val-Val-DaThr-D-Asn] [145].
Figure 6. Cyclic peptides and lipopeptides. Covalent structures of some conventional and lead drug molecules. (A) Gramicidin S; (B) Tyrocidines; (C) Polymyxin B; (D) Daptomycin and (E) cyclo-[D-Ala-(12-guanidinododecanoyl)Thr-D-Val-Val-DaThr-D-Asn] [145].
Pharmaceuticals 08 00366 g006

6.7. Length

The length of the polypeptide chain in antimicrobial peptides varies greatly and there are continuous reports on many active sequences that are not necessarily similar in size. This raises the question whether or not there is an optimal chain length that is beneficial for the observed activity of antimicrobial peptides. To answer this question, a few research studies are presented. For example, in design of a de novo peptide library it was demonstrated that by increasing the chain length of a 12-mer peptide composed of only arginine and valine, one could achieve an increase in antimicrobial activity against P. aeruginosa and S. aureus [111]. Notably, others have reported that an increase in length could give a proportional increase in the toxicity towards mammalian cells [112]. Similarly, Lui et al. showed that by increasing the chain length of a peptide from a 2-mer (RW) to a 10-mer (RW)5 both the antibacterial activity and host cell toxicity increased, however the latter being less affected [146]. Correspondingly, in a study with α-helical model peptides (lysine, leucine, alanine), it appears that 21-mers in general are about two-fold more potent than their 14-mer analogues. Contradictory observations have been made in a different α-helical model peptide library (LARL)3-(LRAL)n (n = 0–3). In this study the peptides exhibited decreased antimicrobial activity and yet increased hemolytic activity with increasing chain length [147]. The above observations are further supported by a study by Gopal et al. where an increase in chain length up to eight residues enhanced antimicrobial activity of peptides composed of repeated lysine and tryptophan (KW)n and the effect on hemolysis increased gradually with the increase of the chain length. However, the increase of chain length from (KW)4 to (KW)5 did not further improve the antimicrobial activity [128]. A similar trend demonstrating increase in antibacterial activity up to a certain chain length before it decreases again, has also been reported for a set of arginine and valine rich β-hairpin like peptides (Ac-C(VR)nDPG(RV)nC-NH2 (n = 1–5) [148]. Additional insight into peptide length can also be learned from peptide sequence truncation. Some examples include truncation of peptide LL-37 [149]. Thus, in conclusion there is no definite length of a peptide sequence that is to be taken into account when one designs novel antimicrobial peptides. However, the studies conducted provide a better understanding of this issue and suggest optimal strand length to improve antibacterial peptide activity for the respective peptide species studied.

7. Structural Properties of Specific Amino Acid Residues in Antimicrobial Peptide Sequences

Identified antimicrobial peptides from plants, insects, fish, frogs and mammals vary greatly with respect to their amino acid composition. Each of the amino acids alone or in synergy with the neighboring amino acid residues contributes to the observed antimicrobial properties conserved in the peptides. The following section highlights the importance of few amino acids commonly found in the structure of many cationic antimicrobial peptides and discusses their contribution to the peptides therapeutic profiles.

7.1. Lysine (Lys, K) and Arginine (Arg, R)

The basic amino acids Lys and Arg are highly conserved residues in antimicrobial peptide structures as they enable electrostatic interactions between the peptide and the negatively charged bacterial membranes [150]. These two amino acids differ in their side chain chemistry. Arginine has a guanidinium group that allows more dispersed positive charge and offers greater directionality and possibility of hydrogen bonding with for example the surrounding water molecules. The unique characteristics of the arginine side chain allow formation of multiple interactions contrasting the mono charge present in lysine. In addition, arginine can also engage in cation–π interaction with tryptophan, where the negative charge clouds in the tryptophan aromatic systems interact with the positively charged side chain [151]. The roles of amino acid substitution and cationicity on antimicrobial and hemolytic activities have been thoroughly investigated over the past 20 years. For this reason, many peptide analogues have been synthesized where known sequences have been modified to include different cationicity using lysine and arginine. Many studies have demonstrated that substitution of arginine with lysine leads to reduced antimicrobial activity [152,153]. Gopal et al. [154] reported on an analogue sequence, (FKKLKKLFKKILKLK-NH2) of HPA3NT3 peptide, where they substituted Trp12 and Trp14 with leucine as well as Arg3 and Asn13 with lysine. They observed an increase in the cationicity due to the substitution of arginine and asparagine with lysine and C-terminal amidation, which resulted in small changes in the antibacterial activity but significant decrease in hemolytic activity [154]. Similarly, substitution of arginine with lysine in cyclic c-RRWWRY resulted in decreased minimum inhibitory concentration and erythrocyte lysis [143]. In addition to the contribution of arginine residues to the initial electrostatic recognition of the membrane surfaces, poly-arginine sequences are able to pass cell membranes more efficiently than poly-lysine or other poly-cationic homopolymers, suggesting that the guanidine group of the arginine side chain is a critical component for the observed biological activity. Furthermore, once the peptides have entered the cell, arginine containing peptide sequences have demonstrated higher affinity for DNA than poly-lysine peptides [155]. Similarly, the arginine-rich peptide (RRWWRRWRR) has been actively used in the intracellular delivery of peptide nucleic acids. Substituting arginine residues in this sequence with lysine decreased the cellular uptake of the conjugate by six-fold [156]. Arginine- and lysine-rich peptide sequences are well described in histones, proteins found in eukaryotic nuclei that contribute to package and organization of DNA into nucleosomes. Lysine and arginine rich histones and histone fragments exhibit broad-spectrum antimicrobial activity [157] as a result of intracellular inhibition of cell functions trough binding of these fragments to bacterial nucleic acids [158,159]. Consequently, Morita et al. demonstrate that the contribution from these fragments to the observed antibacterial activity relies on the arginine-rich histone ability to disrupt the morphology and lysine-rich histone to disrupt the integrity of S. aureus and E. coli [157,159]. One of the limitations of the extracellular applications of antimicrobial peptides as antimicrobial coating agents is the inhibition of the electrostatic interactions with increasing concentrations of sodium chloride. To counteract this limitation, incorporation of arginine residues at the C-terminal end of antimicrobial peptides allows salt-resistance due to the increased cationicity which overall will improve the electrostatic interaction between the peptide and the anionic bacterial membranes [160,161]. To exemplify, human β-defensin has been exposed to such structural changes that led to improved antibacterial potency in ionic environments and further boost the development of many antimicrobial peptides that deal with such obstacles [160]. While tuning the membrane selectivity of antimicrobial peptides, Liu et al. [162] reported on another arginine-rich peptide sequence, (RW)4D, with enhanced antibacterial activity against E. coli and S. aureus and lowered hemolytic profile when compared with the natural antimicrobial peptide indolicidin [162]. Another example of improved membrane permeability of arginine containing peptide sequences is the simple substitution of D-lysine with D-arginine in the short antimicrobial peptide RLA [163].

7.2. Tryptophan (Trp, W)

Another very important residue found in the structure of antimicrobial peptides is tryptophan. The unique side chain containing an indole ring holds hydrogen-bonding potential in addition to other physiochemical properties, e.g., dipole and quadrupole moments [164]. Tryptophan residues show strong membrane-disruptive activities by their ability to interact with the interface of a membrane anchoring the peptide to the surface of the bilayer [165]. Scanning electron microscopy analysis has suggested that two tryptophan-substituted antimicrobial peptides, I1WL5W and I4WL5W, work through disruption of the bacterial cell membranes. In addition, fluorescence and quenching data from liposome studies has indicated possible insertion of these peptides into the lipid bilayers and induction of blue shifts in the tryptophan emission spectra [166]. The hydrogen bonding of tryptophan with the surrounding water molecules diminish upon insertion of the tryptophan residues into the hydrocarbon core of the membrane [167]. However, indole hydrogen bonding and the dipole-dipole interactions are shown not to be primary determinants of the tryptophan interfacial localization [167]. Various experimental and molecular-dynamic simulation studies have demonstrated accommodation of tryptophan residues in the interface layer of membranes [167,168] which could associate with the positively charged choline head groups of the lipid bilayer [151]. Overall, it is suggested that aromaticity and the molecular shape play an important role in explaining the nature of membrane interaction of tryptophan-rich antimicrobial peptides. To exemplify, the flat and rigid structure of tryptophan may influence the positioning in the hydrocarbon core for entropic reasons and the electrical properties of the aromatic system may favor accommodation of tryptophan in the interface regions. In accordance with a plethora of literature studies pointing out the importance of the overall structure and position of certain amino acids in antimicrobial peptide design, the position of tryptophan residues in the tryptophan substituted peptide, L-K6, a peptide derived from temporin-1CEb from skin secretions of the Chinese brown frog, demonstrated to be an important factor for the observed antibacterial activities against Gram-negative and Gram-positive bacteria [169]. The influence of tryptophan residues on antimicrobial potency against P. aeruginosa and S. aureus has been further documented where single substitution with tryptophan significantly increased the anti-pseudomonal potency when compared to the parent peptide [111]. In summary, tryptophan residues have strong preference for the interfacial regions of the lipid bilayer and aid attachment and insertion of the tryptophan containing peptide across the membrane barriers.

7.3. Cysteine (Cys, C) and Disulfide Bonds

Cysteine belongs to the sulfur-containing amino acids which are strongly reactive. The thiol group can be easily oxidized to form a dimer, thus creating a disulfide bridge between two cysteines. Disulfide bridges formed by cysteine residues are strongly hydrophobic (nonpolar) and play an important role in the structures of many antimicrobial peptides. In addition to being important for the overall structural fold of the peptide, these bridges also increase the peptides stability towards proteolytic degradation, e.g., defensins [170]. The tertiary structure stabilization via disulfide bonding of human neutrophile peptide 1 (HNP1) contributes to effective binding to the cell wall precursor lipid II [104] and inhibits TNF-α secretion by human monocyte derived macrophages [171]. In studies where disulfide bonds in HNP1 or human defensin 5 (HD5) were reduced (e.g., by dithiothreitol) or substituted, it has been demonstrated reduced antibacterial activity [170,172]. In contrast, in the highly cationic (+11) human β-defensin 3 (HBD3), the presence, absence or altered pairing of the three disulfide bridges, did not appear to be relevant for the observed antibacterial activity [173]. Some of the most potent natural cysteine containing antimicrobial peptides, are a β-hairpin polyphemusin 1 (RRWC1FRVC2YRGFC2YRKC1R) from horseshoe crab (Figure 1) [5,51] and pig protegrin (RGGRLC1YC2RRRFC2VC1VGR) [46,47]. Polyphemusin I exhibit high antimicrobial activity with MIC ranging from 0.125 to 1 µg/mL against multiple clinical strains of both Gram-negative and Gram-positive bacteria [174]. Similarly, protegrins, especially protegrin-1 is also stabilized by two internal disulfide bridges (Cys6-Cys15 and Cys8-Cys13) and as such is able to permeabilize bacterial membranes including the outer membrane of Gram-negative bacteria [175,176,177]. Loss of bactericidal activity is observed upon removal of cystein residues, especially those in position Cys6 and Cys15 [178,179]. Structural studies have demonstrated that protegrin-1 has the ability to form oligomeric β-sheet like structures in model membranes, inducing pore formation in bacterial membranes [180,181,182]. Structure-activity relationship studies on protegrin-1 have been conducted to evaluate the importance of the cysteins, and the results demonstrated that two truncated linear versions of protegrin-1 retained a broad spectrum activity by disrupting LPS-outer membrane barrier. Furthermore it is assumed that the peptide adopts an amphipathic β-hairpin-like conformation, even in absence of the disulfide bonds in complex with the LPS micelles. Currently, protegrin-1 analogues are the main cysteine containing peptides that hold promise for development of a non-toxic antimicrobial [183]. One example is the protegrin-1 derivative IB-367 (iseganan) that has been the most studied for an effective treatment of oral mucositis due to its broad spectrum antibacterial activity, rapid killing and relative lack of resistance development. However this analogue failed in Phase II clinical trials of oral mucositis [174,184,185].

7.4. Proline (Pro, P)

Proline-rich antimicrobial peptides are a distinctive class of cationic peptides isolated from both insects and mammals, with confirmed antimicrobial activities specifically against Gram-negative bacteria [186,187,188]. Structurally, proline is an unusual amino acid that forms a ring structure with rigid confirmation and a secondary amine compared to the other twenty natural amino acids. This significantly reduces the structural flexibility of the polypeptide chain, and the nitrogen in the idole ring cannot participate in hydrogen-bonding with other residues. Prolines are often considered helix breakers, however, proline-rich sequences tend to adopt distinct PP-II helix (e.g., PR-39 (RRRPRPPYLPRPRPPPFFPPRLPPRIPPGFPPRFPPRFP), apidaecins, drosocin), helix structure with three residues per turn [189]. Retaining highly potent antimicrobial activities, proline-rich antimicrobial peptides subsequently act in a divergent way including stereospecific interaction with membrane translocation system followed by intracellular targeting, compared with the more general membrane disruption mode of action of traditional antimicrobial peptides. Most of the knowledge about the structure-activity relationship for these peptides comes from peptides isolated from insects; e.g., apidaecin, a small 18 amino acid residues peptide (GNNRPVYIPQPRPPHPRL) [190] isolated from honeybees, with proline content of 33% [191], drosocin, a 19-mer glycopeptide (GKPRPYSPRPTSHPRPIRV) [190] isolated from Drosophila [192] and pyrrhocorycin, a glycopeptide (VDKGSYLPRPTPPRPIYNRN) [193] isolated from the firebug Pyrrhocoris apterus [194,195]. In addition, Bac7, Bac5 and PR-39 are representatives of well-studied mammalian proline-rich antimicrobial peptides. Regarding the mechanism of action of these distinct peptides, studies have demonstrated that the members of the proline-rich peptide group and their derivatives act in a completely divergent mechanism than the lytic amphiphilic antimicrobial peptides [192,196,197,198,199]. To exemplify, one of the early observations of the non-lytic mechanism were found in E. coli challenged with apidaecin or bovine PR-39, where the bacteria membranes remained intact during the entire incubation time [188,191]. Similarly, arasin-1, a 37 amino acid long proline-rich peptide (SRWPSPGRPRPFPGRPKPIFRPRPC1NC2YAPPC2PC1DRW) [200] isolated from the spider crab, Hyas araneus, exhibits bactericidal antimicrobial activity not related to membrane disruption, with the proline-rich region (1–23) as the confirmed region responsible for the observed activity [201]. It has been further suggested that proline-rich antimicrobial peptides stereo specifically bind to intracellular targets such is the bacterial heat shock DnaK protein and this binding can be correlated with the observed antimicrobial activity [193,202]. Such observations were made for pyrrhocoricin, and the N-terminal half (Asp2-Pro10) were demonstrated to be responsible for the antimicrobial activity, while the C-terminal part aids internalization into bacterial or mammalian cells [193,195]. This proline-rich antimicrobial peptide protects mice from bacterial infections and is nontoxic to the host. Further studies on artificial proline-rich peptides has verified that a number of proline sequences can travers bacterial membranes, e.g., poly-proline (P14) peptide [203] as well as a polyprolines family of peptides (VXLPPP)n (n = 1, 2, and 3; X correspond to histidine, arginine or lysine) [198]. Contrary to these findings there are also reports of proline-rich peptides and their ability to damage cell membranes, however these effects are highly concentration dependent [201,204,205].
Despite the well-established knowledge about the DnaK being an intracellular target for proline-rich peptides, very few studies suggest other vital intracellular targets in Gram-negative bacteria (for a review see [186,187]). Regardless, in the past five years, researchers have conducted new experiments to introduce proline-rich peptides to the market as promising antibiotics. Analogues of the native peptides, novel structures and synergistic studies between peptides and classical antimicrobials have been conducted [196,199]. In the efforts to design novel pyrrhocoricin derivatives peptides, large libraries have been synthesized and tested. One of the candidate peptides from this study, onconin (VDKPPYLPRPRPPRRIYNR-NH2), shares 70% structural similarity with pyrrhocoricin. It exhibited low toxicity towards mammalian cells and could penetrate lipid membranes without disrupting them [206]. Similarly Fritsche et al. [207] demonstrated that the antibacterial effect exerted by novel onconin and apidaecin derivatives, were not a result of immunomodulatory properties but rather a direct antimicrobial effect. An observation which potentially simplifies the development of novel proline-rich antimicrobial peptides into new pharmacological molecules [207]. Consequently and most recently, Krizsan et al. [208] speculate that onconins and apidaecins act on other targets than the chaperone DnaK alone [208]. In summary, proline-rich peptides are characterized by good water solubility, high potency against bacteria killing and low cytotoxic effects at high concentrations, making them attractive lead candidates for development of novel antimicrobial therapeutic agents.

8. Mechanism of Action of Antimicrobial Peptides

In drug development, a good antimicrobial candidate should exhibit highly specific biological activity followed by good pharmacokinetic profile and low immunogenicity. In order for peptides to be considered as antimicrobial agents of therapeutic relevance, it is essential to dissect their biological activities, specifically their mode of action. This task is not easy as there are over 2000 natural peptides with broad spectrum antimicrobial activities have been isolated. Despite this fact, a surplus of scientific study groups have provided us with structure-activity relationship information that helps the overall understanding of how antimicrobial peptide combat infectious agents. In order to do so, researchers have developed various model membranes that antimicrobial peptides act on. These include, micelles, artificial liposomes with varying lipid compositions (POPE, POPC, POPG, cholesterol etc.) and natural E. coli polar lipid extracts [209,210,211,212].

