Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2010 Aug 1.
Published in final edited form as: Biochim Biophys Acta. 2008 Oct 29;1788(8):1680–1686. doi: 10.1016/j.bbamem.2008.10.009

Structure, Membrane Orientation, Mechanism, and Function of Pexiganan – A Highly Potent Antimicrobial Peptide Designed From Magainin

Lindsey M Gottler 1, Ayyalusamy Ramamoorthy 1,*
PMCID: PMC2726618  NIHMSID: NIHMS127432  PMID: 19010301

Abstract

The growing problem of bacterial resistance to conventional antibiotic compounds and the need for new antibiotics has stimulated interest in the development of antimicrobial peptides (AMPs) as human therapeutics. Development of topically applied agents, such as pexiganan (also known as MSI-78, an analog of the naturally occurring magainin2, extracted from the skin of the African frog Xenopus laevis) has been the focus of pharmaceutical development largely because of the relative safety of topical therapy and the uncertainty surrounding the long-term toxicology of any new class of drug administered systemically. The main hurdle that has hindered the development of antimicrobial peptides is that many of the naturally occurring peptides (such as magainin), although active in vitro, are effective in animal models of infection only at very high doses, often close to the toxic doses of the peptide, reflecting an unacceptable margin of safety. Though MSI-78 did not pass the FDA approval, it is still the best-studied AMP to date for therapeutic purposes. Biophysical studies have shown that this peptide is unstructured in solution, forms an antiparallel dimer of amphipathic helices upon binding to the membrane, and disrupts membrane via toroidal-type pore formation. This article covers functional, biophysical, biochemical and structural studies on pexiganan.

1. Introduction

There is a pressing need for the development of novel antimicrobial therapies due to the emergence of antibiotic resistant bacterial strains. In the United States, the Center for Disease Control (CDC) has reported that approximately 1.7 million cases of healthcare-associate infection resulting in 99,000 deaths are occurring annually [1]. Among these the major sites of infection are urinary tract, surgical site, lung and blood stream [1]. The long standing and continuing problem posed by antimicrobial resistant bacterial strains are exemplified by Staphylococcus aureus and Streptococcus pneumoniae.

Methicillin resistant Staphylococcus aureus (MRSA) was reported as early as 1961 with wide spread occurrences by 1991 [2, 3]. Later, reports of reduced susceptibility to vancomycin were made in 1997 [4] followed by isolation of vancomycin resistant S. aureus in 2003 [5]. Streptococcus pneumoniae serotype 19A was reported in 2007 which is resistant to all US Food and Drug Agency approved antimicrobial agents [6].

The emergence of bacterial strains resistant to most or all of the clinically useful antibiotics has provided the impetus to develop new classes of antibiotics that may combat bacterial resistance more effectively. Antimicrobial peptides (AMPs) show promise as therapeutic agents against a broad spectrum of microbes including bacteria, fungi and viruses [7-9]. Widely distributed in multicellular organisms, they form part of the initial line of defense in the innate immune system and are also implicated in the activation of the adaptive immune response against microbes [10]. The innate and adaptive immune effects of mammalian AMPs, such as defensins and cathelicidin-derived peptides, include antimicrobial activity [8, 11-13], antiviral activity [14], degranulation of mast cells [15, 16], promotion or enhancement of antigen, cytokine and chemokine response [17, 18]. Anticancer activities of AMPs have also been reported [19-21].

Although highly diverse in sequence and structure, almost all AMPs share the property of being highly amphipathic, with one face of the peptide being hydrophobic and the other face presenting a cluster of positively charged residues [22, 23]. AMPs are often classified based on the structural characteristics of the peptides. These classifications include α-helical, linear, or disulfide bonded [8]. The number of disulfide bonds range from 1-4 and result in β-hairpin-like structures [13].

Naturally occurring AMPs with α-helical structures include cecropins from insects and mammals [24], magainins from frogs [25, 26], and cathelicidins from mammals [11, 27]. Linear AMPs include indolicidin, a tryptophan rich peptide from cows and PR-39, a proline/arginine rich peptide from pigs [28, 29]. The disulfide bond containing AMPs include polyphemusin [30, 31] and tachyplesin [32, 33] from horseshoe crab, protegrin-1 from pig [34, 35], human-β-defensin-3 [13], and α-defensin (HNP3) from humans [36].

The length of AMPs ranges from ∼12-50 amino acids making them reasonably easy to synthesize. This has lead to a vast number of studies that systematically investigate the importance of amino acid composition, peptide length, net charge, and hydrophobicity on the antibacterial activity of AMPs [37, 38]. These investigations and mechanistic studies on gene-encoded AMPs have provided much insight into the mechanisms of AMPs.