8.1. Direct Killing

It is generally assumed that cationic antimicrobial peptides interact with membranes where they disturb the amphipathic lipid bilayer, which further leads to disruption of vital bacterial physiological processes and ultimately bacterial death. Bacterial membranes possess large fraction of negatively charged lipids and maintain high electrical potential gradients (transmembrane potential), thus attracting positively charged compounds such as cationic antimicrobial peptides. A result of the electrostatic interaction of positively charged peptides on the negatively charged bacterial surfaces, is destabilization and release of native divalent cations from the membrane. This displacement leads to disruption of the outer membrane barrier in Gram-negative bacteria (Figure 7).
Figure 7. Direct killing mechanisms of antimicrobial peptides. The peptides are typically classified as membrane disrupting (A) or as molecules that act through more specific extra- or intracellular targets (B).
Figure 7. Direct killing mechanisms of antimicrobial peptides. The peptides are typically classified as membrane disrupting (A) or as molecules that act through more specific extra- or intracellular targets (B).
Pharmaceuticals 08 00366 g007
In comparison to bacterial membranes, plant and animal cell membranes are enriched in cholesterol and lipids, have no net charge, and maintain weak transmembrane potential [213]. Studies that refer to the electrostatic interactions between the positively charged amino acids and the negatively charged phospholipid head groups in the membranes have shown that by increasing the buffer salt concentrations the peptide activity on the negatively charged membranes is diminished [214].

8.2. Membrane Interaction

Antimicrobial peptides share common physiochemical features such as cationicity and amphipathicity that allows them to interact with membranes. When peptides come in a close proximity to a model membrane (bacterial or mammalian), the very first interaction is the binding. Information about the strength of this interaction helps in understanding the consecutive processes that eventually lead to cell death.
Isothermal titration calorimetry is one of the techniques applied to extract information about the antimicrobial peptide binding affinity to model membranes. Andrushchenko et al. [212] investigated the thermodynamics of the interaction of tryptophan-rich antimicrobial peptides with large unilamellar vesicles (LUVs). They obtained binding isotherms that could characterize the strength of binding of the peptides to the model membranes. Specifically, all peptides demonstrated to selectively bind stronger to anionic (PE/PG) and E. coli membranes than to zwitterionic (POPC) membranes, due to strong electrostatic interactions. An increase in charge did not improve on binding and replacement of arginine with lysine residues did not change the binding to the anionic membranes but decreased the binding to the zwitterionic LUVs. Other substitutions like proline by alanine in one of the tryptophan-rich peptides allowed α-helix formation of the peptide that resulted in reduced binding selectivity. Furthermore, tryptophan substitution with other aromatic residues significantly reduced the peptides’ ability to interact with both the anionic and the zwitterionic LUVs [212].
Fluorescence spectroscopy_is another technique implemented for studies of peptide interaction with model membranes, which provides useful but limited information about the affinity of the peptide for lipid bilayers. It may also shed light on peptide aggregation, location in the membrane and the effects on membrane integrity. When using fluorescence spectroscopy on tryptophan-rich peptides, variations of blue shifts in tryptophan fluorescence is measured, correspond to binding to model membranes. Series of tryptophan-containing β-hairpin peptides have been assessed for their binding to PE/PG and PC/cholesterol-containing vesicles and their binding properties particularly to PE/PG correlates well with their antimicrobial activity [148]. Furthermore, peptides like VRW3 (Ac-C(VR)3DPG(RV)3CW-NH2) peptide exhibited stronger binding to PE/PG compared to PC/cholesterol-containing vesicles, while in parallel demonstrating high antibacterial properties and low hemolytic activity [148].
Lipopolysaccharide (LPS) is the endotoxin found on the outer face of the outer membrane of Gram-negative bacteria that helps cell wall stabilization and increases the overall negative charge on the bacterial surface. It is responsible for systemic inflammatory response in mammalians, sometimes leading to septic shock. As a result of simple charge distribution, many cationic antimicrobial peptides are found to exhibit high affinity to LPS. Peptide interaction with LPS is facilitated by displacement of Mg2+ which naturally stabilize and cross-bridge adjacent LPS molecules in the outer membrane [123,124]. This interaction further aids the peptides self-promoted uptake granting them access to the inner membrane where they can exert bacterial killing. A contradictory theory suggests that when antimicrobial peptides bind to LPS they tend to form aggregates that are unable to cross the outer membrane and therefore are incapable of killing bacteria [215]. Regardless, when LPS are released as endotoxins, the binding of many cationic antimicrobial peptides will have the potential of neutralizing the toxin, thus they hold promise as new leads for development of improved strategies for clinical treatment of sepsis [216].

8.3. Membrane Disruption by Antimicrobial Peptides

Upon membrane binding, antimicrobial peptides contribute to possible alternations of the membrane structure such as thinning, pore formation, altered curvatures, etc. This results in overall membrane disruption by lowering the proton gradient (loss of membrane potential) that ultimately stops ATP production and cellular metabolism, leading to cell death [217,218,219,220]. As is the case with many other antimicrobial peptides, membrane permeabilization is a crucial step in the microbicidal activity observed for defensins [18]. The permeability effect has been demonstrated on both mammalian (cell line K562) and bacterial (E. coli ML-35) membranes [221,222]. In addition, experiments where defensins were exposed to artificial membranes has shown that channels were formed when negative potential had been applied on the opposite site to a defensin containing solvent [214]. A plethora of evidence supporting various proposed models of specific membrane disruption mechanism exist in the literature. The following section should give a brief overview of some of the acknowledged models.
One of the first models for pore formation mechanism proposed for antimicrobial peptides is the barrel-stave model [223]. This model has been extensively studied and describes a scenario where after initial binding of the peptides to the membrane, they align perpendicularly to the membrane and aggregate on the surface leading to formation of channels or pores (Figure 8). It has been demonstrated that antimicrobial peptides with defined secondary structures use this mechanism as their hydrophobic parts interact with the lipids of the membrane and the hydrophilic part line the lumen of the pore [105,224]. Alamethicin (Ac-Aib-Pro-Aib-Ala-Aib-Ala-Gln-Aib-Val-Aib-Gly-Leu-Aib-Pro-Val-Aib-Aib-Glu-Gln-Phl) [225] is an antimicrobial peptide that has been extensively studied for its ability to disrupt membranes using this model [226]. Recently, Bobone et al. demonstrated and proposed that trichogin GA IV (n-Oct-Aib-Gly-Leu-Aib-Gly-Gly-Leu-Aib-Gly-Ile-Lol) [227] and similarly short peptaibols make use of barrel-stave model to exert their membrane disruption abilities [217]. The two-state model is another perspective on how peptides interact with membranes [211]. Here, the first physical state of interaction is the peptide binding to the membranes (as explained in the earlier section) followed by multi-pore formation at a threshold concentration (or ratio) of peptide to lipid. Furthermore, it has been shown that the lipid composition of the cell membrane is the main determinant for the susceptibility of the cell to an antimicrobial peptide and not the peptides binding affinity [211]. The third interpretation of membrane disruption by antimicrobial peptides supports the toroidal pore model where the peptides cause the membrane to bend inwards so that the pores formed consist of peptides and hydrophilic lipid head groups (Figure 8) [228]. Disordered toroidal pore models have been observed by Sengupta et al. using molecular dynamics simulations for melittin and DPPC membranes where peptides were observed to locate near the pore center, while other molecules were positioned close to the pore edge (Figure 8) [229]. The forth model which involves more dynamic disruption of the membrane is the carpet model (Figure 8). Here the peptides bind parallel to the membrane and upon reaching a certain threshold concentration they break the membrane into micelles, resembling a detergent–like mechanism of membrane permeabilization [230,231]. Recently, this model has been proposed for an analogue of PMAP-23 peptide (cathelicidin; RIIDLLWRVRRPQKPKFVTVWV) on its action towards E. coli [210]. Killing of the bacteria had only taken place when a complete saturation of the membrane with PMAP-23 molecules occurred, indicating compatible mechanism of membrane destabilization with the carpet model as it has been described using artificial membrane systems [210].
Five other models are suggested by different research groups and those include; interfacial activity [2,232], sinking raft [233], leaky slit [234], lipid clustering [235] and sand in a gearbox [236] models. Reflecting on these conflicting proposed models, one should keep in mind that the challenge to provide universally accepted knowledge regarding precise mechanism of action lies in the different techniques that are applied to this complexity.

8.4. Other Mechanisms and Intracellular Antibacterial Targets

Despite the ability to successfully lyse bacterial membranes, there are antimicrobial peptides that can effectively cross the membrane barrier without disrupting it and exhibit their antimicrobial modes of action through intracellular targeting. There is increasing evidence reported in the literature for such antimicrobial peptides and some are described in this section. Human neutrophil peptide 1 (HNP1) and human-β defensin 3 (HBD3) have been reported to bind lipid II, a bacterial cell wall precursor [104,237]. Human β-defensin 3 also affects the electron transport in S. aureus [238]. Many antimicrobial peptides have immunomodualtory functions. One example are the human neutrophil peptides 1-3 that when released by tissue of invading granulocytes, trigger secretion of tumor necrosis factor (TNF-α) and interferon-gamma (IFN-γ) from macrophages. This enhances the clearance of bacteria as observed in a murine in vivo model [239]. Human β-defensin 3 also activates specialized antigen presenting cells (monocytes, dendritic cells) thus stimulating the adaptive immune system [240]. Buforin II, a 21-amino acid peptide (TRSSRAGLQFPVGRVHRLLRK) [241] with potent broad spectrum antimicrobial activity, is also able to traverse the cell membrane and inhibit cellular function by binding to DNA and RNA of the cells, resulting in rapid cell death [242]. Most recently, it has been documented that indolicidin binds double stranded DNA, thereby inhibiting DNA replication and transcription [39]. Another non-lytic antimicrobial peptides is the small proline-rich apidaecin peptide that kills bacteria by binding to a cytoplasmic target, most likely DnaK [243].
Figure 8. Proposed pathways for peptide interaction and disruption of lipid bilayers. The schematic overview illustrates some of the most well acknowledged pathways. After initial peptide lipid interaction (A) different models have been proposed, depending on e.g. peptide type and peptide-lipid ratios. (B) The barrel stave model, (C) the toroidal pore model, (D) disordered toroidal pores and (E) the carpet or detergent model. For other suggested models the reader are refereed to [2,232].
Figure 8. Proposed pathways for peptide interaction and disruption of lipid bilayers. The schematic overview illustrates some of the most well acknowledged pathways. After initial peptide lipid interaction (A) different models have been proposed, depending on e.g. peptide type and peptide-lipid ratios. (B) The barrel stave model, (C) the toroidal pore model, (D) disordered toroidal pores and (E) the carpet or detergent model. For other suggested models the reader are refereed to [2,232].
Pharmaceuticals 08 00366 g008

9. Mechanism of Bacterial Resistance towards Antimicrobial Peptides

Antibiotic treatments to fight various infections have greatly contributed to increased human life expectancy throughout the years. However, only a year after the discovery and the commercial use of penicillin, resistant strains were already isolated due to high selection pressure and rapid resistance development. The same applies to other antibiotics to which many bacteria have developed resistance, thus there are many infections today that lack treatment options. Antimicrobial peptides hold potential for being part of the novel anti-infective strategies. The majority of the peptides act on the microbial membrane and not on specific intracellular targets, with a mechanism less prone to development of resistance. However, to resist the action of antimicrobial peptides, bacteria have evolved various strategies. These include, proteolytic degradation, shielding bacterial cell surface, surface modification of membrane structure, active efflux and down-regulation of antimicrobial peptide expression [244]. Proteolytic degradation has been reported for LL-37 and β-defensin 2, by proteases from P. mirabilis [245] in addition to many proteases secreted by bacteria that cleave peptides after specific amino acid residues. In Gram-negative bacteria, specifically, many of the proteases that inactivate antimicrobial peptides are found at the outer membrane [246]. Extracellular structures such as capsule polysaccharides, fimbriae, exopolysaccharides and O-polysaccharide of LPS aid bacteria by binding antimicrobial peptides and thereby reducing the amount of peptide reaching the bacterial membrane [247,248,249,250,251]. Addition of positively charged groups or removal of phosphate group of lipid A, thereby neutralizing and decreasing the negative charge that attracts cationic antimicrobial peptides, respectively, has been reported as a strategy to resist the action of polymyxin B (Figure 6) [252,253]. Decreased overall negative charge has further been reported as a resistance mechanism for cationic peptides and related mimetics [254,255,256,257]. Gram-positive bacteria also resist the action of antimicrobial peptides also by enzymatic degradation by extracellular proteases [258] and cell wall and membrane modifications. The latter is prevalent by deacetylation of N-acetylglucosamine or O-acetylation of N-acetylmuramyl residues in the cell wall and alternation of charge by d-analynation of the lipoteichoic acids [259]. One example is d-alanylation of anionic lipoteichoic acids in group B Streptococcus via increasing the cell wall density and thereby reducing the penetration of cationic antimicrobial peptides [260]. Another way to resist antimicrobial peptide activity is to pump these antimicrobials out of the bacterial cytoplasm. Bacteria use import and export pumps for the movement of different molecules across their membrane. By using bacterial strains with deleted or inactive pumps, it has been demonstrated that these mutant strains are significantly more susceptibility of antimicrobial peptide killing [261,262]. In parallel to regulation of its own defense strategies, several bacteria also produce toxins that affect host recognition of bacteria. e.g., The exotoxins of Vibrio cholerae and enterotoxigenic E. coli have both been reported to be involved in down-regulation of antimicrobial peptides, i.e., LL-37 and hBD1, expression by host cells, although the precise mechanism behind this still is unknown [263]. In summary, bacteria as any living organisms, will always adjust to a selective pressure through various modes of resistance. The chemical evolution of antimicrobial peptides has proven to counteract some of the resistance mechanisms through various structural modifications and thus keeps antimicrobial peptides in the game as potential antimicrobial alternatives to the conventional antibiotics.

10. Peptidomimetics for Antimicrobial Research

In the continuous research for novel antimicrobial drugs, antimicrobial peptides serve as a pharmacophore for structural optimization. In order to retain the activity and selectivity of antimicrobial peptides, while improving bioavailability, metabolic stability and immunogenicity, researchers has challenged the peptide structures with various diverse modifications. Any compound that is able to imitate the structural properties and/or biological activities of a peptide is referred to as a peptidomimetic. Modifications of peptide structures in antimicrobial research involve backbone and/or side chain modifications. Some examples include incorporation of unnatural amino acids (e.g., d-amino acids), β-peptides, peptoids (N-substituted glycines), hybrid peptide-peptidomimetic structures, lipidation etc. (Figure 9). This section highlights some of the most common peptidomimetic strategies in antimicrobial drug research and development.
Figure 9. Current peptidomimetic structures with potent antimicrobial activity. The schematic illustrates the different backbone compositions of a variety of peptide and peptidomimetic structures that has been proven to possess antibacterial properties.
Figure 9. Current peptidomimetic structures with potent antimicrobial activity. The schematic illustrates the different backbone compositions of a variety of peptide and peptidomimetic structures that has been proven to possess antibacterial properties.
Pharmaceuticals 08 00366 g009

10.1. Unnatural Amino Acid Sequences

Some of the different strategies to modify and improve the antimicrobial properties of antimicrobial peptides include substitution of the natural L-amino acids with unnatural analogs. However, one should keep in mind that the unnatural amino acids are not always synthetic and thus several different types are found in nature. One example is the natural derivatives of L-proline that are found in leucinostatine, a peptide antibiotic produced by an endophytic fungus of European yew (Taxus baccata), which has demonstrated broad spectrum antimicrobial activities [264]. Structural modification of the side chains of arginine and lysine has also been investigated in a synthetic library of tritpticin (VRRFPWWWPFLRR)-derived peptides. By methylation of the side chain and also alternation of the length, it was demonstrated that the antimicrobial activity of the peptides could be slightly improved while the hemolytic activity was decreased when compared to the parent tritpticin peptide [265].
Another very common modification strategy used to improve on traditional antimicrobial peptides has been to change from l- to d-amino acids. d-Amino acids are very rare in Nature and incorporation of the D-amino acids changes the side chain and the backbone properties of the peptide. In addition, retro-inverso peptides have also been introduced, representing a reversed peptide sequences from N- to C-terminus exchanging l-amino acids to d-amino acids [266]. One of the first examples of a peptidomimetic is the all-d-magainin peptide, however this peptide failed to show any significant improvement of the antimicrobial activity when compared to the all-l-enantiomer. However, all-d-magainin demonstrated high resistance to proteolysis and exhibited no hemolytic activity, two properties which are considered valuable for therapeutic applications [267].

10.2. Addition of Lipid Moieties

Natural lipopeptides are well characterized in the literature as promising antimicrobial compounds, often produced by different bacterial strains to give an advantage over other closely related strains. This family of bacterial compounds generally includes cationic, cyclic compounds, such as polymyxins (B and E), daptomycin, lipopeptaibol and others (Figure 6) [268]. Consequently, addition of an aliphatic chain to the N-terminus of an active antimicrobial peptide can be a strategy to improve the peptides overall antimicrobial properties [269,270]. Cudic et al. have actively investigated novel lipopeptides by incorporating synthetic analogues to structures of known antimicrobials [145]. To exemplify, they recently designed a novel cyclic lipopeptides (cyclo-[D-Ala-(12-guanidino-dodecanoyl)Thr-D-Val-Val-DaThr-D-Asn]) (Figure 6) derived from fusaricidin, which proved to inhibit both growth of S. aureus biofilms in vitro and proliferation of S. aureus in vivo [271]. By using combinations of hydrocarbon tails and other hydrophobic moieties in conjugation with polylysine and lysine analogues, Ahn et al. [272] have demonstrated successful design of very simple antimicrobial peptides, with no detectable hemolytic activity [272]. Though their precise mode of action still needs to be characterized, they present convincing data demonstrating that the peptides work through different mechanisms than the classical membrane targeting peptides like melittin [272].