Whereas some AMPs have been determined to act intracellularly [39], most appear to function primarily by disrupting bacterial cell membranes [40, 41]. Bacterial cell membranes contain predominantly negatively charged phospholipids that give rise to an electrostatic attraction to the highly cationic AMPs. On the other hand, eukaryotic membranes, which contain predominantly neutral phospholipids, are usually less susceptible to disruption by AMPs. In addition the presence of cholesterol in eukaryotic membranes increases the resistance against membrane disruption by AMPs. Upon association with the membrane, unstructured peptides become structured and begin thinning the bacterial membrane and proceed to disrupt the membrane through one of three broadly defined methods. The barrel stave method involves peptide insertion into the membrane parallel to the lipid bilayer normal, the toroidal pore method induces bending of the lipid bilayer resulting in pores in the membrane where lipids tilt in such a way that the lipid head groups define the surface of the pore, and the micellization model results in the degradation of membranes through the formation of lipid encompassed peptides [11, 42, 43]. There are other mechanisms, generally categorized as carpet mechanism, that destabilize the membrane structure to cause cell death.

There is extensive literature and general reviews on AMPs and their mechanisms. In this article, we have specifically focus on Pexiganan or MSI-78 and its very promising attributes. Within this review we describe the process by which the AMP Pexiganan was developed, the antimicrobial activity of the peptide and the high-resolution biophysical characterization of the structure and mechanism of Pexiganan. The potential therapeutic applications of Pexiganan and further developments of Pexiganan analogs are also discussed.

Pexiganan or MSI-78

Among the hundreds of gene-encoded and designed AMPs, magainin-2 and its analogs have been very well studied; amino acids sequences are given in Figure 1. Magainin-2 was co-discovered in 1987 after scientists found that Xenopus laevis were able to remain infection-free upon making incisions in the frogs' skin and placing the frog in water containing high levels of microbes [43, 44]. Two peptides were isolated from Xenopus laevis, magainin-1 and 2. The 23 amino acid magainin-2 was soon found to have broad-spectrum antibacterial and antifungal activity [45, 46]. Many synthetic analogs of magainin-2 have been developed to maximize the broad spectrum activity of the peptide in hopes of developing a clinically useful antimicrobial therapeutic agent.

Figure 1.

Figure 1

Amino acid sequences of naturally-occurring magainins (magainin-1 and 2) and MSI-78 or pexiganan, designed based on magainin 2. Helical wheel diagrams of Maigainin 2 and MSI-78 illustrate the amphipathicity of the peptide in helical form.

Zasloff et al. discovered that removal of amino acid residues from the N-terminus of magainin-2 resulted in a loss of activity [45]. Omission of residues past Lys4 was particularly detrimental to the activity of the peptide lower the MIC at least 30× for E. coli in comparison to magainin-2 [45]. Removal of amino acid residues from the C-terminus of the peptide also negatively affected the activity of the peptide significantly [45]. These results suggested that the minimal peptide length was important potentially due to the mechanism of the peptide.

The helical content of the peptide was explored as a potential target for improving the activity of magainin by Chen et al [46]. Gly to Ala substitution, which were made to increase the stability of the α-helical structure, resulted in improved activity and attempts to disrupt the helicity of the peptide by substituting D-amino acids proved to decrease the activity of the peptide [46]. These studies established the importance of the secondary structure of the peptide upon association with the cell membrane [46, 47].

Systematic single amino acid mutations to the peptide were performed by Cuervo, et al. C-terminal amidation was found to increase peptide activity as well as removal of Glu19 [48]. Poly-lysine and poly-arginine sequences appended to the termini of magainin-2 were also developed with the intent of improving the electrostatic attraction of AMP for anionic bacterial membranes [49]. The resulting peptides did not show increased activity.

With the information available from previous research Zasloff and coworkers of Magainin Pharmaceuticals did an extensive SAR study that resulted in the development of MSI-78 or Pexiganan which entered clinical trials for topical treatment of diabetic foot ulcers [50]. In 1999 the FDA denied approval of pexiganan after completion of two phase III clinical trials that revealed pexiganan was no more effective than already approved treatments for diabetic foot ulcers and required addition clinical trials for consideration [51]. Magainin Pharmaceuticals became Genaera Corporation followed by the recent acquisition of worldwide rights to pexiganan by MacroChem. Improvements in clinical trial design, greater understanding of diabetic foot ulcers and topical anti-infective treatments, and advances in peptide manufacturing keep hopes alive regarding the potential FDA approval of pexiganan [52].

Activities of Pexiganan

Extensive in vitro studies have been conducted to determine the minimum inhibitory concentration of pexiganan for a diverse array of microbes. The major categories of microbes that are of interest as potential targets for Pexiganan have included both aerobic and anaerobic, as well as Gram positive and Gram negative bacteria. Table I summarizes the MIC values obtained for Pexiganan from two sources with comparative figures for ofloxacin, the FDA approved drug used as a comparative treatment in the phase III clinical trials of Pexiganan [53, 54]. The results clearly show the effectiveness and broad spectrum activity of Pexiganan in vitro. In addition, attempts to generate resistance in bacteria by repeated treatment at subinhibitory concentration of the peptide were unsuccessful. Importantly, S. aureus, a bacteria that has quickly developed multiple resistance to current antimicrobial compounds including methicillin and vencomycin [25] was included in these attempts and showed no resistance to Pexiganan [54].

Table I.

Minimum inhibitory concentrations for aerobic and anaerobic bacterial strains and percentage of strains susceptible to inhibition by Pexiganan or Ofloxacin from literature values.

Bacteria Strain MIC Pexiganan (μg/mL) % Strains susceptible at ≤ 64 μg/mL % Strains susceptible to ofloxacin at ≤ 2 μg/mL

Reference 53 Reference 54 Reference 53 Reference 54 Reference 53
Aerobic bacteria

Acinetobacter sp. 8 8 100 100 100
Alcaligenes faecalis N/A 256 100 71 80
Citrobacter diversus 16 8 100 99 100
Citrobacter freudii 16 8 96 100 92
Corynebacterium jeikeium 4 8 100 100 100
Enterobacter aerogenes 16 32 100 100 100
Enterobacter cloacae 32 64 96 92 100
Escherichia coli 16 32 100 100 83
Klebsiella oxytoca 16 16 100 99 100
Klebsiella pneumoniae 16 16 100 100 100
Psuedomonas aeruginosa 16 16 100 99 100
Staphylococcus aureus (MRSA) 64 100 20
Staphylococcus aureus (MSSA) 16 16 100 100 93
Staphylococcus epidermidis 8 8 100 100 100
Staphylococcus haemolyticus 8 8 100 100 100
Streptococcus agalactiae 32 16 100 100 92
Streptococcus pyogenes 32 8 97 99 96

Anaerobic bacteria

Bacteroides fragilis 6 4 100 100 67
Bacteroides ovatus 6 8 100 100 0
Clostridium perfringens 18 64 100 90 100
Clostridium ramosum 16 16 100 100 0
Clostridium sporogenes 6 16 100 100 100
Peptostretococcus anaerobius 45 32 100 91 100
Peptostreptococcus magnus 1 8 100 97 0
Prevotella bivia 69 32 78 94 100
Prevotella melaninogenica 51 64 100 91 33

Kinetcis of Pexiganan cytotoxicity were studied for E. coli and S. aureus at 50 μg/mL. Colony forming units were reduced to zero by 30 minutes for E. coli and 60 minutes for S. aureus [55]. These results show the rapid onset of activity displayed by Pexiganan in vitro. The broad-spectrum activity, low propensity for generating resistance and quick onset of activity are ideal characteristics for potential therapeutics. The potential toxicity of Pexiganan has been established by measuring the hemolytic activity of the peptide against human red blood cells. Reported numbers suggest a concentration of at least 250 μg/mL are necessary to induce 100% hemolysis [55-57], much below the MIC for many of the bacterial strains listed in Table I. This exemplifies the high selectivity against erythrocytes and, in turn, low likelihood of toxicity. In addition, no adverse side effects were reported during the 2 phase III clinical trials described in detail in reference 58. The promising attributes of Pexiganan have lead to detailed studies regarding the structure and mechanism of the peptide.

Structural Studies

Solving the secondary structure of an AMP has been considered to be an important step in understanding its function and will be useful in developing potent peptides for pharmaceutical applications. While global structure analysis of AMPs using low-resolution techniques like CD and FTIR usually provides a quick information on the experimental conditions under which a conformational change occur, atomistic-level resolution three-dimensional structures can provide high-resolution information on peptide-peptide and peptide-membrane interactions. Such high-resolution structural information are powerful in understanding the role of individual amino acids in the formation of oligomers in solution or in a membrane environment and in providing insights into the mechanism of cell lysis. Below we cover the structural studies on pexiganan in solution and in model membranes.

Circular dichroism studies have shown that MSI-78 is unstructured in solution and forms an α-helix in the presence of lipids or detergents [59]; a number of biophysical studies reported similar behavior for Magainin-2 and its membrane interaction is also well established [6062]. Shanmugam et al. also showed that Pexiganan is able to adopt a β-turn structure when in methanol or dimethylsulfoxide [63]. The highly cationic nature of the peptide results in a random coil structure in aqueous solution due to electrostatic repulsion between lysine side chains. While this property is similar to most linear AMPs that are not structured in water or ionic solutions, the only human member of the cathelicidin-derived peptides, LL-37, forms a helical structure either in the presence ions or at high peptide concentration as it forms helical oligomers [64-66]. Upon association with the membrane surface, particularly anionic lipid head groups, a charge balance is achieved and Pexiganan assumes an α-helical structure [59].