10.3. β-Peptidomimetics

The modification achieved in β-peptidomimetics involves a change in the backbone of the natural peptide structures without changing the side chain chemistry, i.e., addition of one, two or three carbons along the peptide chain (Figure 9). A small library of highly potent, short β-peptidomimetics has been synthesized by Hansen et al. [273]. The library was designed to mimic the minimal pharmacophores model for ultra-short peptides with activity against S. aureus. The β-peptidomimetics consisted of different lipophilic β2,2-amino acid coupled to a C-terminal amidated L-Arg residue. The most potent peptidomimetic from this library displayed MIC values of 2.7–7.2 µM against S. aureus, methicillin-resistant S. aureus and S. epidermidis, as well as and E. coli, thus presenting one promising class of antimicrobial agents with improved enzymatic stability and a low cost production [273]. Another interesting novel class of active peptidomimetics has been introduced by Mosca et al. [274]. This class presents amphiphilic cationic β3R3-peptides. These peptides demonstrate high selectivity with low cytotoxicity profile and thus illustrate a class of antimicrobial candidates with high therapeutic index with potential for further development.

10.4. Peptoids

Peptoids are oligomers of N-substituted glycines. They comprise a new class of unnatural compounds that mimic peptide structures by relocating the side chain from the α-carbon to the nitrogen (Figure 9). Structurally, the change in the amide bond corresponds to a loss of backbone chirality which can be compensated for by introducing chiral side chains that allow peptoids to fold into stable secondary structures [275,276,277]. Two of the most appreciated advantages with the backbone modification in peptoids are that side chains appended at the nitrogens render them less prone to enzymatic and proteolytic degradation and they are often more membrane permeable than peptides [278]. In this manner, Bang et al. used the peptoid conversion strategy of the novel tryptophan-rich model peptide to increase its protease stability while retaining its biological activity [279]. For more than 15 years, the research in peptoid synthesis and application increased dramatically due to their potential as compounds with broad antimicrobial activity profiles [280]. Their synthesis comprises a two steps process in which various side chains of commercial availability are incorporated and this versatility enables peptoids to hold status as promising structures for antimicrobial drug development. The careful work by the Barron group provides important information on several helical, cationic, facially amphipathic peptoid mimics of magainin-2 amide. Certain analogues exhibited potent antimicrobial activities against both Gram-negative and Gram-positive bacteria with low toxicity against human red blood cells [281]. In addition, application of peptoids as effective antimicrobials against Mycobacterium tuberculosis and inhibitors of biofilm formation by P. aeruginosa has been reported [282,283]. High synergistic interactions between nine antimicrobial peptides and peptoids have been further demonstrated suggesting that these two classes of antimicrobials are functionally and mechanistically analogous [284]. Recently another study demonstrated synthesis of peptoids that mimic the structure of antimicrobial peptides while retaining good potency against broad spectrum bacterial and low toxicity against human cells [285]. Overall, the important and conserved properties of antimicrobial peptides are also seen in peptoids and as such provide adequate information for future design of potent antimicrobials.

10.5. Cyclic Peptoids

Another structural challenge that has been overcome by Kirshenbaum et al. is cyclization of peptoids via a head to tail macrocyclization reaction [286]. Soon after an efficient synthetic approach was established, a comparison of linear and cyclic peptoids (6-10 residues) with potent antimicrobial activity demonstrated that the cyclic counterparts exhibited higher antimicrobial activities with no significant lysis of human erythrocytes [287]. New studies have shown that cyclic peptoids can damage methicillin-resistant S. aureus membranes trough pore formation mechanism with low MIC and low hemolytic activity [288]. Furthermore, these study challenge many researches in the peptoid field to embrace the cyclization awareness in developing peptoids as candidates for antimicrobial drugs.

10.6. Hybrids

Another attractive approach in peptidomimetics synthesis is merging structures of native peptides and their mimics which results in hybrid structures. For example, substituting arginine/leucine residues in apidaecin Ib with peptoid residues has shown to circumvent the resistance problem without any significant cytotoxic effects, however this has led to reduced antimicrobial activity at specific positional substitutions due to the inability of the novel peptide-peptoid hybrid to translocate into bacterial cells [289]. Research on piscidin 1, a novel cytotoxic peptide with cationic α-helical structure, provided support of an increased antimicrobial activity and lower cytotoxicity in mammalian cells by substitution of Pro8 with lysine mimicking peptoid residues (Nlys). This modification provided structural flexibility of the novel compound that exhibited better membrane permeability in bacteria conferring better overall selectivity [290]. The study of antimicrobial profiles of peptoid hybrids draws on research conducted by Franzyk et al. which described the first generation of oligomers consisting of alternating repeats of α-amino acids and β-peptoid residues. Such structures showed stability toward proteolysis, and exhibited high antimicrobial and no hemolytic activities. Additionally, more in depth evaluation of the synthesized library of hybrids challenged the importance of the length, choice of cationic side chain, presence of chiral side chains as well as lipophilicity on both the antimicrobial and hemolytic activities observed by this library of compounds [291,292,293,294]. Hansen et al. demonstrated synthesis of 20 novel lysine-peptoid hybrids designed on the basis of an active parent structure [N-(1-naphahalenemethyl)glycyl] - [N-(4-methylbenzyl)glycyl] - [N-(1-naphthalenemethyl)glycyl]-N-(butyl) -glycine amide 1, with antimicrobial properties against both Gram-negative and Gram-positive bacteria [295]. Recently, more profound information on HDM-4 peptidomimetic, composed of repeating units of lysine and NPhe (benzylamine), was demonstrated. This peptidomimetic exhibits low toxicity against mammalian cells and kills Gram-negative bacteria by membrane disruption. DNA binding, induction of chemokine production in immune cells and inhibition of LPS induced pro-inflammatory response has also been reported for this peptidomimetic [296]. In parallel to these observations, similar behaviour for DNA binding of peptide-peptoid hybrid has been established for lysine-peptoid hybrid, LP5. LP5 inhibited the growth of S. aureus in vitro by binding to DNA and inhibiting macromolecular synthesis. In addition it inhibited DNA gyrase and topoisomerase IV causing SOS response [297]. In summary, synthetic analogues and mimetics of antimicrobial peptides are being actively developed and studied to improve the antimicrobial and pharmacokinetic properties and lower the cost of production.

10.7. AApeptides

Another class of peptidomimetics with broad spectrum antimicrobial activity are AApeptides. This class of antimicrobials have been developed by Cai et al. [298] based on the chiral peptide nucleic acid backbone (Figure 9). Various structural modifications such as lipidation and cyclization have been illustrated for AApeptides and these new structures have been shown to be active against a range of community-acquired multidrug resistant pathogens [299,300,301].

11. Towards the Design of Novel Antimicrobial Peptides

The main goal of all researchers in the field of antimicrobial peptides is to contribute to the antimicrobial peptide range with novel strategies for the development of pharmacologically relevant antimicrobial peptide structures. The design of de novo structures is not always evident and it demands extensive experimental work. To meet this challenge and aid the discovery of de novo peptides, various computer-based strategies have been employed. Most involve building quantitative structure-activity relationship parameters for computer aided design of antimicrobial peptides. The general idea is to combine chemical sequence and structure of antimicrobial peptides, described with physiochemical parameters (descriptors) and correlate them with their respective biological activities using mathematical models [304,305]. One of the mathematical models used for prediction of novel polypeptide sequences with potentially improved biological activities is the Designer algorithm [306,307]. Using the Designer algorithm the most recent work of Ilic et al. reports on a series of peptide sequences with high antibacterial activity against Gram-negative bacteria (0.5–4 µM) and low hemolytic properties (HC50 > 400 µM) [308]. Fjell et al. have also created a software system that could identify peptides with antibacterial activity with up to 94% accuracy [309,310]. As a new method that further improved the identification of novel antimicrobials, they describe the programming method of genetic algorithms, improving on the number of active peptides within a semi-random library [311]. In summary, this kind of models can be useful for prediction of antimicrobial peptides with potentially higher biological activities; however the challenge to develop a robust model still prevails.

12. Concluding Remarks

There is increased evidence of emergence of bacteria resistant to conventional antibiotics illustrating the importance of research on antimicrobial drug development. Antimicrobial peptides and their synthetic derivatives hold vast potential in the development of novel antimicrobial drugs due to their (1) high biological activities; (2) low cost of production when compared with that of proteins and antibodies; (3) ease of structural modifications and stability improvements; (4) promising pharmacokinetic profiles; (5) degradation that leads to amino acids which are less toxic for the organism resulting in low immunogenicity and (6) good organ penetration.
The variety of structure-activity relationship studies for antimicrobial peptides demonstrate that the observed antibacterial activity is usually the outcome of multiple factors such as; secondary structures, amphipathicity, charge, length and hydrophobicity. The sequence composition together with the importance of specific residues and their contribution to the optimal therapeutic profiles of the peptides reveal important scaffolds for design of novel compounds to successfully combat and eliminate infectious diseases in the future.