The high-resolution structure, oligomerization state, and orientation of membrane associated Pexiganan has been studied by NMR spectroscopy [59, 67]. A dimeric antiparallel α -helical coiled-coil structure is formed on association with dodecylphosphocholine micelles and bilayers [67]. The interface of the dimer is a ‘phenylalanine zipper’ composed of three phenylalanine side chains per helix (Figure 2). The leucine and isoleucine residues near the termini of the helices also pack together at the dimer interface. The importance of the three phenylalanine residues in the self-association of Pexiganan is comfirmed by the fact that MSI-594, which lacks 2 of the 3 phenylalanine residues, does not oligomerize in a membrane environment [67]. Solid-state NMR studies on mechanically aligned model membranes and multilamellar vesicles of phospholipids [68], and chemical crosslinking with glutaraldehyde [69], also suggest that Pexiganan self-associates to form dimers in the presence of lipid vesicles. Interestingly, magainin-2 is also a random coil in solution and forms a dimeric antiparallel helical structure in dilauroylphosphatidylcholine vesicles at a higher concentration than that of MSI-78 (Figure 2) [70]. A recent solid-state NMR study utilized REDOR (rotational echo double resonance) [71] MAS (magic angle spinning) experiments on selectively labeled peptides to determine the backbone conformation of pexganan in phospholipids bilayers [67]. Structural analysis revealed that the structure of pexiganan is the same in detergent micelles and in lipid bilayers.

Figure 2.

Figure 2

Monomeric (A) and antiparallel dimeric (B) helical structures of MSI-78 determined by NMR experiments in a membrane environment [67]. (C) A surface representation showing the hydrophobic interface (green) and hydrophilic exterior (blue) of the dimeric helical structure. The formation of a dimer is a key step in its activity. Since the dimer has more hydrophilic surface exposed for the membrane interaction and hydrophobic residues are not exposed outside, the selectivity of the peptide towards negatively charged (both Gram positive and Gram negative) bacterial membranes is increased. Therefore, the toxicity of the peptide is further reduced.

Topology of pexiganan in membranes

In addition to the high-resolution structure, folding and topology (or membrane orientation) of an AMP is essential to fully understand the functional properties of an AMP. For example, the exact membrane orientation of an AMP can provide insights into the mechanism by which an AMP lyse bacterial cells. Solid-state NMR experiments on aligned samples have been used to obtain this information on several AMPs. Difficulties related to the preparation of fully-hydrated and stable bilayers for this purpose have been overcome by a newly developed naphthalene procedure [72]. This successful preparation of mechanically aligned glass-plate bilayer samples has been vital in the investigation of a number of AMPs [59, 64, 72, 73]. 2D PISEMA (polarization inversion spin exchange at the magic angle) [74] solid-state NMR experiments on mechanically aligned bilayers containing 15N-labeled pexiganan peptides were used to measure 15N chemical shifts and 1H-15N dipolar couplings associated with amide sites of the peptide. These NMR parameters revealed that the helical pexiganan is oriented near the surface of the membrane [59]. These results ruled out the barrel-stave type mechanism of membrane disruption (or channel formation) by pexiganan. Similar observations have also reported for magainin-2 peptide [75, 76]. On the other hand, based on 19F [77] and 2H [78] solid-state NMR experiments on PGLa embedded in lipid bilayers the authors have suggested that the helical peptide initially binds to the membrane surface and then forms dimers that are tiled by inserting into the hydrophobic region of the membrane. These studies have shown that solid-state NMR experiments on fluid and physiologically-relevant lipid bilayers provide insights into the function of antimicrobial peptides that are difficult to obtain by other means. It should also be mentioned here that care must be taken in interpreting solid-state NMR data obtained from samples that are not physiologically relevant.

Mechanistic Studies

The mechanism of membrane interaction and disruption has been established via fluorescence assays, calorimetric techniques, microscopy, solid-state NMR spectroscopy and neutron diffraction. The consensus is that Pexiganan exerts its antibacterial effect by forming toroidal pores in the bacterial membrane (Figure 3).

Figure 3.

Figure 3

Mechanism of membrane disruption by MSI-78. (A) NMR studies have shown that MSI-78 is unstructured in solution and forms a helix in a membrane environment [59, 67]. The amphipathic peptide is aligned near the surface of the membrane [59]. (B) Positive curvature strain induced by the peptide is determined from 31P solid-state NMR on POPE bilayers and differential scanning calorimetry experiments on DiPoPE bilayers [81]. (C) Formation of toroidal pores was determined from solid-state NMR studies [81]. (D) Solid-state NMR studies revealed the formation of normal hexagonal phase structure of lipids at higher concentrations (>10 mole %) of MSI-78 [59, 81]. (E) Solid-state NMR experiments revealed that a couple of weeks old samples exhibited the formation of bicelles and then micellization due to the detergent-like behavior of the peptide [results unpublished].

Cell membrane disruption by Pexiganan was confirmed by monitoring the uptake or leakage of fluorescent molecules from either E. coli and lipid vesicles, respectively [59]. Pexiganan effectively induced the uptake of ANS into E. coli cell membranes as well as inducing the leakage of carboxyfluorescene from POPC/POPG (3:1) vesicles, a model system for bacterial cell membranes. In both studies, the membranes were maximally affected within 5 minutes of exposure to Pexiganan, indicating the rapid onset of membrane disruption by the peptide [59].

Isothermal titration calorimetry reveals that the association of Pexiganan to lipid vesicles is exothermic with a binding enthalpy of -14.4 kcal/mol [79] and is in agreement with measurements on Magainin-2 which has a ΔH = -17.0 kcal/mol [80]. The binding enthalpies of cationic peptides have been attributed primarily to the electrostatic interactions between the peptide and the lipid head groups.

The effect of the peptide on lipid bilayers has been very informative of the mechanistic route of Pexiganan [81]. Differential scanning calorimetry (DSC) and NMR spectroscopic studies have pinpointed the formation of toroidal pore formation by Pexiganan in lipid bilayers [81]. DSC experiments in DiPoPE show a concentration dependant increase in the fluid lamellar to inverted hexagonal phase transition of the bilayer which supports the induction of positive curvature strain on the lipid bilayer upon incorporation of Pexiganan [81]. 31P NMR also showed that Pexiganan inhibited the fluid lamellar to inverted hexagonal phase transition and 15N NMR supporting the formation of toroidal pores in the lipid bilayer due to the orientation of the peptide perpendicular to the bilayer normal [59, 81]. Solid-state NMR results from mechanically aligned bilayers were used to determine the toroidal pore geometry in the presence of pexiganan. Solid-state NMR studies on bicelles containing MSI-78 showed the peptide-induced disorders in the hydrophobic region of the lipid bilayer and also the detergent behavior of the peptide [82]. The significant reduction in the 14N quadrupole coupling observed from the choline head group of the lipid in bilayers containing pexiganan [66] and molecular dynamic simulations [83] revealed the electrostatic interactions between the positively charged residues of the peptide and the phosphate O- atom of the lipid head group.

Membrane thinning effects of Pexiganan have been visually observed using atomic force microscopy (AFM). Mecke et al present images of supported lipid bilayers composed of DMPC in the presence and absence of Pexiganan [84]. There is an obvious thinning of the bilayer as a function of time in the presence of the peptide.

Can Pexiganan be made more potent?

Although promising as broad-spectrum antibiotics, MSI-78 and other AMPs are susceptible to proteolysis in vivo by endogenous or bacterial proteases, which may considerably diminish their effectiveness for intravenous applications. Studies on Leishmania pinpoint leishmanolysin as the preventative factor in AMP induced apoptosis of the bacteria [85]. Attempts to overcome this problem by increasing the dose of the AMP often leads to toxic side effects, most notably lysis of red blood cells, which has been attributed to non-specific hydrophobic interactions between the peptide and the eukaryotic cell membrane [86]. Improvements to the stability and/or activity of Pexiganan would be significant for the potential application of Pexiganan, or derivatives thereof, as antimicrobial therapeutic agents.

Derivatives of Pexiganan-related peptides designed to further optimize the stability and activity of the peptide for potential applications in vivo include: acylated analogs [50, 87, 88] and non-natural amino acid analogs [35, 89]. Incorporation of the 12 carbon lauryl group or amino lauryl group to the N terminus of MSI-78 resulted in increased hemolytic activity and no substantial gains in antibacterial activity [87]. Non-natural amino acid analogs, specifically fluorinated amino acides, including hexafluoroleucine and pentafluorophenylalanine [69] have been substituted into MSI-78 for the hydrophobic residues leucine, isoleucine and phenylalanine [35, 89]. These substitutions allow for retained helical structure in lipids, micelles, or trifluoroethanol [35, 89]. In addition, antimicrobial activity is generally retained or improved in hexafluoroleucine variants with no additional hemolytic activity. Also, the proteolytic stability of the fluorous peptides is greatly enhanced giving rise to potentially improved bioavailability of the peptides [35, 89]. β-peptides [90], peptoids composed of poly-N-substituted glycines [91], and cationic oligourea polymers [92] have also been designed to mimic the structure and amphipathic nature of Pexiganan, but have shown to be not susceptible to proteolytic degradation.