Acknowledgments

BM holds a Ph.D. scholarship from The Danish Council for Independent Research (grant # 10-085287).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jenssen, H.; Hamill, P.; Hancock, R.E. Peptide antimicrobial agents. Clin. Microbiol. Rev. 2006, 19, 491–511. [Google Scholar] [CrossRef] [PubMed]
  2. Wimley, W.C. Describing the mechanism of antimicrobial peptide action with the interfacial activity model. ACS Chem. Biol. 2010, 5, 905–917. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, G. Database-guided discovery of potent peptides to combat hiv-1 or superbugs. Pharmaceuticals (Basel) 2013, 6, 728–758. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, G. Structures of human host defense cathelicidin ll-37 and its smallest antimicrobial peptide kr-12 in lipid micelles. J. Biol. Chem. 2008, 283, 32637–32643. [Google Scholar] [CrossRef] [PubMed]
  5. Powers, J.P.; Rozek, A.; Hancock, R.E. Structure-activity relationships for the beta-hairpin cationic antimicrobial peptide polyphemusin i. Biochim. Biophys. Acta 2004, 1698, 239–250. [Google Scholar] [CrossRef] [PubMed]
  6. Rozek, A.; Friedrich, C.L.; Hancock, R.E. Structure of the bovine antimicrobial peptide indolicidin bound to dodecylphosphocholine and sodium dodecyl sulfate micelles. Biochemistry 2000, 39, 15765–15774. [Google Scholar] [CrossRef] [PubMed]
  7. Sawai, M.V.; Jia, H.P.; Liu, L.; Aseyev, V.; Wiencek, J.M.; McCray, P.B., Jr.; Ganz, T.; Kearney, W.R.; Tack, B.F. The nmr structure of human beta-defensin-2 reveals a novel alpha-helical segment. Biochemistry 2001, 40, 3810–3816. [Google Scholar] [CrossRef] [PubMed]
  8. Andreu, D.; Rivas, L. Animal antimicrobial peptides: An overview. Biopolymers 1998, 47, 415–433. [Google Scholar] [CrossRef]
  9. Wang, G.; Li, X.; Wang, Z. Apd2: The updated antimicrobial peptide database and its application in peptide design. Nucleic Acids Res. 2009, 37, D933–D937. [Google Scholar] [CrossRef] [PubMed]
  10. Steiner, H.; Hultmark, D.; Engstrom, A.; Bennich, H.; Boman, H.G. Sequence and specificity of two antibacterial proteins involved in insect immunity. Nature 1981, 292, 246–248. [Google Scholar] [CrossRef] [PubMed]
  11. Pasupuleti, M.; Schmidtchen, A.; Malmsten, M. Antimicrobial peptides: Key components of the innate immune system. Crit. Rev. Biotechnol. 2012, 32, 143–171. [Google Scholar] [CrossRef] [PubMed]
  12. Boman, H.G. Peptide antibiotics and their role in innate immunity. Ann. Rev. immunol. 1995, 13, 61–92. [Google Scholar] [CrossRef] [PubMed]
  13. Thomas, E.L.; Lehrer, R.I.; Rest, R.F. Human neutrophil antimicrobial activity. Rev. Infect. Dis. 1988, 10, S450–S456. [Google Scholar] [CrossRef] [PubMed]
  14. Selsted, M.E.; Ouellette, A.J. Mammalian defensins in the antimicrobial immune response. Nat. Immunol. 2005, 6, 551–557. [Google Scholar] [CrossRef] [PubMed]
  15. Skerlavaj, B.; Romeo, D.; Gennaro, R. Rapid membrane permeabilization and inhibition of vital functions of gram-negative bacteria by bactenecins. Infect. Immun. 1990, 58, 3724–3730. [Google Scholar] [PubMed]
  16. Ganz, T.; Selsted, M.E.; Szklarek, D.; Harwig, S.S.; Daher, K.; Bainton, D.F.; Lehrer, R.I. Defensins. Natural peptide antibiotics of human neutrophils. J. Clin. Investig. 1985, 76, 1427–1435. [Google Scholar] [CrossRef] [PubMed]
  17. Selsted, M.E.; Harwig, S.S.; Ganz, T.; Schilling, J.W.; Lehrer, R.I. Primary structures of three human neutrophil defensins. J. Clin. Investig. 1985, 76, 1436–1439. [Google Scholar] [CrossRef] [PubMed]
  18. Ganz, T. Defensins: Antimicrobial peptides of innate immunity. Nature Rev. Immunol. 2003, 3, 710–720. [Google Scholar] [CrossRef] [PubMed]
  19. Bowdish, D.M.; Davidson, D.J.; Hancock, R.E. Immunomodulatory properties of defensins and cathelicidins. Curr. Top. Microbiol. Immunol. 2006, 306, 27–66. [Google Scholar] [PubMed]
  20. Pazgier, M.; Hoover, D.M.; Yang, D.; Lu, W.; Lubkowski, J. Human beta-defensins. Cell. Mol. Life Sci. 2006, 63, 1294–1313. [Google Scholar] [CrossRef] [PubMed]
  21. Ouellette, A.J.; Selsted, M.E. Paneth cell defensins: Endogenous peptide components of intestinal host defense. FASEB J. 1996, 10, 1280–1289. [Google Scholar] [PubMed]
  22. Date, Y.; Nakazato, M.; Shiomi, K.; Toshimori, H.; Kangawa, K.; Matsuo, H.; Matsukura, S. Localization of human neutrophil peptide (hnp) and its messenger rna in neutrophil series. Ann. Hematol. 1994, 69, 73–77. [Google Scholar] [CrossRef] [PubMed]
  23. Ericksen, B.; Wu, Z.; Lu, W.; Lehrer, R.I. Antibacterial activity and specificity of the six human {alpha}-defensins. Antimicrob. Agents Chemother. 2005, 49, 269–275. [Google Scholar] [CrossRef] [PubMed]
  24. Porter, E.M.; van Dam, E.; Valore, E.V.; Ganz, T. Broad-spectrum antimicrobial activity of human intestinal defensin 5. Infect. Immun. 1997, 65, 2396–2401. [Google Scholar] [PubMed]
  25. Xu, Z.; Zhong, Z.; Huang, L.; Peng, L.; Wang, F.; Cen, P. High-level production of bioactive human beta-defensin-4 in escherichia coli by soluble fusion expression. Appl. Microbiol. Biotechnol. 2006, 72, 471–479. [Google Scholar] [CrossRef] [PubMed]
  26. Nomura, I.; Goleva, E.; Howell, M.D.; Hamid, Q.A.; Ong, P.Y.; Hall, C.F.; Darst, M.A.; Gao, B.; Boguniewicz, M.; Travers, J.B.; et al. Cytokine milieu of atopic dermatitis, as compared to psoriasis, skin prevents induction of innate immune response genes. J. Immunol. 2003, 171, 3262–3269. [Google Scholar] [CrossRef] [PubMed]
  27. Lehrer, R.I.; Ganz, T. Cathelicidins: A family of endogenous antimicrobial peptides. Curr. Opin. Hematol. 2002, 9, 18–22. [Google Scholar] [CrossRef] [PubMed]
  28. Zaiou, M.; Gallo, R.L. Cathelicidins, essential gene-encoded mammalian antibiotics. J. Mol. Med. (Berl) 2002, 80, 549–561. [Google Scholar] [CrossRef] [PubMed]
  29. Scocchi, M.; Pallavicini, A.; Salgaro, R.; Bociek, K.; Gennaro, R. The salmonid cathelicidins: A gene family with highly varied c-terminal antimicrobial domains. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 2009, 152, 376–381. [Google Scholar] [CrossRef] [PubMed]
  30. Putsep, K.; Carlsson, G.; Boman, H.G.; Andersson, M. Deficiency of antibacterial peptides in patients with morbus kostmann: An observation study. Lancet 2002, 360, 1144–1149. [Google Scholar] [CrossRef]
  31. Szyk, A.; Wu, Z.; Tucker, K.; Yang, D.; Lu, W.; Lubkowski, J. Crystal structures of human alpha-defensins hnp4, hd5, and hd6. Protein Sci. 2006, 15, 2749–2760. [Google Scholar] [CrossRef] [PubMed]
  32. Hoover, D.M.; Chertov, O.; Lubkowski, J. The structure of human beta-defensin-1: New insights into structural properties of beta-defensins. J. Biol. Chem. 2001, 276, 39021–39026. [Google Scholar] [CrossRef] [PubMed]
  33. Trabi, M.; Schirra, H.J.; Craik, D.J. Three-dimensional structure of rtd-1, a cyclic antimicrobial defensin from rhesus macaque leukocytes. Biochemistry 2001, 40, 4211–4221. [Google Scholar] [CrossRef] [PubMed]
  34. Falla, T.J.; Karunaratne, D.N.; Hancock, R.E.W. Mode of action of the antimicrobial peptide indolicidin. J. Biol. Chem. 1996, 271, 19298–19303. [Google Scholar] [CrossRef] [PubMed]
  35. Selsted, M.E.; Novotny, M.J.; Morris, W.L.; Tang, Y.Q.; Smith, W.; Cullor, J.S. Indolicidin, a novel bactericidal tridecapeptide amide from neutrophils. J. Biol. Chem. 1992, 267, 4292–4295. [Google Scholar] [PubMed]
  36. Hsu, C.H.; Chen, C.; Jou, M.L.; Lee, A.Y.; Lin, Y.C.; Yu, Y.P.; Huang, W.T.; Wu, S.H. Structural and DNA-binding studies on the bovine antimicrobial peptide, indolicidin: Evidence for multiple conformations involved in binding to membranes and DNA. Nucleic Acids Res. 2005, 33, 4053–4064. [Google Scholar] [CrossRef] [PubMed]
  37. Subbalakshmi, C.; Sitaram, N. Mechanism of antimicrobial action of indolicidin. FEMS Microbiol. Lett. 1998, 160, 91–96. [Google Scholar] [CrossRef] [PubMed]
  38. Wu, M.; Maier, E.; Benz, R.; Hancock, R.E. Mechanism of interaction of different classes of cationic antimicrobial peptides with planar bilayers and with the cytoplasmic membrane of escherichia coli. Biochemistry 1999, 38, 7235–7242. [Google Scholar] [CrossRef] [PubMed]
  39. Ghosh, A.; Kar, R.K.; Jana, J.; Saha, A.; Jana, B.; Krishnamoorthy, J.; Kumar, D.; Ghosh, S.; Chatterjee, S.; Bhunia, A. Indolicidin targets duplex DNA: Structural and mechanistic insight through a combination of spectroscopy and microscopy. ChemMedChem 2014, 9, 2052–2058. [Google Scholar] [CrossRef] [PubMed]
  40. Ryge, T.S.; Doisy, X.; Ifrah, D.; Olsen, J.E.; Hansen, P.R. New indolicidin analogues with potent antibacterial activity. J. Pept. Res. 2004, 64, 171–185. [Google Scholar] [CrossRef] [PubMed]
  41. Friedrich, C.L.; Rozek, A.; Patrzykat, A.; Hancock, R.E. Structure and mechanism of action of an indolicidin peptide derivative with improved activity against gram-positive bacteria. J. Biol. Chem. 2001, 276, 24015–24022. [Google Scholar] [CrossRef] [PubMed]
  42. Falla, T.J.; Hancock, R.E. Improved activity of a synthetic indolicidin analog. Antimicrob. Agents Chemother. 1997, 41, 771–775. [Google Scholar] [PubMed]
  43. Rozek, A.; Powers, J.P.; Friedrich, C.L.; Hancock, R.E. Structure-based design of an indolicidin peptide analogue with increased protease stability. Biochemistry 2003, 42, 14130–14138. [Google Scholar] [CrossRef] [PubMed]
  44. Selsted, M.E.; Harwig, S.S. Determination of the disulfide array in the human defensin hnp-2. A covalently cyclized peptide. J. Biol. Chem. 1989, 264, 4003–4007. [Google Scholar] [PubMed]
  45. Tjabringa, G.S.; Rabe, K.F.; Hiemstra, P.S. The human cathelicidin ll-37: A multifunctional peptide involved in infection and inflammation in the lung. Pulm. Pharmacol. Ther. 2005, 18, 321–327. [Google Scholar] [CrossRef] [PubMed]
  46. Zhao, C.; Liu, L.; Lehrer, R.I. Identification of a new member of the protegrin family by cdna cloning. FEBS Lett. 1994, 346, 285–288. [Google Scholar] [CrossRef]
  47. Cole, A.M.; Waring, A.J. The role of defensins in lung biology and therapy. Am. J. Respir. Med. 2002, 1, 249–259. [Google Scholar] [CrossRef] [PubMed]
  48. Zasloff, M. Magainins, a class of antimicrobial peptides from xenopus skin: Isolation, characterization of two active forms, and partial cdna sequence of a precursor. Proc. Natl. Acad. Sci. USA 1987, 84, 5449–5453. [Google Scholar] [CrossRef] [PubMed]
  49. Steiner, H. Secondary structure of the cecropins: Antibacterial peptides from the moth hyalophora cecropia. FEBS Lett. 1982, 137, 283–287. [Google Scholar] [CrossRef]
  50. Dempsey, C.E. The actions of melittin on membranes. Biochim. Biophys. Acta 1990, 1031, 143–161. [Google Scholar] [CrossRef]
  51. Miyata, T.; Tokunaga, F.; Yoneya, T.; Yoshikawa, K.; Iwanaga, S.; Niwa, M.; Takao, T.; Shimonishi, Y. Antimicrobial peptides, isolated from horseshoe crab hemocytes, tachyplesin ii, and polyphemusins i and ii: Chemical structures and biological activity. J. Biochem. 1989, 106, 663–668. [Google Scholar] [PubMed]
  52. Gause, G.F.; Brazhnikova, M.G. Gramicin s - origin and mode of action. Lancet 1944, 2, 715–716. [Google Scholar]
  53. Gross, E.; Morell, J.L. The structure of nisin. J. Am. Chem. Soc. 1971, 93, 4634–4635. [Google Scholar] [CrossRef] [PubMed]
  54. Zanetti, M. Cathelicidins, multifunctional peptides of the innate immunity. J. Leukoc. Biol. 2004, 75, 39–48. [Google Scholar] [CrossRef] [PubMed]
  55. Gombart, A.F.; O'Kelly, J.; Saito, T.; Koeffler, H.P. Regulation of the camp gene by 1,25(oh)2d3 in various tissues. J. Steroid Biochem. Mol. Biol. 2007, 103, 552–557. [Google Scholar] [CrossRef] [PubMed]
  56. Gombart, A.F.; Borregaard, N.; Koeffler, H.P. Human cathelicidin antimicrobial peptide (camp) gene is a direct target of the vitamin d receptor and is strongly up-regulated in myeloid cells by 1,25-dihydroxyvitamin d3. FASEB J. 2005, 19, 1067–1077. [Google Scholar] [CrossRef] [PubMed]
  57. Carlberg, C. Current understanding of the function of the nuclear vitamin d receptor in response to its natural and synthetic ligands. Recent Results Cancer Res. 2003, 164, 29–42. [Google Scholar] [PubMed]
  58. Guo, C.; Rosoha, E.; Lowry, M.B.; Borregaard, N.; Gombart, A.F. Curcumin induces human cathelicidin antimicrobial peptide gene expression through a vitamin d receptor-independent pathway. J. Nutr. Biochem. 2012, 24, 754–759. [Google Scholar] [CrossRef] [PubMed]
  59. Barlow, P.G.; Li, Y.; Wilkinson, T.S.; Bowdish, D.M.; Lau, Y.E.; Cosseau, C.; Haslett, C.; Simpson, A.J.; Hancock, R.E.; Davidson, D.J. The human cationic host defense peptide ll-37 mediates contrasting effects on apoptotic pathways in different primary cells of the innate immune system. J. Leukoc. Biol. 2006, 80, 509–520. [Google Scholar] [CrossRef] [PubMed]
  60. Hase, K.; Murakami, M.; Iimura, M.; Cole, S.P.; Horibe, Y.; Ohtake, T.; Obonyo, M.; Gallo, R.L.; Eckmann, L.; Kagnoff, M.F. Expression of ll-37 by human gastric epithelial cells as a potential host defense mechanism against helicobacter pylori. Gastroenterology 2003, 125, 1613–1625. [Google Scholar] [CrossRef] [PubMed]
  61. Travis, S.M.; Anderson, N.N.; Forsyth, W.R.; Espiritu, C.; Conway, B.D.; Greenberg, E.P.; McCray, P.B., Jr.; Lehrer, R.I.; Welsh, M.J.; Tack, B.F. Bactericidal activity of mammalian cathelicidin-derived peptides. Infect. Immun. 2000, 68, 2748–2755. [Google Scholar] [CrossRef] [PubMed]
  62. Scott, M.G.; Davidson, D.J.; Gold, M.R.; Bowdish, D.; Hancock, R.E. The human antimicrobial peptide ll-37 is a multifunctional modulator of innate immune responses. J. Immunol. 2002, 169, 3883–3891. [Google Scholar] [CrossRef] [PubMed]
  63. Bowdish, D.M.; Davidson, D.J.; Lau, Y.E.; Lee, K.; Scott, M.G.; Hancock, R.E. Impact of ll-37 on anti-infective immunity. J. Leukoc. Biol. 2005, 77, 451–459. [Google Scholar] [CrossRef] [PubMed]
  64. Barlow, P.G.; Beaumont, P.E.; Cosseau, C.; Mackellar, A.; Wilkinson, T.S.; Hancock, R.E.; Haslett, C.; Govan, J.R.; Simpson, A.J.; Davidson, D.J. The human cathelicidin ll-37 preferentially promotes apoptosis of infected airway epithelium. Am. J. Respir. Cell. Mol. Biol. 2010, 43, 692–702. [Google Scholar] [CrossRef] [PubMed]
  65. Yu, J.; Mookherjee, N.; Wee, K.; Bowdish, D.M.; Pistolic, J.; Li, Y.; Rehaume, L.; Hancock, R.E. Host defense peptide ll-37, in synergy with inflammatory mediator il-1beta, augments immune responses by multiple pathways. J. Immunol. 2007, 179, 7684–7691. [Google Scholar] [CrossRef] [PubMed]
  66. Zheng, Y.; Niyonsaba, F.; Ushio, H.; Nagaoka, I.; Ikeda, S.; Okumura, K.; Ogawa, H. Cathelicidin ll-37 induces the generation of reactive oxygen species and release of human alpha-defensins from neutrophils. Br. J. Dermatol. 2007, 157, 1124–1131. [Google Scholar] [CrossRef] [PubMed]
  67. Bals, R.; Weiner, D.J.; Moscioni, A.D.; Meegalla, R.L.; Wilson, J.M. Augmentation of innate host defense by expression of a cathelicidin antimicrobial peptide. Infect. Immun. 1999, 67, 6084–6089. [Google Scholar] [PubMed]
  68. Niyonsaba, F.; Iwabuchi, K.; Someya, A.; Hirata, M.; Matsuda, H.; Ogawa, H.; Nagaoka, I. A cathelicidin family of human antibacterial peptide ll-37 induces mast cell chemotaxis. Immunology 2002, 106, 20–26. [Google Scholar] [CrossRef] [PubMed]