Synergistic capability of Pexiganan has been evaluated against bacteria present in the bloodsteams of neutopenic febrile patients including P. aeruginosa, E. coli, S. aureus (methicillin resistant and methicillin susceptible) and S. epidermidis (methicillin resistant and methicillin susceptible) [93]. MSI-78 appears to act in synergy with β -lactams which has been attributed to increased ability of the β-lactams to penetrate the bacterial membrane due to thinning of the membrane in the presence of the AMP [93]. Sepsis rat models have also been used to determine the effectiveness of Pexiganan alone and in combination with β -lactams [94]. The synergistic activity of the Pexiganan/β-lactam combination is observed for the treatment of endotoxic shock. The LPS binding ability of Pexiganan is implicated as a very important aspect of the synergistic effect of these antibiotics due to the association of endotoxin release with β-lactam activity [94].

Potential Applications

The antimicrobial activity of Pexiganan against a broad spectrum of bacterial species makes it a promising candidate for the treatment of bacterial infections. In addition to topical application for the treatment of diabetic foot ulcers as was studied in clinical trials, the above described studies imply the potential applicability of Pexiganan for the treatment of sepsis, particularly in combination with approved antibacterial agents such as β -lactams. The development of non-natural amino acid containing or non-peptidic mimics of Pexiganan could be key to improving biological stability and bioavailability while retaining the broad-spectrum activity and low toxicity characteristics of Pexiganan.

Acknowledgments

We thank Ravi Nanga for molecular dynamics simulation studies on MSI-78 and help with figures of this manuscript. This research was supported by the NIH grant (AI054515 to A.R.) and a grant in aid from the American Heart Association to A.R.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Center for Disease Control, http://www.cdc.gov/ncidod/dhqp/ar_vre.html.
  • 2.Jevons MP. Celbenin-resistant staphylococci. British Med J. 1961;1:124–125. [Google Scholar]
  • 3.Johnson AP, Aucken HM, Cavendish S, Ganner M, Wale MC, Warner M, Livermore DM, Cookson BD. Dominance of EMRSA-15 and -16 among MRSA causing nosocomial bacteraemia in the UK: analysis of isolates from the European Antimicrobial Resistance Surveillance System (EARSS) J Antimicrob Chemother. 2001;48:143–144. doi: 10.1093/jac/48.1.143. [DOI] [PubMed] [Google Scholar]
  • 4.Hiramatsu K, Hanaki H, Ino T, Yabuta K, Oguri T, Tenover FC. Methicillin-resistant Staphylococcus aureus clinical strain with reduced vancomycin susceptibility. J Antimicrob Chemother. 1990;40:135–136. doi: 10.1093/jac/40.1.135. [DOI] [PubMed] [Google Scholar]
  • 5.Chang S, Sievert DM, Hageman JC, Boulton ML, Tenover FC, Downes FP, Shah S, Rudrik JT, Pupp GR, Brown WJ, Cardo D, Fridkin SK. Infection with vancomycin-resistant Staphylococcus aureus containing the vanA resistance gene. N Engl J Med. 2003;348:1342–1347. doi: 10.1056/NEJMoa025025. [DOI] [PubMed] [Google Scholar]
  • 6.Pichichero M, Casey JR. Emergence of a Multiresistant Serotype 19A Pneumococcal Strain Not Included in the 7-Valent Conjugate Vaccine as an Otopathogen in Children. J Am Med Assoc. 2007;298:1772–1778. doi: 10.1001/jama.298.15.1772. [DOI] [PubMed] [Google Scholar]
  • 7.Shai Y. From Innate Immunity to de-Novo Designed Antimicrobial Peptides. Curr Pharm Design. 2002;8:715–725. doi: 10.2174/1381612023395367. [DOI] [PubMed] [Google Scholar]
  • 8.Zasloff M. Antimicrobial peptides of multicellular organisms. Nature. 2002;415:389–395. doi: 10.1038/415389a. [DOI] [PubMed] [Google Scholar]
  • 9.Boman HG. Peptide Antibiotics and their Role in Innate Immunity. Ann Rev Immunol. 1995;13:61–92. doi: 10.1146/annurev.iy.13.040195.000425. [DOI] [PubMed] [Google Scholar]; Hancock REW, Lehrer R. Cationic peptides: a new source of antibiotics. Trends in Biotech. 1998;16:82–88. doi: 10.1016/s0167-7799(97)01156-6. [DOI] [PubMed] [Google Scholar]; Hancock REW, Diamond G. The role of cationic antimicrobial peptides in innate host defences. Trends Microbiol. 2000;8:402–10. doi: 10.1016/s0966-842x(00)01823-0. [DOI] [PubMed] [Google Scholar]; Epand RF, Ramamoorthy A, Epand RM. Membrane lipid composition and the interaction with Pardaxin: the role of cholesterol. Protein Peptide Letters. 2006;13:1–5. [PubMed] [Google Scholar]; Epand RM, Vogel HJ. Diversity of antimicrobial peptides and their mechanisms of action. BBA Biomembranes. 1999;1462:11–28. doi: 10.1016/s0005-2736(99)00198-4. [DOI] [PubMed] [Google Scholar]
  • 10.Oppenheim JJ, Biragyn A, Kwak LW, Yang D. Roles of antimicrobial peptides such as defensins in innate and adaptive immunity. Ann Rheum Dis. 2003;62:17–21. doi: 10.1136/ard.62.suppl_2.ii17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Dürr UH, Sudheendra US, Ramamoorthy A. LL-37, the only human member of the cathelicidin family of antimicrobial peptides. BBA Biomembranes. 2006;1758:1408–1425. doi: 10.1016/j.bbamem.2006.03.030. [DOI] [PubMed] [Google Scholar]
  • 12.Chan DI, Prenner EJ, Vogel HJ. Tryptophan- and arginine-rich antimicrobial peptides: Structures and mechanisms of action. BBA Biomembranes. 2006;1758:1184–1202. doi: 10.1016/j.bbamem.2006.04.006. [DOI] [PubMed] [Google Scholar]
  • 13.Dhople V, Krukemeyer A, Ramamoorthy A. The human beta-defensin-3, an antibacterial peptide with multiple biological functions. BBA Biomembranes. 2006;1758:1408–1425. doi: 10.1016/j.bbamem.2006.07.007. [DOI] [PubMed] [Google Scholar]
  • 14.Rugeles MT, Trubey CM, Bedoya VI, Pinto LA, Oppenheim JJ, Rybak SM. Ribonuclease is partly responsible for the HIV-1 inhibitory effect activated by HLA alloantigen recognition. AIDS. 2003;17:481–486. doi: 10.1097/00002030-200303070-00002. [DOI] [PubMed] [Google Scholar]
  • 15.Niyonsaba F, Someya A, Hirata M, Ogawa H, Nagaoka I. Roles of antimicrobial peptides such as defensins in innate and adaptive immunity. Eur J Immunol. 2001;31:1066–1075. doi: 10.1002/1521-4141(200104)31:4<1066::aid-immu1066>3.0.co;2-#. [DOI] [PubMed] [Google Scholar]
  • 16.Huang HJ, Ross CR, Blecha F. Chemoattractant properties of PR-39, a neutrophil antibacterial peptide. J Leukoc Biol. 1997;61:624–629. doi: 10.1002/jlb.61.5.624. [DOI] [PubMed] [Google Scholar]
  • 17.Lillard JW, Jr, Boyaka PN, Chertov O, Oppenheim JJ, McGhee JR. The Role of Antimicrobial Peptides in Innate Immunity. Proc Natl Acad Sci. 1999;96:651–656. doi: 10.1073/pnas.96.2.651. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Tani K, Murphy WJ, Chertov O, Salcedo R, Koh CY, Utsunomiya I. Defensins act as potent adjuvants that promote cellular and humoral immune responses in mice to a lymphoma idiotype and carrier antigens. Int Immunol. 2000;12:691–700. doi: 10.1093/intimm/12.5.691. [DOI] [PubMed] [Google Scholar]