  69. Nijnik, A.; Pistolic, J.; Wyatt, A.; Tam, S.; Hancock, R.E. Human cathelicidin peptide ll-37 modulates the effects of ifn-gamma on apcs. J. Immunol. 2009, 183, 5788–5798. [Google Scholar] [CrossRef] [PubMed]
  70. Bulet, P.; Hetru, C.; Dimarcq, J.L.; Hoffmann, D. Antimicrobial peptides in insects; structure and function. Dev. Comp. Immunol. 1999, 23, 329–344. [Google Scholar] [CrossRef]
  71. Bulet, P.; Stocklin, R. Insect antimicrobial peptides: Structures, properties and gene regulation. Protein Pept. Lett. 2005, 12, 3–11. [Google Scholar] [CrossRef] [PubMed]
  72. Vizioli, J.; Bulet, P.; Charlet, M.; Lowenberger, C.; Blass, C.; Muller, H.M.; Dimopoulos, G.; Hoffmann, J.; Kafatos, F.C.; Richman, A. Cloning and analysis of a cecropin gene from the malaria vector mosquito, anopheles gambiae. Insect Mol. Biol. 2000, 9, 75–84. [Google Scholar] [CrossRef] [PubMed]
  73. Raghuraman, H.; Chattopadhyay, A. Melittin: A membrane-active peptide with diverse functions. Biosci. Rep. 2007, 27, 189–223. [Google Scholar] [CrossRef] [PubMed]
  74. Schroder, E.; Lubke, K.; Lehmann, M.; Beetz, I. Haemolytic activity and action on the surface tension of aqueous solutions of synthetic melittins and their derivatives. Experientia 1971, 27, 764–765. [Google Scholar] [CrossRef] [PubMed]
  75. Blondelle, S.E.; Houghten, R.A. Probing the relationships between the structure and hemolytic activity of melittin with a complete set of leucine substitution analogs. Peptide Res. 1991, 4, 12–18. [Google Scholar]
  76. Blondelle, S.E.; Houghten, R.A. Hemolytic and antimicrobial activities of the twenty-four individual omission analogues of melittin. Biochemistry 1991, 30, 4671–4678. [Google Scholar] [CrossRef] [PubMed]
  77. Cao, Y.; Yu, R.Q.; Liu, Y.; Zhou, H.X.; Song, L.L.; Qiao, D.R. Design, recombinant expression, and antibacterial activity of the cecropins-melittin hybrid antimicrobial peptides. Curr. Microbiol. 2010, 61, 169–175. [Google Scholar] [CrossRef] [PubMed]
  78. Ji, S.; Li, W.; Zhang, L.; Zhang, Y.; Cao, B. Cecropin a-melittin mutant with improved proteolytic stability and enhanced antimicrobial activity against bacteria and fungi associated with gastroenteritis in vitro. Biochem. Biophys. Res. Commun. 2014, 451, 650–655. [Google Scholar] [CrossRef] [PubMed]
  79. Marion, D.; Bakan, B.; Elmorjani, K. Plant lipid binding proteins: Properties and applications. Biotechnol. Adv. 2007, 25, 195–197. [Google Scholar] [CrossRef] [PubMed]
  80. Stotz, H.U.; Thomson, J.G.; Wang, Y. Plant defensins: Defense, development and application. Plant Signal. Behav. 2009, 4, 1010–1012. [Google Scholar] [CrossRef] [PubMed]
  81. Beintema, J.J. Structural features of plant chitinases and chitin-binding proteins. FEBS Lett. 1994, 350, 159–163. [Google Scholar] [CrossRef]
  82. Burman, R.; Gunasekera, S.; Stromstedt, A.A.; Goransson, U. Chemistry and biology of cyclotides: Circular plant peptides outside the box. J. Nat. Prod. 2014, 77, 724–736. [Google Scholar] [CrossRef] [PubMed]
  83. Stec, B. Plant thionins--the structural perspective. Cell. Mol. Life Sci. 2006, 63, 1370–1385. [Google Scholar] [CrossRef] [PubMed]
  84. Clifton, L.A.; Sanders, M.R.; Hughes, A.V.; Neylon, C.; Frazier, R.A.; Green, R.J. Lipid binding interactions of antimicrobial plant seed defence proteins: Puroindoline-a and beta-purothionin. Phys. Chem. Chem. Physics 2011, 13, 17153–17162. [Google Scholar] [CrossRef] [PubMed]
  85. Samuelsson, G.; Jayawardene, A.L. Isolation and characterization of viscotoxin 1-ps from viscum album l. Ssp. Austriacum (wiesb.) vollmann, growing on pinus silvestris. Acta Pharm. Suec. 1974, 11, 175–184. [Google Scholar] [PubMed]
  86. Giudici, M.; Pascual, R.; de la Canal, L.; Pfuller, K.; Pfuller, U.; Villalain, J. Interaction of viscotoxins a3 and b with membrane model systems: Implications to their mechanism of action. Biophys. J. 2003, 85, 971–981. [Google Scholar] [CrossRef]
  87. Stec, B.; Rao, U.; Teeter, M.M. Refinement of purothionins reveals solute particles important for lattice formation and toxicity. Part 2: Structure of beta-purothionin at 1.7 a resolution. Acta Crystallogr. D Biol. Crystallogr. 1995, 51, 914–924. [Google Scholar] [CrossRef] [PubMed]
  88. Bruix, M.; Jimenez, M.A.; Santoro, J.; Gonzalez, C.; Colilla, F.J.; Mendez, E.; Rico, M. Solution structure of gamma 1-h and gamma 1-p thionins from barley and wheat endosperm determined by 1h-nmr: A structural motif common to toxic arthropod proteins. Biochemistry 1993, 32, 715–724. [Google Scholar] [CrossRef] [PubMed]
  89. Milbradt, A.G.; Kerek, F.; Moroder, L.; Renner, C. Structural characterization of hellethionins from helleborus purpurascens. Biochemistry 2003, 42, 2404–2411. [Google Scholar] [CrossRef] [PubMed]
  90. Cotter, P.D.; Hill, C.; Ross, R.P. Bacteriocins: Developing innate immunity for food. Nat. Rev. Microbiol. 2005, 3, 777–788. [Google Scholar] [CrossRef] [PubMed]
  91. Rogers, L.A. The inhibiting effect of streptococcus lactis on lactobacillus bulgaricus. J. Bacteriol. 1928, 16, 321–325. [Google Scholar] [PubMed]
  92. Wiedemann, I.; Breukink, E.; van Kraaij, C.; Kuipers, O.P.; Bierbaum, G.; de Kruijff, B.; Sahl, H.G. Specific binding of nisin to the peptidoglycan precursor lipid ii combines pore formation and inhibition of cell wall biosynthesis for potent antibiotic activity. J. Biol. Chem. 2001, 276, 1772–1779. [Google Scholar] [CrossRef] [PubMed]
  93. Hasper, H.E.; de Kruijff, B.; Breukink, E. Assembly and stability of nisin-lipid ii pores. Biochemistry 2004, 43, 11567–11575. [Google Scholar] [CrossRef] [PubMed]
  94. Breukink, E.; Ganz, P.; de Kruijff, B.; Seelig, J. Binding of nisin z to bilayer vesicles as determined with isothermal titration calorimetry. Biochemistry 2000, 39, 10247–10254. [Google Scholar] [CrossRef] [PubMed]
  95. Hasper, H.E.; Kramer, N.E.; Smith, J.L.; Hillman, J.D.; Zachariah, C.; Kuipers, O.P.; de Kruijff, B.; Breukink, E. An alternative bactericidal mechanism of action for lantibiotic peptides that target lipid ii. Science 2006, 313, 1636–1637. [Google Scholar] [CrossRef] [PubMed]
  96. Chan, W.C.; Leyland, M.; Clark, J.; Dodd, H.M.; Lian, L.Y.; Gasson, M.J.; Bycroft, B.W.; Roberts, G.C. Structure-activity relationships in the peptide antibiotic nisin: Antibacterial activity of fragments of nisin. FEBS Lett. 1996, 390, 129–132. [Google Scholar] [CrossRef]
  97. Strother, T.; Hamers, R.J.; Smith, L.M. Covalent attachment of oligodeoxyribonucleotides to amine-modified si (001) surfaces. Nucleic Acids Res. 2000, 28, 3535–3541. [Google Scholar] [CrossRef] [PubMed]
  98. Hillman, J.D.; Novak, J.; Sagura, E.; Gutierrez, J.A.; Brooks, T.A.; Crowley, P.J.; Hess, M.; Azizi, A.; Leung, K.; Cvitkovitch, D.; et al. Genetic and biochemical analysis of mutacin 1140, a lantibiotic from streptococcus mutans. Infect. Immun. 1998, 66, 2743–2749. [Google Scholar] [PubMed]
  99. Fox, J.L. Antimicrobial peptides stage a comeback. Nat. Biotechnol. 2013, 31, 379–382. [Google Scholar] [CrossRef] [PubMed]
  100. Chen, C.; Fan, H.; Huang, Y.; Peng, F.; Yuan, S.; Tong, Y. Recombinant lysostaphin protects mice from methicillin-resistant staphylococcus aureus pneumonia. BioMed Res. Int. 2014, 2014, 602185. [Google Scholar] [PubMed]
  101. Hill, C.P.; Yee, J.; Selsted, M.E.; Eisenberg, D. Crystal structure of defensin hnp-3, an amphiphilic dimer: Mechanisms of membrane permeabilization. Science 1991, 251, 1481–1485. [Google Scholar] [CrossRef] [PubMed]
  102. Zhang, X.L.; Selsted, M.E.; Pardi, A. Nmr studies of defensin antimicrobial peptides. 1. Resonance assignment and secondary structure determination of rabbit np-2 and human hnp-1. Biochemistry 1992, 31, 11348–11356. [Google Scholar] [CrossRef] [PubMed]
  103. Pardi, A.; Zhang, X.L.; Selsted, M.E.; Skalicky, J.J.; Yip, P.F. Nmr studies of defensin antimicrobial peptides. 2. Three-dimensional structures of rabbit np-2 and human hnp-1. Biochemistry 1992, 31, 11357–11364. [Google Scholar] [CrossRef] [PubMed]
  104. De Leeuw, E.; Li, C.; Zeng, P.; Diepeveen-de Buin, M.; Lu, W.Y.; Breukink, E.; Lu, W. Functional interaction of human neutrophil peptide-1 with the cell wall precursor lipid ii. FEBS Lett. 2010, 584, 1543–1548. [Google Scholar] [CrossRef] [PubMed]
  105. Yeaman, M.R.; Yount, N.Y. Mechanisms of antimicrobial peptide action and resistance. Pharmacol. Rev. 2003, 55, 27–55. [Google Scholar] [CrossRef] [PubMed]
  106. Epand, R.M.; Shai, Y.; Segrest, J.P.; Anantharamaiah, G.M. Mechanisms for the modulation of membrane bilayer properties by amphipathic helical peptides. Biopolymers 1995, 37, 319–338. [Google Scholar] [CrossRef] [PubMed]
  107. Jahnsen, R.D.; Frimodt-Moller, N.; Franzyk, H. Antimicrobial activity of peptidomimetics against multidrug-resistant escherichia coli : A comparative study of different backbones. J. Med. Chem. 2012, 55, 7253–7261. [Google Scholar] [CrossRef] [PubMed]
  108. 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]
  109. 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] [PubMed]
  110. 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]
  111. Deslouches, B.; Phadke, S.M.; Lazarevic, V.; Cascio, M.; Islam, K.; Montelaro, R.C.; Mietzner, T.A. De novo generation of cationic antimicrobial peptides: Influence of length and tryptophan substitution on antimicrobial activity. Antimicrob. Agents Chemother. 2005, 49, 316–322. [Google Scholar] [CrossRef] [PubMed]
  112. Javadpour, M.M.; Juban, M.M.; Lo, W.C.; Bishop, S.M.; Alberty, J.B.; Cowell, S.M.; Becker, C.L.; McLaughlin, M.L. De novo antimicrobial peptides with low mammalian cell toxicity. J. Med. Chem. 1996, 39, 3107–3113. [Google Scholar] [CrossRef] [PubMed]
  113. Rajabi, M.; de Leeuw, E.; Pazgier, M.; Li, J.; Lubkowski, J.; Lu, W. The conserved salt bridge in human alpha-defensin 5 is required for its precursor processing and proteolytic stability. J. Biol. Chem. 2008, 283, 21509–21518. [Google Scholar] [CrossRef] [PubMed]
  114. Hunter, H.N.; Demcoe, A.R.; Jenssen, H.; Gutteberg, T.J.; Vogel, H.J. Human lactoferricin is partially folded in aqueous solution and is better stabilized in a membrane mimetic solvent. Antimicrob. Agents Chemother. 2005, 49, 3387–3395. [Google Scholar] [CrossRef] [PubMed]
  115. Wommack, A.J.; Robson, S.A.; Wanniarachchi, Y.A.; Wan, A.; Turner, C.J.; Wagner, G.; Nolan, E.M. Nmr solution structure and condition-dependent oligomerization of the antimicrobial peptide human defensin 5. Biochemistry 2012, 51, 9624–9637. [Google Scholar] [CrossRef] [PubMed]
  116. De Leeuw, E.; Rajabi, M.; Zou, G.; Pazgier, M.; Lu, W. Selective arginines are important for the antibacterial activity and host cell interaction of human alpha-defensin 5. FEBS Lett. 2009, 583, 2507–2512. [Google Scholar] [CrossRef] [PubMed]
  117. Wei, G.; de Leeuw, E.; Pazgier, M.; Yuan, W.R.; Zou, G.Z.; Wang, J.F.; Ericksen, B.; Lu, W.Y.; Lehrer, R.I.; Lu, W.Y. Through the looking glass, mechanistic insights from enantiomeric human defensins. J. Biol. Chem. 2009, 284, 29180–29192. [Google Scholar] [CrossRef] [PubMed]
  118. Pazgier, M.; Prahl, A.; Hoover, D.M.; Lubkowski, J. Studies of the biological properties of human beta-defensin 1. J. Biol. Chem. 2007, 282, 1819–1829. [Google Scholar] [CrossRef] [PubMed]
  119. Harder, J.; Bartels, J.; Christophers, E.; Schroder, J.M. A peptide antibiotic from human skin. Nature 1997, 387, 861–861. [Google Scholar] [CrossRef] [PubMed]
  120. Schibli, D.J.; Hunter, H.N.; Aseyev, V.; Starner, T.D.; Wiencek, J.M.; McCray, P.B.; Tack, B.F.; Vogel, H.J. The solution structures of the human beta-defensins lead to a better understanding of the potent bactericidal activity of hbd3 against staphylococcus aureus. J. Biol. Chem. 2002, 277, 8279–8289. [Google Scholar] [CrossRef] [PubMed]
  121. Dathe, M.; Nikolenko, H.; Meyer, J.; Beyermann, M.; Bienert, M. Optimization of the antimicrobial activity of magainin peptides by modification of charge. FEBS Lett. 2001, 501, 146–150. [Google Scholar] [CrossRef]
  122. Hilpert, K.; Elliott, M.R.; Volkmer-Engert, R.; Henklein, P.; Donini, O.; Zhou, Q.; Winkler, D.F.H.; Hancock, R.E.W. Sequence requirements and an optimization strategy for short antimicrobial peptides. Chem. Biol. 2006, 13, 1101–1107. [Google Scholar] [CrossRef] [PubMed]
  123. Sawyer, J.G.; Martin, N.L.; Hancock, R.E.W. Interaction of macrophage cationic proteins with the outer-membrane of pseudomonas-aeruginosa. Infect. Immun. 1988, 56, 693–698. [Google Scholar] [PubMed]
  124. Piers, K.L.; Hancock, R.E.W. The interaction of a recombinant cecropin/melittin hybrid peptide with the outer-membrane of pseudomonas-aeruginosa. Mol. Microbiol. 1994, 12, 951–958. [Google Scholar] [CrossRef] [PubMed]
  125. Matsuzaki, K. Control of cell selectivity of antimicrobial peptides. Biochim. Biophys. Acta 2009, 1788, 1687–1692. [Google Scholar] [CrossRef] [PubMed]
  126. Yin, L.M.; Edwards, M.A.; Li, J.; Yip, C.M.; Deber, C.M. Roles of hydrophobicity and charge distribution of cationic antimicrobial peptides in peptide-membrane interactions. J. Biol. Chem. 2012, 287, 7738–7745. [Google Scholar] [CrossRef] [PubMed]
  127. Wieprecht, T.; Dathe, M.; Krause, E.; Beyermann, M.; Maloy, W.L.; MacDonald, D.L.; Bienert, M. Modulation of membrane activity of amphipathic, antibacterial peptides by slight modifications of the hydrophobic moment. FEBS Lett. 1997, 417, 135–140. [Google Scholar] [CrossRef]
  128. Gopal, R.; Seo, C.H.; Song, P.I.; Park, Y. Effect of repetitive lysine-tryptophan motifs on the bactericidal activity of antimicrobial peptides. Amino acids 2013, 44, 645–660. [Google Scholar] [CrossRef] [PubMed]
  129. Eisenberg, D. Three-dimensional structure of membrane and surface proteins. Ann. Rev. Biochem. 1984, 53, 595–623. [Google Scholar] [CrossRef] [PubMed]
  130. Fernandez-Vidall, M.; Jayasinghe, S.; Ladokhin, A.S.; White, S.H. Folding amphipathic helices into membranes: Amphiphilicity trumps hydrophobicity. J. Mol. Biol. 2007, 370, 459–470. [Google Scholar] [CrossRef] [PubMed]
  131. Chen, Y.; Mant, C.T.; Farmer, S.W.; Hancock, R.E.; Vasil, M.L.; Hodges, R.S. Rational design of alpha-helical antimicrobial peptides with enhanced activities and specificity/therapeutic index. J. Biol. Chem. 2005, 280, 12316–12329. [Google Scholar] [CrossRef] [PubMed]
  132. Thaker, H.D.; Cankaya, A.; Scott, R.W.; Tew, G.N. Role of amphiphilicity in the design of synthetic mimics of antimicrobial peptides with gram-negative activity. ACS Med. Chem. Lett. 2013, 4, 481–485. [Google Scholar] [CrossRef] [PubMed]
  133. Wessolowski, A.; Bienert, M.; Dathe, M. Antimicrobial activity of arginine- and tryptophan-rich hexapeptides: The effects of aromatic clusters, d-amino acid substitution and cyclization. J. Pept. Res. 2004, 64, 159–169. [Google Scholar] [CrossRef] [PubMed]
  134. Pathak, N.; Salas-Auvert, R.; Ruche, G.; Janna, M.H.; McCarthy, D.; Harrison, R.G. Comparison of the effects of hydrophobicity, amphiphilicity, and alpha-helicity on the activities of antimicrobial peptides. Proteins 1995, 22, 182–186. [Google Scholar] [CrossRef] [PubMed]
  135. Andreu, D.; Ubach, J.; Boman, A.; Wahlin, B.; Wade, D.; Merrifield, R.B.; Boman, H.G. Shortened cecropin a-melittin hybrids. Significant size reduction retains potent antibiotic activity. FEBS Lett. 1992, 296, 190–194. [Google Scholar] [CrossRef]
  136. Kondejewski, L.H.; Jelokhani-Niaraki, M.; Farmer, S.W.; Lix, B.; Kay, C.M.; Sykes, B.D.; Hancock, R.E.W.; Hodges, R.S. Dissociation of antimicrobial and hemolytic activities in cyclic peptide diastereomers by systematic alterations in amphipathicity. J. Biol. Chem. 1999, 274, 13181–13192. [Google Scholar] [CrossRef] [PubMed]
  137. Kushibiki, T.; Kamiya, M.; Aizawa, T.; Kumaki, Y.; Kikukawa, T.; Mizuguchi, M.; Demura, M.; Kawabata, S.; Kawano, K. Interaction between tachyplesin i, an antimicrobial peptide derived from horseshoe crab, and lipopolysaccharide. Biochim. Biophys. Acta 2014, 1844, 527–534. [Google Scholar] [CrossRef] [PubMed]
  138. Fahrner, R.L.; Dieckmann, T.; Harwig, S.S.; Lehrer, R.I.; Eisenberg, D.; Feigon, J. Solution structure of protegrin-1, a broad-spectrum antimicrobial peptide from porcine leukocytes. Chem. Biol. 1996, 3, 543–550. [Google Scholar] [CrossRef]
  139. Bagheri, M.; Keller, S.; Dathe, M. Interaction of w-substituted analogs of cyclo-rrrwfw with bacterial lipopolysaccharides: The role of the aromatic cluster in antimicrobial activity. Antimicrob. Agents Chemother. 2011, 55, 788–797. [Google Scholar] [CrossRef] [PubMed]
  140. Scheinpflug, K.; Nikolenko, H.; Komarov, I.V.; Rautenbach, M.; Dathe, M. What goes around comes around-a comparative study of the influence of chemical modifications on the antimicrobial properties of small cyclic peptides. Pharmaceuticals (Basel) 2013, 6, 1130–1144. [Google Scholar] [CrossRef] [PubMed]
  141. Kondejewski, L.H.; Farmer, S.W.; Wishart, D.S.; Hancock, R.E.; Hodges, R.S. Gramicidin s is active against both gram-positive and gram-negative bacteria. Int. J. Pept. Protein Res. 1996, 47, 460–466. [Google Scholar] [CrossRef] [PubMed]
  142. Kondejewski, L.H.; Farmer, S.W.; Wishart, D.S.; Kay, C.M.; Hancock, R.E.; Hodges, R.S. Modulation of structure and antibacterial and hemolytic activity by ring size in cyclic gramicidin s analogs. J. Biol. Chem. 1996, 271, 25261–25268. [Google Scholar] [CrossRef] [PubMed]
  143. Appelt, C.; Wessolowski, A.; Soderhall, J.A.; Dathe, M.; Schmieder, P. Structure of the antimicrobial, cationic hexapeptide cyclo(rrwwrf) and its analogues in solution and bound to detergent micelles. Chembiochem 2005, 6, 1654–1662. [Google Scholar] [CrossRef] [PubMed]
  144. Trabi, M.; Craik, D.J. Circular proteins--no end in sight. Trends Biochem. Sci. 2002, 27, 132–138. [Google Scholar] [CrossRef]
  145. Bionda, N.; Stawikowski, M.; Stawikowska, R.; Cudic, M.; Lopez-Vallejo, F.; Treitl, D.; Medina-Franco, J.; Cudic, P. Effects of cyclic lipodepsipeptide structural modulation on stability, antibacterial activity, and human cell toxicity. ChemMedChem 2012, 7, 871–882. [Google Scholar] [CrossRef] [PubMed]
  146. Liu, Z.G.; Brady, A.; Young, A.; Rasimick, B.; Chen, K.; Zhou, C.H.; Kallenbach, N.R. Length effects in antimicrobial peptides of the (rw)(n) series. Antimicrob. Agents Chemother. 2007, 51, 597–603. [Google Scholar] [CrossRef] [PubMed]
  147. Niidome, T.; Matsuyama, N.; Kunihara, M.; Hatakeyama, T.; Aoyagi, H. Effect of chain length of cationic model peptides on antibacterial activity. Bull. Chem. Soc. Jpn. 2005, 78, 473–476. [Google Scholar] [CrossRef]
  148. Dong, N.; Ma, Q.; Shan, A.; Lv, Y.; Hu, W.; Gu, Y.; Li, Y. Strand length-dependent antimicrobial activity and membrane-active mechanism of arginine- and valine-rich beta-hairpin-like antimicrobial peptides. Antimicrob. Agents Chemother. 2012, 56, 2994–3003. [Google Scholar] [CrossRef] [PubMed]
  149. Wang, G.; Mishra, B.; Epand, R.F.; Epand, R.M. High-quality 3d structures shine light on antibacterial, anti-biofilm and antiviral activities of human cathelicidin ll-37 and its fragments. Biochim. Biophys. Acta 2014, 1838, 2160–2172. [Google Scholar] [CrossRef] [PubMed]
  150. Shepherd, C.M.; Schaus, K.A.; Vogel, H.J.; Juffer, A.H. Molecular dynamics study of peptide-bilayer adsorption. Biophys. J. 2001, 80, 579–596. [Google Scholar] [CrossRef]
  151. Aliste, M.P.; MacCallum, J.L.; Tieleman, D.P. Molecular dynamics simulations of pentapeptides at interfaces: Salt bridge and cation-pi interactions. Biochemistry 2003, 42, 8976–8987. [Google Scholar] [CrossRef] [PubMed]
  152. Zou, G.Z.; de Leeuw, E.; Li, C.; Pazgier, M.; Li, C.Q.; Zeng, P.Y.; Lu, W.Y.; Lubkowski, J.; Lu, W.Y. Toward understanding the cationicity of defensins. arg and lys versus their noncoded analogs. J. Biol. Chem. 2007, 282, 19653–19665. [Google Scholar] [CrossRef] [PubMed]
  153. Kang, J.H.; Lee, M.K.; Kim, K.L.; Hahm, K.S. Structure-biological activity relationships of 11-residue highly basic peptide segment of bovine lactoferrin. Int. J. Pept. Protein Res. 1996, 48, 357–363. [Google Scholar] [CrossRef] [PubMed]
  154. Gopal, R.; Park, S.C.; Ha, K.J.; Cho, S.J.; Kim, S.W.; Song, P.I.; Nah, J.W.; Park, Y.; Hahm, K.S. Effect of leucine and lysine substitution on the antimicrobial activity and evaluation of the mechanism of the hpa3nt3 analog peptide. J. Pept. Sci. 2009, 15, 589–594. [Google Scholar] [CrossRef] [PubMed]
  155. Mitchell, D.J.; Kim, D.T.; Steinman, L.; Fathman, C.G.; Rothbard, J.B. Polyarginine enters cells more efficiently than other polycationic homopolymers. J. Pept. Res. 2000, 56, 318–325. [Google Scholar] [CrossRef] [PubMed]
  156. Cordier, C.; Boutimah, F.; Bourdeloux, M.; Dupuy, F.; Met, E.; Alberti, P.; Loll, F.; Chassaing, G.; Burlina, F.; Saison-Behmoaras, T.E. Delivery of antisense peptide nucleic acids to cells by conjugation with small arginine-rich cell-penetrating peptide (r/w)9. PloS ONE 2014, 9, e104999. [Google Scholar] [CrossRef] [PubMed]
  157. Morita, S.; Tagai, C.; Shiraishi, T.; Miyaji, K.; Iwamuro, S. Differential mode of antimicrobial actions of arginine-rich and lysine-rich histones against gram-positive staphylococcus aureus. Peptides 2013, 48, 75–82. [Google Scholar] [CrossRef] [PubMed]
  158. Kawasaki, H.; Koyama, T.; Conlon, J.M.; Yamakura, F.; Iwamuro, S. Antimicrobial action of histone h2b in escherichia coli: Evidence for membrane translocation and DNA-binding of a histone h2b fragment after proteolytic cleavage by outer membrane proteinase t. Biochimie 2008, 90, 1693–1702. [Google Scholar] [CrossRef] [PubMed]
  159. Tagai, C.; Morita, S.; Shiraishi, T.; Miyaji, K.; Iwamuro, S. Antimicrobial properties of arginine- and lysine-rich histones and involvement of bacterial outer membrane protease t in their differential mode of actions. Peptides 2011, 32, 2003–2009. [Google Scholar] [CrossRef] [PubMed]
  160. Li, X.; Saravanan, R.; Kwak, S.K.; Leong, S.S.J. Biomolecular engineering of a human beta defensin model for increased salt resistance. Chem. Eng. Sci. 2013, 95, 128–137. [Google Scholar] [CrossRef]
  161. Scudiero, O.; Galdiero, S.; Cantisani, M.; Di Noto, R.; Vitiello, M.; Galdiero, M.; Naclerio, G.; Cassiman, J.J.; Pedone, C.; Castaldo, G.; et al. Novel synthetic, salt-resistant analogs of human beta-defensins 1 and 3 endowed with enhanced antimicrobial activity. Antimicrob. Agents Chemother. 2010, 54, 2312–2322. [Google Scholar] [CrossRef] [PubMed]
  162. Liu, Z.; Young, A.W.; Hu, P.; Rice, A.J.; Zhou, C.; Zhang, Y.; Kallenbach, N.R. Tuning the membrane selectivity of antimicrobial peptides by using multivalent design. Chembiochem 2007, 8, 2063–2065. [Google Scholar] [CrossRef] [PubMed]
  163. Nakase, I.; Okumura, S.; Katayama, S.; Hirose, H.; Pujals, S.; Yamaguchi, H.; Arakawa, S.; Shimizu, S.; Futaki, S. Transformation of an antimicrobial peptide into a plasma membrane-permeable, mitochondria-targeted peptide via the substitution of lysine with arginine. Chem. Commun. (Camb) 2012, 48, 11097–11099. [Google Scholar] [CrossRef] [PubMed]
  164. Vogel, H.J.; Schibli, D.J.; Jing, W.; Lohmeier-Vogel, E.M.; Epand, R.F.; Epand, R.M. Towards a structure-function analysis of bovine lactoferricin and related tryptophan- and arginine-containing peptides. Biochem. Cell Biol. 2002, 80, 49–63. [Google Scholar] [CrossRef] [PubMed]
  165. Khandelia, H.; Kaznessis, Y.N. Cation-pi interactions stabilize the structure of the antimicrobial peptide indolicidin near membranes: Molecular dynamics simulations. J. Phys. Chemi. B 2007, 111, 242–250. [Google Scholar] [CrossRef] [PubMed]
  166. Bi, X.; Wang, C.; Dong, W.; Zhu, W.; Shang, D. Antimicrobial properties and interaction of two trp-substituted cationic antimicrobial peptides with a lipid bilayer. J. Antibiot. 2014, 67, 361–368. [Google Scholar] [CrossRef] [PubMed]
  167. Yau, W.M.; Wimley, W.C.; Gawrisch, K.; White, S.H. The preference of tryptophan for membrane interfaces. Biochemistry 1998, 37, 14713–14718. [Google Scholar] [CrossRef] [PubMed]
  168. Shepherd, C.M.; Vogel, H.J.; Tieleman, D.P. Interactions of the designed antimicrobial peptide mb21 and truncated dermaseptin s3 with lipid bilayers: Molecular-dynamics simulations. Biochem. J. 2003, 370, 233–243. [Google Scholar] [CrossRef] [PubMed]
  169. Bi, X.; Wang, C.; Ma, L.; Sun, Y.; Shang, D. Investigation of the role of tryptophan residues in cationic antimicrobial peptides to determine the mechanism of antimicrobial action. J. Appl. Microbiol. 2013, 115, 663–672. [Google Scholar] [CrossRef] [PubMed]
  170. Tanabe, H.; Ayabe, T.; Maemoto, A.; Ishikawa, C.; Inaba, Y.; Sato, R.; Moriichi, K.; Okamoto, K.; Watari, J.; Kono, T.; et al. Denatured human alpha-defensin attenuates the bactericidal activity and the stability against enzymatic digestion. Biochem. Biophys. Res. Commun. 2007, 358, 349–355. [Google Scholar] [CrossRef] [PubMed]
  171. Miles, K.; Clarke, D.J.; Lu, W.; Sibinska, Z.; Beaumont, P.E.; Davidson, D.J.; Barr, T.A.; Campopiano, D.J.; Gray, M. Dying and necrotic neutrophils are anti-inflammatory secondary to the release of alpha-defensins. J. Immunol. 2009, 183, 2122–2132. [Google Scholar] [CrossRef] [PubMed]
  172. Varkey, J.; Nagaraj, R. Antibacterial activity of human neutrophil defensin hnp-1 analogs without cysteines. Antimicrob. Agents Chemother. 2005, 49, 4561–4566. [Google Scholar] [CrossRef] [PubMed]
  173. Kluver, E.; Schulz-Maronde, S.; Scheid, S.; Meyer, B.; Forssmann, W.G.; Adermann, K. Structure-activity relation of human beta-defensin 3: Influence of disulfide bonds and cysteine substitution on antimicrobial activity and cytotoxicity. Biochemistry 2005, 44, 9804–9816. [Google Scholar] [CrossRef] [PubMed]
  174. Zhang, L.; Scott, M.G.; Yan, H.; Mayer, L.D.; Hancock, R.E. Interaction of polyphemusin i and structural analogs with bacterial membranes, lipopolysaccharide, and lipid monolayers. Biochemistry 2000, 39, 14504–14514. [Google Scholar] [CrossRef] [PubMed]
  175. Ostberg, N.; Kaznessis, Y. Protegrin structure-activity relationships: Using homology models of synthetic sequences to determine structural characteristics important for activity. Peptides 2005, 26, 197–206. [Google Scholar] [CrossRef] [PubMed]
  176. Gottler, L.M.; de la Salud Bea, R.; Shelburne, C.E.; Ramamoorthy, A.; Marsh, E.N. Using fluorous amino acids to probe the effects of changing hydrophobicity on the physical and biological properties of the beta-hairpin antimicrobial peptide protegrin-1. Biochemistry 2008, 47, 9243–9250. [Google Scholar] [CrossRef] [PubMed]
  177. Chen, J.; Falla, T.J.; Liu, H.; Hurst, M.A.; Fujii, C.A.; Mosca, D.A.; Embree, J.R.; Loury, D.J.; Radel, P.A.; Cheng Chang, C.; et al. Development of protegrins for the treatment and prevention of oral mucositis: Structure-activity relationships of synthetic protegrin analogues. Biopolymers 2000, 55, 88–98. [Google Scholar] [CrossRef]
  178. Mangoni, M.E.; Aumelas, A.; Charnet, P.; Roumestand, C.; Chiche, L.; Despaux, E.; Grassy, G.; Calas, B.; Chavanieu, A. Change in membrane permeability induced by protegrin 1: Implication of disulphide bridges for pore formation. FEBS Lett. 1996, 383, 93–98. [Google Scholar] [CrossRef]
  179. Harwig, S.S.; Waring, A.; Yang, H.J.; Cho, Y.; Tan, L.; Lehrer, R.I. Intramolecular disulfide bonds enhance the antimicrobial and lytic activities of protegrins at physiological sodium chloride concentrations. Eur. J. Biochem. 1996, 240, 352–357. [Google Scholar] [CrossRef] [PubMed]
  180. Lazaridis, T.; He, Y.; Prieto, L. Membrane interactions and pore formation by the antimicrobial peptide protegrin. Biophys. J. 2013, 104, 633–642. [Google Scholar] [CrossRef] [PubMed]
  181. Mani, R.; Cady, S.D.; Tang, M.; Waring, A.J.; Lehrert, R.I.; Hong, M. Membrane-dependent oligomeric structure and pore formation of beta-hairpin antimicrobial peptide in lipid bilayers from solid-state nmr. Proc. Natl. Acad. Sci. USA 2006, 103, 16242–16247. [Google Scholar] [CrossRef] [PubMed]
  182. Aumelas, A.; Mangoni, M.; Roumestand, C.; Chiche, L.; Despaux, E.; Grassy, G.; Calas, B.; Chavanieu, A. Synthesis and solution structure of the antimicrobial peptide protegrin-1. Eur. J. Biochem. 1996, 237, 575–583. [Google Scholar] [CrossRef] [PubMed]
  183. Mohanram, H.; Bhattacharjya, S. Cysteine deleted protegrin-1 (cdp-1): Anti-bacterial activity, outer-membrane disruption and selectivity. Biochim. Biophys. Acta 2014, 1840, 3006–3016. [Google Scholar] [CrossRef] [PubMed]
  184. Mosca, D.A.; Hurst, M.A.; So, W.; Viajar, B.S.; Fujii, C.A.; Falla, T.J. Ib-367, a protegrin peptide with in vitro and in vivo activities against the microflora associated with oral mucositis. Antimicrob. Agents Chemother. 2000, 44, 1803–1808. [Google Scholar] [CrossRef] [PubMed]
  185. Trotti, A.; Garden, A.; Warde, P.; Symonds, P.; Langer, C.; Redman, R.; Pajak, T.F.; Fleming, T.R.; Henke, M.; Bourhis, J.; et al. A multinational, randomized phase iii trial of iseganan hcl oral solution for reducing the severity of oral mucositis in patients receiving radiotherapy for head-and-neck malignancy. Int. J. Radiat. Oncol. Biol. Phys. 2004, 58, 674–681. [Google Scholar] [CrossRef]
  186. Scocchi, M.; Tossi, A.; Gennaro, R. Proline-rich antimicrobial peptides: Converging to a non-lytic mechanism of action. Cell. Mol. Life Sci. 2011, 68, 2317–2330. [Google Scholar] [CrossRef] [PubMed]
  187. Podda, E.; Benincasa, M.; Pacor, S.; Micali, F.; Mattiuzzo, M.; Gennaro, R.; Scocchi, M. Dual mode of action of bac7, a proline-rich antibacterial peptide. Biochim. Biophys. Acta 2006, 1760, 1732–1740. [Google Scholar] [CrossRef] [PubMed]
  188. Boman, H.G.; Agerberth, B.; Boman, A. Mechanisms of action on escherichia coli of cecropin p1 and pr-39, two antibacterial peptides from pig intestine. Infect. Immun. 1993, 61, 2978–2984. [Google Scholar] [PubMed]
  189. Cabiaux, V.; Agerberth, B.; Johansson, J.; Homble, F.; Goormaghtigh, E.; Ruysschaert, J.M. Secondary structure and membrane interaction of pr-39, a pro+arg-rich antibacterial peptide. Eur. J. Biochem. 1994, 224, 1019–1027. [Google Scholar] [CrossRef] [PubMed]
  190. Taguchi, S.; Mita, K.; Ichinohe, K.; Hashimoto, S. Targeted engineering of the antibacterial peptide apidaecin, based on an in vivo monitoring assay system. Appl. Environ. Microbiol. 2009, 75, 1460–1464. [Google Scholar] [CrossRef] [PubMed]
  191. Casteels, P.; Ampe, C.; Jacobs, F.; Vaeck, M.; Tempst, P. Apidaecins: Antibacterial peptides from honeybees. EMBO J. 1989, 8, 2387–2391. [Google Scholar] [PubMed]
  192. Bulet, P.; Urge, L.; Ohresser, S.; Hetru, C.; Otvos, L., Jr. Enlarged scale chemical synthesis and range of activity of drosocin, an o-glycosylated antibacterial peptide of drosophila. Eur. J. Biochem. 1996, 238, 64–69. [Google Scholar] [CrossRef] [PubMed]
  193. Kragol, G.; Hoffmann, R.; Chattergoon, M.A.; Lovas, S.; Cudic, M.; Bulet, P.; Condie, B.A.; Rosengren, K.J.; Montaner, L.J.; Otvos, L., Jr. Identification of crucial residues for the antibacterial activity of the proline-rich peptide, pyrrhocoricin. Eur. J. Biochem. 2002, 269, 4226–4237. [Google Scholar] [CrossRef] [PubMed]
  194. Cociancich, S.; Dupont, A.; Hegy, G.; Lanot, R.; Holder, F.; Hetru, C.; Hoffmann, J.A.; Bulet, P. Novel inducible antibacterial peptides from a hemipteran insect, the sap-sucking bug pyrrhocoris apterus. Biochem. J. 1994, 300(Pt 2), 567–575. [Google Scholar] [PubMed]
  195. Otvos, L., Jr.; Bokonyi, K.; Varga, I.; Otvos, B.I.; Hoffmann, R.; Ertl, H.C.; Wade, J.D.; McManus, A.M.; Craik, D.J.; Bulet, P. Insect peptides with improved protease-resistance protect mice against bacterial infection. Protein Sci. 2000, 9, 742–749. [Google Scholar] [CrossRef] [PubMed]
  196. Otvos, L.; Snyder, C.; Condie, B.; Bulet, P.; Wade, J.D. Chimeric antimicrobial peptides exhibit multiple modes of action. Int. J. Pept. Res. Ther. 2005, 11, 29–42. [Google Scholar] [CrossRef]
  197. Casteels, P.; Tempst, P. Apidaecin-type peptide antibiotics function through a non-poreforming mechanism involving stereospecificity. Biochem. Biophys. Res. Commun. 1994, 199, 339–345. [Google Scholar] [CrossRef] [PubMed]
  198. Fernandez-Carneado, J.; Kogan, M.J.; Castel, S.; Giralt, E. Potential peptide carriers: Amphipathic proline-rich peptides derived from the n-terminal domain of gamma-zein. Angew. Chem. Int. Ed. Engl. 2004, 43, 1811–1814. [Google Scholar] [CrossRef] [PubMed]
  199. Cassone, M.; Vogiatzi, P.; La Montagna, R.; De Olivier Inacio, V.; Cudic, P.; Wade, J.D.; Otvos, L., Jr. Scope and limitations of the designer proline-rich antibacterial peptide dimer, a3-apo, alone or in synergy with conventional antibiotics. Peptides 2008, 29, 1878–1886. [Google Scholar] [CrossRef] [PubMed]
  200. Stensvag, K.; Haug, T.; Sperstad, S.V.; Rekdal, O.; Indrevoll, B.; Styrvold, O.B. Arasin 1, a proline-arginine-rich antimicrobial peptide isolated from the spider crab, hyas araneus. Dev. Comp. Immunol. 2008, 32, 275–285. [Google Scholar] [CrossRef] [PubMed]
  201. Paulsen, V.S.; Blencke, H.M.; Benincasa, M.; Haug, T.; Eksteen, J.J.; Styrvold, O.B.; Scocchi, M.; Stensvag, K. Structure-activity relationships of the antimicrobial peptide arasin 1 - and mode of action studies of the n-terminal, proline-rich region. PloS ONE 2013, 8, e53326. [Google Scholar] [CrossRef] [PubMed]
  202. Kragol, G.; Lovas, S.; Varadi, G.; Condie, B.A.; Hoffmann, R.; Otvos, L., Jr. The antibacterial peptide pyrrhocoricin inhibits the atpase actions of dnak and prevents chaperone-assisted protein folding. Biochemistry 2001, 40, 3016–3026. [Google Scholar] [CrossRef] [PubMed]
  203. Crespo, L.; Sanclimens, G.; Montaner, B.; Perez-Tomas, R.; Royo, M.; Pons, M.; Albericio, F.; Giralt, E. Peptide dendrimers based on polyproline helices. J. Am. Chem. Soc. 2002, 124, 8876–8883. [Google Scholar] [CrossRef] [PubMed]
  204. Rozgonyi, F.; Szabo, D.; Kocsis, B.; Ostorhazi, E.; Abbadessa, G.; Cassone, M.; Wade, J.D.; Otvos, L., Jr. The antibacterial effect of a proline-rich antibacterial peptide a3-apo. Curr. Med. Chem. 2009, 16, 3996–4002. [Google Scholar] [CrossRef] [PubMed]
  205. Shen, X.; Ye, G.; Cheng, X.; Yu, C.; Altosaar, I.; Hu, C. Characterization of an abaecin-like antimicrobial peptide identified from a pteromalus puparum cdna clone. J. Invertebr. Pathol. 2010, 105, 24–29. [Google Scholar] [CrossRef] [PubMed]
  206. Schneider, M.; Dorn, A. Differential infectivity of two pseudomonas species and the immune response in the milkweed bug, oncopeltus fasciatus (insecta: Hemiptera). J. Invertebr. Pathol. 2001, 78, 135–140. [Google Scholar] [CrossRef] [PubMed]
  207. Fritsche, S.; Knappe, D.; Berthold, N.; von Buttlar, H.; Hoffmann, R.; Alber, G. Absence of in vitro innate immunomodulation by insect-derived short proline-rich antimicrobial peptides points to direct antibacterial action in vivo. J. Pept. Sci. 2012, 18, 599–608. [Google Scholar] [CrossRef] [PubMed]
  208. Krizsan, A.; Volke, D.; Weinert, S.; Strater, N.; Knappe, D.; Hoffmann, R. Insect-derived proline-rich antimicrobial peptides kill bacteria by inhibiting bacterial protein translation at the 70 s ribosome. Angew. Chem. Int. Ed. Engl. 2014, 53, 12236–12239. [Google Scholar] [CrossRef] [PubMed]
  209. Bobone, S.; Roversi, D.; Giordano, L.; De Zotti, M.; Formaggio, F.; Toniolo, C.; Park, Y.; Stella, L. The lipid dependence of antimicrobial peptide activity is an unreliable experimental test for different pore models. Biochemistry 2012, 51, 10124–10126. [Google Scholar] [CrossRef] [PubMed]
  210. Roversi, D.; Luca, V.; Aureli, S.; Park, Y.; Mangoni, M.L.; Stella, L. How many antimicrobial peptide molecules kill a bacterium? The case of pmap-23. ACS Chem. Biol. 2014, 9, 2003–2007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  211. Huang, H.W. Action of antimicrobial peptides: Two-state model. Biochemistry 2000, 39, 8347–8352. [Google Scholar] [CrossRef] [PubMed]
  212. Andrushchenko, V.V.; Aarabi, M.H.; Nguyen, L.T.; Prenner, E.J.; Vogel, H.J. Thermodynamics of the interactions of tryptophan-rich cathelicidin antimicrobial peptides with model and natural membranes. Biochim. Biophys. Acta 2008, 1778, 1004–1014. [Google Scholar] [CrossRef] [PubMed]
  213. Zasloff, M. Antimicrobial peptides of multicellular organisms. Nature 2002, 415, 389–395. [Google Scholar] [CrossRef] [PubMed]
  214. Kagan, B.L.; Selsted, M.E.; Ganz, T.; Lehrer, R.I. Antimicrobial defensin peptides form voltage-dependent ion-permeable channels in planar lipid bilayer membranes. Proc. Natl. Acad. Sci. USA 1990, 87, 210–214. [Google Scholar] [CrossRef] [PubMed]
  215. Papo, N.; Shai, Y. A molecular mechanism for lipopolysaccharide protection of gram-negative bacteria from antimicrobial peptides. J. Biol. Chem. 2005, 280, 10378–10387. [Google Scholar] [CrossRef] [PubMed]
  216. Andra, J.; Gutsmann, T.; Garidel, P.; Brandenburg, K. Mechanisms of endotoxin neutralization by synthetic cationic compounds. J. Endotoxin Res. 2006, 12, 261–277. [Google Scholar] [CrossRef] [PubMed]
  217. Bobone, S.; Gerelli, Y.; De Zotti, M.; Bocchinfuso, G.; Farrotti, A.; Orioni, B.; Sebastiani, F.; Latter, E.; Penfold, J.; Senesi, R.; et al. Membrane thickness and the mechanism of action of the short peptaibol trichogin ga iv. Biochim. Biophys. Acta Biomembr. 2013, 1828, 1013–1024. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Kourie, J.I.; Shorthouse, A.A. Properties of cytotoxic peptide-formed ion channels. Am. J. Physiol. Cell. Physiol. 2000, 278, C1063–C1087. [Google Scholar] [PubMed]
  219. Tarao, K.; Shimizu, A.; Ohkawa, S.; Harada, M.; Ito, Y.; Tamai, S.; Kuni, Y.; Nagaoka, T.; Hoshino, H. Increased uptake of bromodeoxyuridine by hepatocytes from early stage of primary biliary-cirrhosis. Gastroenterology 1991, 100, 725–730. [Google Scholar] [PubMed]
  220. Steiner, H.; Andreu, D.; Merrifield, R.B. Binding and action of cecropin and cecropin analogues: Antibacterial peptides from insects. Biochim. Biophys. Acta 1988, 939, 260–266. [Google Scholar] [CrossRef]
  221. Lichtenstein, A. Mechanism of mammalian cell lysis mediated by peptide defensins. Evidence for an initial alteration of the plasma membrane. J. Clin. Investig. 1991, 88, 93–100. [Google Scholar] [CrossRef] [PubMed]
  222. Lehrer, R.I.; Barton, A.; Daher, K.A.; Harwig, S.S.; Ganz, T.; Selsted, M.E. Interaction of human defensins with escherichia coli. Mechanism of bactericidal activity. J. Clin. Investig. 1989, 84, 553–561. [Google Scholar] [CrossRef] [PubMed]
  223. Ehrenstein, G.; Lecar, H. Electrically gated ionic channels in lipid bilayers. Q. Rev. Biophys. 1977, 10, 1–34. [Google Scholar] [CrossRef] [PubMed]
  224. Brogden, K.A. Antimicrobial peptides: Pore formers or metabolic inhibitors in bacteria? Nat. Rev. Microbiol. 2005, 3, 238–250. [Google Scholar] [CrossRef] [PubMed]
  225. Wang, K.F.; Nagarajan, R.; Camesano, T.A. Antimicrobial peptide alamethicin insertion into lipid bilayer: A qcm-d exploration. Colloid. Surface. B 2014, 116, 472–481. [Google Scholar] [CrossRef] [PubMed]
  226. Laver, D.R. The barrel-stave model as applied to alamethicin and its analogs reevaluated. Biophys. J. 1994, 66, 355–359. [Google Scholar] [CrossRef]
  227. De Zotti, M.; Biondi, B.; Peggion, C.; Formaggio, F.; Park, Y.; Hahm, K.S.; Toniolo, C. Trichogin ga iv: A versatile template for the synthesis of novel peptaibiotics. Org. Biomol. Chem. 2012, 10, 1285–1299. [Google Scholar] [CrossRef] [PubMed]
  228. 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]
  229. Sengupta, D.; Leontiadou, H.; Mark, A.E.; Marrink, S.J. Toroidal pores formed by antimicrobial peptides show significant disorder. Biochim. Biophys. Acta 2008, 1778, 2308–2317. [Google Scholar] [CrossRef] [PubMed]
  230. Bechinger, B.; Lohner, K. Detergent-like actions of linear amphipathic cationic antimicrobial peptides. Biochim. Biophys. Acta Biomembr. 2006, 1758, 1529–1539. [Google Scholar] [CrossRef] [PubMed]
  231. Bechinger, B. Rationalizing the membrane interactions of cationic amphipathic antimicrobial peptides by their molecular shape. Curr. Opin. Colloid IN 2009, 14, 349–355. [Google Scholar] [CrossRef]
  232. Orioni, B.; Bocchinfuso, G.; Kim, J.Y.; Palleschi, A.; Grande, G.; Bobone, S.; Park, Y.; Kim, J.I.; Hahm, K.S.; Stella, L. Membrane perturbation by the antimicrobial peptide pmap-23: A fluorescence and molecular dynamics study. Biochim. Biophys. Acta Biomembr. 2009, 1788, 1523–1533. [Google Scholar] [CrossRef] [PubMed]
  233. Pokorny, A.; Birkbeck, T.H.; Almeida, P.F.F. Mechanism and kinetics of delta-lysin interaction with phospholipid vesicles. Biochemistry 2002, 41, 11044–11056. [Google Scholar] [CrossRef] [PubMed]
  234. Zhao, H.; Sood, R.; Jutila, A.; Bose, S.; Fimland, G.; Nissen-Meyer, J.; Kinnunen, P.K.J. Interaction of the antimicrobial peptide pheromone plantaricin a with model membranes: Implications for a novel mechanism of action. Biochim. Biophys. Acta Biomembr. 2006, 1758, 1461–1474. [Google Scholar] [CrossRef] [PubMed]
  235. Epand, R.M.; Epand, R.F. Bacterial membrane lipids in the action of antimicrobial agents. J. Pept. Sci. 2011, 17, 298–305. [Google Scholar] [CrossRef] [PubMed]
  236. Pag, U.; Oedenkoven, M.; Sass, V.; Shai, Y.; Shamova, O.; Antcheva, N.; Tossi, A.; Sahl, H.G. Analysis of in vitro activities and modes of action of synthetic antimicrobial peptides derived from an alpha-helical 'sequence template'. J. Antimicrob. Chemother. 2008, 61, 341–352. [Google Scholar] [CrossRef] [PubMed]
  237. Sass, V.; Schneider, T.; Wilmes, M.; Korner, C.; Tossi, A.; Novikova, N.; Shamova, O.; Sahl, H.G. Human beta-defensin 3 inhibits cell wall biosynthesis in staphylococci. Infect. Immun. 2010, 78, 2793–2800. [Google Scholar] [CrossRef] [PubMed]
  238. Sass, V.; Pag, U.; Tossi, A.; Bierbaum, G.; Sahl, H.G. Mode of action of human beta-defensin 3 against staphylococcus aureus and transcriptional analysis of responses to defensin challenge. Int. J. Med. Microbiol. 2008, 298, 619–633. [Google Scholar] [CrossRef] [PubMed]
  239. Soehnlein, O.; Kai-Larsen, Y.; Frithiof, R.; Sorensen, O.E.; Kenne, E.; Scharffetter-Kochanek, K.; Eriksson, E.E.; Herwald, H.; Agerberth, B.; Lindbom, L. Neutrophil primary granule proteins hbp and hnp1–3 boost bacterial phagocytosis by human and murine macrophages. J. Clin. Investig. 2008, 118, 3491–3502. [Google Scholar] [CrossRef] [PubMed]
  240. Funderburg, N.; Lederman, M.M.; Feng, Z.; Drage, M.G.; Jacllowsky, J.; Harding, C.V.; Weinberg, A.; Sieg, S.F. Human beta-defensin-3 activates professional antigen-presenting cells via toll-like receptors 1 and 2. Proc. Natl. Acad. Sci. USA 2007, 104, 18631–18635. [Google Scholar] [CrossRef] [PubMed]
  241. 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] [PubMed]
  242. 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]
  243. Li, W.F.; Ma, G.X.; Zhou, X.X. Apidaecin-type peptides: Biodiversity, structure-function relationships and mode of action. Peptides 2006, 27, 2350–2359. [Google Scholar] [CrossRef] [PubMed]
  244. Gruenheid, S.; Le Moual, H. Resistance to antimicrobial peptides in gram-negative bacteria. FEMS Microbiol. Lett. 2012, 330, 81–89. [Google Scholar] [CrossRef] [PubMed]
  245. Belas, R.; Manos, J.; Suvanasuthi, R. Proteus mirabilis zapa metalloprotease degrades a broad spectrum of substrates, including antimicrobial peptides. Infect. Immun. 2004, 72, 5159–5167. [Google Scholar] [CrossRef] [PubMed]
  246. Hritonenko, V.; Stathopoulos, C. Omptin proteins: An expanding family of outer membrane proteases in gram-negative enterobacteriaceae. Mol. Membr. Biol. 2007, 24, 395–406. [Google Scholar] [CrossRef] [PubMed]
  247. Chromek, M.; Kai-Larsen, Y.; Luthje, P.; Wang, X.; Holm, A.A.; Hedlund, K.O.; Johansson, J.; Chapman, M.R.; Jacobson, S.H.; Romling, U.; et al. The antimicrobial peptide cathelicidin interferes with polymerization of curli fimbrie and thus inhibits the formation of uropathogenic e-coli biofilm. Pediatr. Nephrol. 2010, 25, 1849–1849. [Google Scholar]
  248. Ilg, K.; Endt, K.; Misselwitz, B.; Stecher, B.; Aebi, M.; Hardt, W.D. O-antigen-negative salmonella enterica serovar typhimurium is attenuated in intestinal colonization but elicits colitis in streptomycin-treated mice. Infect. Immun. 2009, 77, 2568–2575. [Google Scholar] [CrossRef] [PubMed]
  249. Foschiatti, M.; Cescutti, P.; Tossi, A.; Rizzo, R. Inhibition of cathelicidin activity by bacterial exopolysaccharides. Mol. Microbiol. 2009, 72, 1137–1146. [Google Scholar] [CrossRef] [PubMed]
  250. Llobet, E.; Tomas, J.M.; Bengoechea, J.A. Capsule polysaccharide is a bacterial decoy for antimicrobial peptides. Microbiology. 2008, 154, 3877–3886. [Google Scholar] [CrossRef] [PubMed]
  251. Campos, M.A.; Vargas, M.A.; Regueiro, V.; Llompart, C.M.; Alberti, S.; Bengoechea, J.A. Capsule polysaccharide mediates bacterial resistance to antimicrobial peptides. Infect. Immun. 2004, 72, 7107–7114. [Google Scholar] [CrossRef] [PubMed]
  252. Gunn, J.S. The salmonella pmrab regulon: Lipopolysaccharide modifications, antimicrobial peptide resistance and more. Trends Microbiol. 2008, 16, 284–290. [Google Scholar] [CrossRef] [PubMed]
  253. Wang, X.Y.; Ribeiro, A.A.; Guan, Z.Q.; Abraham, S.N.; Raetz, C.R.H. Attenuated virulence of a francisella mutant lacking the lipid a 4'-phosphatase. Proc. Natl. Acad. Sci. USA 2007, 104, 4136–4141. [Google Scholar] [CrossRef] [PubMed]
  254. Hein-Kristensen, L.; Franzyk, H.; Holch, A.; Gram, L. Adaptive evolution of escherichia coli to an alpha-peptide/beta-peptoid peptidomimetic induces stable resistance. PloS ONE 2013, 8, e73620. [Google Scholar] [CrossRef] [PubMed]
  255. Ernst, R.K.; Guina, T.; Miller, S.I. How intracellular bacteria survive: Surface modifications that promote resistance to host innate immune responses. J. Infect. Dis. 1999, 179, S326–S330. [Google Scholar] [CrossRef] [PubMed]
  256. Guo, L.; Lim, K.B.; Gunn, J.S.; Bainbridge, B.; Darveau, R.P.; Hackett, M.; Miller, S.I. Regulation of lipid a modifications by salmonella typhimurium virulence genes phop-phoq. Science 1997, 276, 250–253. [Google Scholar] [CrossRef] [PubMed]
  257. Guo, L.; Lim, K.B.; Poduje, C.M.; Daniel, M.; Gunn, J.S.; Hackett, M.; Miller, S.I. Lipid a acylation and bacterial resistance against vertebrate antimicrobial peptides. Cell 1998, 95, 189–198. [Google Scholar] [CrossRef]
  258. Nawrocki, K.L.; Crispell, E.K.; McBride, S.M. Antimicrobial peptide resistance mechanisms of gram-positive bacteria. Antibiotics (Basel) 2014, 3, 461–492. [Google Scholar] [CrossRef] [PubMed]
  259. Bernard, E.; Rolain, T.; Courtin, P.; Guillot, A.; Langella, P.; Hols, P.; Chapot-Chartier, M.P. Characterization of o-acetylation of n-acetylglucosamine: A novel structural variation of bacterial peptidoglycan. J. Biol. Chem. 2011, 286, 23950–23958. [Google Scholar]
  260. Saar-Dover, R.; Bitler, A.; Nezer, R.; Shmuel-Galia, L.; Firon, A.; Shimoni, E.; Trieu-Cuot, P. D-Alanylation of lipoteichoic acids confers resistance to cationic peptides in group b streptococcus by increasing the cell wall density. PLoS Pathog 2012, 8, e1002891. [Google Scholar]
  261. Shafer, W.M.; Qu, X.D.; Waring, A.J.; Lehrer, R.I. Modulation of neisseria gonorrhoeae susceptibility to vertebrate antibacterial peptides due to a, member of the resistance/nodulation/division efflux pump family. Proc. Natl. Acad. Sci. USA 1998, 95, 1829–1833. [Google Scholar] [CrossRef] [PubMed]
  262. Parralopez, C.; Baer, M.T.; Groisman, E.A. Molecular-genetic analysis of a locus required for resistance to antimicrobial peptides in salmonella-typhimurium. EMBO J. 1993, 12, 4053–4062. [Google Scholar]
  263. Chakraborty, K.; Ghosh, S.; Koley, H.; Mukhopadhyay, A.K.; Ramamurthy, T.; Saha, D.R.; Mukhopadhyay, D.; Roychowdhury, S.; Hamabata, T.; Takeda, Y.; et al. Bacterial exotoxins downregulate cathelicidin (hcap-18/ll-37) and human beta-defensin 1 (hbd-1) expression in the intestinal epithelial cells. Cell. Microbiol. 2008, 10, 2520–2537. [Google Scholar] [CrossRef] [PubMed]
  264. Arai, T.; Mikami, Y.; Fukushima, K.; Utsumi, T.; Yazawa, K. A new antibiotic, leucinostatin, derived from penicillium lilacinum. J. Antibiot. 1973, 26, 157–161. [Google Scholar] [CrossRef] [PubMed]
  265. Nguyen, L.T.; de Boer, L.; Zaat, S.A.; Vogel, H.J. Investigating the cationic side chains of the antimicrobial peptide tritrpticin: Hydrogen bonding properties govern its membrane-disruptive activities. Biochim. Biophys. Acta 2011, 1808, 2297–2303. [Google Scholar] [CrossRef] [PubMed]
  266. Grauer, A.; Konig, B. Peptidomimetics - a versatile route to biologically active compounds. Eur. J. Org. Chem. 2009, 5099–5111. [Google Scholar] [CrossRef]
  267. Bessalle, R.; Kapitkovsky, A.; Gorea, A.; Shalit, I.; Fridkin, M. All-d-magainin: Chirality, antimicrobial activity and proteolytic resistance. FEBS Lett. 1990, 274, 151–155. [Google Scholar] [CrossRef]
  268. Giuliani, A.; Rinaldi, A.C. Beyond natural antimicrobial peptides: Multimeric peptides and other peptidomimetic approaches. Cell. Mol. Life Sci. 2011, 68, 2255–2266. [Google Scholar] [CrossRef] [PubMed]
  269. Shalev, D.E.; Rotem, S.; Fish, A.; Mor, A. Consequences of n-acylation on structure and membrane binding properties of dermaseptin derivative k4-s4-(1–13). J. Biol. Chem. 2006, 281, 9432–9438. [Google Scholar] [CrossRef] [PubMed]
  270. Giuliani, A.; Pirri, G.; Bozzi, A.; Di Giulio, A.; Aschi, M.; Rinaldi, A.C. Antimicrobial peptides: Natural templates for synthetic membrane-active compounds. Cell. Mol. Life Sci. 2008, 65, 2450–2460. [Google Scholar] [CrossRef] [PubMed]
  271. Bionda, N.; Pastar, I.; Davis, S.C.; Cudic, P. In vitro and in vivo activities of novel cyclic lipopeptides against staphylococcal biofilms. Protein pept. Lett. 2014, 21, 352–356. [Google Scholar] [CrossRef] [PubMed]
  272. Ahn, M.; Jacob, B.; Gunasekaran, P.; Murugan, R.N.; Ryu, E.K.; Lee, G.H.; Hyun, J.K.; Cheong, C.; Kim, N.H.; Shin, S.Y.; et al. Poly-lysine peptidomimetics having potent antimicrobial activity without hemolytic activity. Amino Acids 2014, 46, 2259–2269. [Google Scholar] [CrossRef] [PubMed]
  273. Hansen, T.; Alst, T.; Havelkova, M.; Strom, M.B. Antimicrobial activity of small beta-peptidomimetics based on the pharmacophore model of short cationic antimicrobial peptides. J. Med. Chem. 2010, 53, 595–606. [Google Scholar] [CrossRef] [PubMed]
  274. Mosca, S.; Keller, J.; Azzouz, N.; Wagner, S.; Titz, A.; Seeberger, P.H.; Brezesinski, G.; Hartmann, L. Amphiphilic cationic beta(3r3)-peptides: Membrane active peptidomimetics and their potential as antimicrobial agents. Biomacromolecules 2014, 15, 1687–1695. [Google Scholar] [CrossRef] [PubMed]
  275. Wu, C.W.; Sanborn, T.J.; Zuckermann, R.N.; Barron, A.E. Peptoid oligomers with alpha-chiral, aromatic side chains: Effects of chain length on secondary structure. J. Am. Chem. Soc. 2001, 123, 2958–2963. [Google Scholar] [CrossRef] [PubMed]
  276. Wu, C.W.; Sanborn, T.J.; Huang, K.; Zuckermann, R.N.; Barron, A.E. Peptoid oligomers with alpha-chiral, aromatic side chains: Sequence requirements for the formation of stable peptoid helices. J. Am. Chem. Soc. 2001, 123, 6778–6784. [Google Scholar] [CrossRef] [PubMed]
  277. Sanborn, T.J.; Wu, C.W.; Zuckermann, R.N.; Barron, A.E. Extreme stability of helices formed by water-soluble poly-n-substituted glycines (polypeptoids) with alpha-chiral side chains. Biopolymers 2002, 63, 12–20. [Google Scholar] [CrossRef] [PubMed]
  278. Tan, N.C.; Yu, P.; Kwon, Y.U.; Kodadek, T. High-throughput evaluation of relative cell permeability between peptoids and peptides. Bioorg. Med. Chem. 2008, 16, 5853–5861. [Google Scholar] [CrossRef] [PubMed]
  279. Bang, J.K.; Nan, Y.H.; Lee, E.K.; Shin, S.Y. A novel trp-rich model antimicrobial peptoid with increased protease stability. Bull. Korean Chem. Soc. 2010, 31, 2509–2513. [Google Scholar] [CrossRef]
  280. Chongsiriwatana, N.P.; Patch, J.A.; Czyzewski, A.M.; Dohm, M.T.; Ivankin, A.; Gidalevitz, D.; Zuckermann, R.N.; Barron, A.E. Peptoids that mimic the structure, function, and mechanism of helical antimicrobial peptides. Proc. Natl. Acad. Sci. USA 2008, 105, 2794–2799. [Google Scholar] [CrossRef] [PubMed]
  281. Patch, J.A.; Barron, A.E. Helical peptoid mimics of magainin-2 amide. J. Am. Chem. Soc. 2003, 125, 12092–12093. [Google Scholar] [CrossRef] [PubMed]
  282. Kapoor, R.; Eimerman, P.R.; Hardy, J.W.; Cirillo, J.D.; Contag, C.H.; Barron, A.E. Efficacy of antimicrobial peptoids against mycobacterium tuberculosis. Antimicrob. Agents Chemother. 2011, 55, 3058–3062. [Google Scholar] [CrossRef] [PubMed]
  283. Kapoor, R.; Wadman, M.W.; Dohm, M.T.; Czyzewski, A.M.; Spormann, A.M.; Barron, A.E. Antimicrobial peptoids are effective against pseudomonas aeruginosa biofilms. Antimicrob. Agents Chemother. 2011, 55, 3054–3057. [Google Scholar] [CrossRef] [PubMed]
  284. Chongsiriwatana, N.P.; Wetzler, M.; Barron, A.E. Functional synergy between antimicrobial peptoids and peptides against gram-negative bacteria. Antimicrob. Agents Chemother. 2011, 55, 5399–5402. [Google Scholar] [CrossRef] [PubMed]
  285. Mojsoska, B.; Zuckermann, R.N.; Jenssen, H. Structure-activity relationship study of novel peptoids that mimic the structure of antimicrobial peptides. Antimicrob. Agents Chemother. 2015, 59, 4112–4120. [Google Scholar] [CrossRef] [PubMed]
  286. Shin, S.B.; Yoo, B.; Todaro, L.J.; Kirshenbaum, K. Cyclic peptoids. J. Am. Chem. Soc. 2007, 129, 3218–3225. [Google Scholar] [CrossRef] [PubMed]
  287. Huang, M.L.; Shin, S.B.; Benson, M.A.; Torres, V.J.; Kirshenbaum, K. A comparison of linear and cyclic peptoid oligomers as potent antimicrobial agents. ChemMedChem 2012, 7, 114–122. [Google Scholar] [CrossRef] [PubMed]
  288. Huang, M.L.; Benson, M.A.; Shin, S.B.Y.; Torres, V.J.; Kirshenbaum, K. Amphiphilic cyclic peptoids that exhibit antimicrobial activity by disrupting staphylococcus aureus membranes. Eur. J. Org. Chem. 2013, 3560–3566. [Google Scholar] [CrossRef]
  289. Gobbo, M.; Benincasa, M.; Bertoloni, G.; Biondi, B.; Dosselli, R.; Papini, E.; Reddi, E.; Rocchi, R.; Tavano, R.; Gennaro, R. Substitution of the arginine/leucine residues in apidaecin ib with peptoid residues: Effect on antimicrobial activity, cellular uptake, and proteolytic degradation. J. Med. Chem. 2009, 52, 5197–5206. [Google Scholar] [CrossRef] [PubMed]
  290. Kim, J.K.; Lee, S.A.; Shin, S.; Lee, J.Y.; Jeong, K.W.; Nan, Y.H.; Park, Y.S.; Shin, S.Y.; Kim, Y. Structural flexibility and the positive charges are the key factors in bacterial cell selectivity and membrane penetration of peptoid-substituted analog of piscidin 1. Biochim. Biophys. Acta Biomembr. 2010, 1798, 1913–1925. [Google Scholar] [CrossRef] [PubMed]
  291. Liu, Y.; Knapp, K.M.; Yang, L.; Molin, S.; Franzyk, H.; Folkesson, A. High in vitro antimicrobial activity of beta-peptoid-peptide hybrid oligomers against planktonic and biofilm cultures of staphylococcus epidermidis. Int. J. Antimicrob. Agents 2013, 41, 20–27. [Google Scholar] [CrossRef] [PubMed]
  292. Olsen, C.A.; Bonke, G.; Vedel, L.; Adsersen, A.; Witt, M.; Franzyk, H.; Jaroszewski, J.W. Alpha-peptide/beta-peptoid chimeras. Org. Lett. 2007, 9, 1549–1552. [Google Scholar] [CrossRef] [PubMed]
  293. Olsen, C.A.; Ziegler, H.L.; Nielsen, H.M.; Frimodt-Moller, N.; Jaroszewski, J.W.; Franzyk, H. Antimicrobial, hemolytic, and cytotoxic activities of beta-peptoid-peptide hybrid oligomers: Improved properties compared to natural amps. Chembiochem 2010, 11, 1356–1360. [Google Scholar] [CrossRef] [PubMed]
  294. Bonke, G.; Vedel, L.; Witt, M.; Jaroszewski, J.W.; Olsen, C.A.; Franzyk, H. Dimeric building blocks for solid-phase synthesis of alpha-peptide-beta-peptoid chimeras. Synth. Stuttg. 2008, 2381–2390. [Google Scholar]
  295. Ryge, T.S.; Hansen, P.R. Novel lysine-peptoid hybrids with antibacterial properties. J. Pept. Sci. 2005, 11, 727–734. [Google Scholar] [CrossRef] [PubMed]
  296. Jahnsen, R.D.; Haney, E.F.; Franzyk, H.; Hancock, R.E. Characterization of a proteolytically stable multifunctional host defense peptidomimetic. Chem. Biol. 2013, 20, 1286–1295. [Google Scholar] [CrossRef] [PubMed]
  297. Gottschalk, S.; Ifrah, D.; Lerche, S.; Gottlieb, C.T.; Cohn, M.T.; Hiasa, H.; Hansen, P.R.; Gram, L.; Ingmer, H.; Thomsen, L.E. The antimicrobial lysine-peptoid hybrid lp5 inhibits DNA replication and induces the sos response in staphylococcus aureus. BMC microbiology 2013, 13. [Google Scholar] [CrossRef] [PubMed]
  298. Yuan, P.; Di, L.; Zhang, X.; Yan, M.; Wan, D.; Li, L.; Zhang, Y.; Cai, J.; Dai, H.; Zhu, Q.; et al. Efficacy of oral etoposide in pretreated metastatic breast cancer: A multicenter phase 2 study. Medicine 2015, 94, e774. [Google Scholar] [CrossRef] [PubMed]
  299. Li, Y.; Smith, C.; Wu, H.; Padhee, S.; Manoj, N.; Cardiello, J.; Qiao, Q.; Cao, C.; Yin, H.; Cai, J. Lipidated cyclic gamma-aapeptides display both antimicrobial and anti-inflammatory activity. ACS Chem. Biol. 2014, 9, 211–217. [Google Scholar] [CrossRef] [PubMed]
  300. Chen, X.; Cai, J.; Zhou, X.; Chen, L.; Gong, Y.; Gao, Z.; Zhang, H.; Huang, W.; Zhou, H. Protective effect of spironolactone on endothelial-to-mesenchymal transition in huvecs via notch pathway. Cell. Physiol. Biochem. 2015, 36, 191–200. [Google Scholar] [CrossRef] [PubMed]
  301. Zhang, C.; Cui, H.; Cai, J.; Duan, Y.; Liu, Y. Development of fluorescence sensing material based on cdse/zns quantum dots and molecularly imprinted polymer for the detection of carbaryl in rice and chinese cabbage. J. Agric. Food Chem. 2015, 63, 4966–4972. [Google Scholar] [CrossRef] [PubMed]
  302. Cherkasov, A.; Jankovic, B. Application of 'inductive' qsar descriptors for quantification of antibacterial activity of cationic polypeptides. Molecules 2004, 9, 1034–1052. [Google Scholar] [CrossRef] [PubMed]
  303. Taboureau, O.; Olsen, O.H.; Nielsen, J.D.; Raventos, D.; Mygind, P.H.; Kristensen, H.H. Design of novispirin antimicrobial peptides by quantitative structure-activity relationship. Chem. Biol. Drug Des. 2006, 68, 48–57. [Google Scholar] [CrossRef] [PubMed]
  304. Juretic, D.; Vukicevic, D.; Ilic, N.; Antcheva, N.; Tossi, A. Computational design of highly selective antimicrobial peptides. J. Chem. Inf. Model. 2009, 49, 2873–2882. [Google Scholar] [CrossRef] [PubMed]
  305. Juretic, D.; Vukicevic, D.; Petrov, D.; Novkovic, M.; Bojovic, V.; Lucic, B.; Ilic, N.; Tossi, A. Knowledge-based computational methods for identifying or designing novel, non-homologous antimicrobial peptides. Eur. Biophys. J. 2011, 40, 371–385. [Google Scholar] [CrossRef] [PubMed]
  306. 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] [PubMed]
  307. Cherkasov, A.; Hilpert, K.; Jenssen, H.; Fjell, C.D.; Waldbrook, M.; Mullaly, S.C.; Volkmer, R.; Hancock, R.E.W. Use of artificial intelligence in the design of small peptide antibiotics effective against a broad spectrum of highly antibiotic-resistant superbugs. ACS Chem. Biol. 2009, 4, 65–74. [Google Scholar] [CrossRef] [PubMed]
  308. Fjell, C.D.; Jenssen, H.; Hilpert, K.; Cheung, W.A.; Pante, N.; Hancock, R.E.; Cherkasov, A. Identification of novel antibacterial peptides by chemoinformatics and machine learning. J. Med. Chem. 2009, 52, 2006–2015. [Google Scholar] [CrossRef] [PubMed]
  309. Fjell, C.D.; Jenssen, H.; Cheung, W.A.; Hancock, R.E.; Cherkasov, A. Optimization of antibacterial peptides by genetic algorithms and cheminformatics. Chem. Biol. Drug Des. 2011, 77, 48–56. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Mojsoska, B.; Jenssen, H. Peptides and Peptidomimetics for Antimicrobial Drug Design. Pharmaceuticals 2015, 8, 366-415. https://doi.org/10.3390/ph8030366

AMA Style

Mojsoska B, Jenssen H. Peptides and Peptidomimetics for Antimicrobial Drug Design. Pharmaceuticals. 2015; 8(3):366-415. https://doi.org/10.3390/ph8030366

Chicago/Turabian Style

Mojsoska, Biljana, and Håvard Jenssen. 2015. "Peptides and Peptidomimetics for Antimicrobial Drug Design" Pharmaceuticals 8, no. 3: 366-415. https://doi.org/10.3390/ph8030366

Article Metrics

Back to TopTop