  • 19.Baker MA, Maloy WL, Zasloff M, Jacob LS. Anticancer efficacy of magainin 2 and analogue peptides. Cancer Res. 1993;53:3052–3057. [PubMed] [Google Scholar]; Papo N, Shai Y. Host defense peptides as new weapons in cancer treatment. Cell Mol Life Sci. 2005;62:784–790. doi: 10.1007/s00018-005-4560-2. [DOI] [PubMed] [Google Scholar]
  • 20.Papo N, Shai Y. New lytic peptides based on the D,L-amphipathic helix motif preferentially kill tumor cells compared to normal cells. Biochemistry. 2003;42:9346–9354. doi: 10.1021/bi027212o. [DOI] [PubMed] [Google Scholar]; Papo N, Seger D, Makovitzki A, Kalchenko V, Eshhar Z, Degani H, Shai Y. Inhibition of tumor growth and elimination of multiple metastases in human prostate and breast xenografts by systemic inoculation of a host defense-like lytic peptide. Cancer Res. 2006;66:5371–5378. doi: 10.1158/0008-5472.CAN-05-4569. [DOI] [PubMed] [Google Scholar]
  • 21.Hoskin DW, Ramamoorthy A. Studies on Anticancer Activities of Antimicrobial Peptides. BBA Biomembranes. 2008;1778:357–375. doi: 10.1016/j.bbamem.2007.11.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Shai Y. Mechanism of the binding, insertion and destabilization of phospholipid bilayer membranes by alpha-helical antimicrobial and cell non-selective membrane-lytic peptides. Biochim Biophys Acta. 1999;1462:55–70. doi: 10.1016/s0005-2736(99)00200-x. [DOI] [PubMed] [Google Scholar]
  • 23.Wu MH, Maier E, Benz R, Hancock REW. 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. doi: 10.1021/bi9826299. [DOI] [PubMed] [Google Scholar]
  • 24.Lee JY, Boman A, Sun C, Andersson M, Jornvall H, Mutt V, Boman HG. Antibacterial peptides from pig intestine: isolation of a mammalian cecropin. Proc Natl Acad Sci. 1989;86:9159–9162. doi: 10.1073/pnas.86.23.9159. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Bevins CL, Zasloff M. Peptides from Frog Skin. Ann Rev Biochem. 1990;59:395–414. doi: 10.1146/annurev.bi.59.070190.002143. [DOI] [PubMed] [Google Scholar]
  • 26.Barra D, Simmaco M. Amphibian skin: a promising resource for antimicrobial peptides. Tren Biotech. 1998;13:205–209. doi: 10.1016/S0167-7799(00)88947-7. [DOI] [PubMed] [Google Scholar]
  • 27.Bals R, Wang X, Zasloff M, Wilson J. The peptide antibiotic LL-37/hCAP-18 is expressed in epithelia of the human lung where it has broad antimicrobial activity at the airway surface. Proc Natl Acad Sci. 1998;95:9541–9546. doi: 10.1073/pnas.95.16.9541. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Selsted ME, Novotny MJ, Morris WL, Tang YQ, Smith W, Cullor JS. Indolicidin, a novel bactericidal tridecapeptide amide from neutrophils. J Biol Chem. 1992;267:4292–4295. [PubMed] [Google Scholar]
  • 29.Agerberth B, Lee JY, Bergman T, Carlquist M, Boman HG, Mutt V, Jornvall H. Amino acid sequence of PR-39-Isolation from pig intestine of a new member of the family of proline-arginine-rich antibacterial peptides. Eur J Biochem. 1991;202:849–854. doi: 10.1111/j.1432-1033.1991.tb16442.x. [DOI] [PubMed] [Google Scholar]
  • 30.Powers JS, Martin MM, Goosney DL, Hancock RE. The Antimicrobial Peptide Polyphemusin Localizes to the Cytoplasm of Escherichia coli following Treatment. Antimicrob Agents Chemother. 2006;50:1522–1524. doi: 10.1128/AAC.50.4.1522-1524.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Powers JP, Tan A, Ramamoorthy A, Hancock RE. Solution structure and interaction of the antimicrobial polyphemusins with lipid membranes. Biochemistry. 2005;44:15504–15513. doi: 10.1021/bi051302m. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Nakamura T, Furunaka H, Miyata T, Tokunaga F, Muta T, Iwanaga S, Niwa M, Takao T, Shimonishi Y. Tachyplesin, a class of antimicrobial peptide from the hemocytes of the horseshoe crab (Tachypleus tridentatus). Isolation and chemical structure. J Biol Chem. 1988;263:16709–16713. [PubMed] [Google Scholar]
  • 33.Imura Y, Nishida M, Ogawa Y, Takakura Y, Matsuzaki K. Action mechanism of tachyplesin I and effects of PEGylation. BBA Biomembranes. 2007;1768:1160–1169. doi: 10.1016/j.bbamem.2007.01.005. [DOI] [PubMed] [Google Scholar]; Ramamoorthy A, Thennarasu S, Tan A, Gottipati K, Sreekumar S, Heyl DL, An FY, Shelburne CE. Deletion of All Cysteines in Tachyplesin I Abolishes Hemolytic Activity and Retains Antimicrobial Activity and Lipopolysaccharide Selective Binding. Biochemistry. 2006;45:6529–6540. doi: 10.1021/bi052629q. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Harwig SS, Waring A, Yang HJ, Cho Y, Tan L, Lehrer RI. Intramolecular disulfide bonds enhance the antimicrobial and lytic activities of protegrins at physiological sodium chloride concentrations. Eur J Biochem. 1996;240:352–357. doi: 10.1111/j.1432-1033.1996.0352h.x. [DOI] [PubMed] [Google Scholar]
  • 35.Gottler LM, de la Salud Bea R, Shelburne CE, Ramamoorthy A, Marsh EN. 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. doi: 10.1021/bi801045n. [DOI] [PubMed] [Google Scholar]
  • 36.Ganz T, Selsted ME, Szklarek D, Harwig SS, Daher K, Bainton DF, Lehrer RI. Defensins. Natural peptide antibiotics of human neutrophils. J Clin Invest. 1985;76:1427–1435. doi: 10.1172/JCI112120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Tossi A, Sandri L, Giangaspero A. Amphipathic, α-helical antimicrobial peptides. Biopolymers. 2000;44:4–30. doi: 10.1002/1097-0282(2000)55:1<4::AID-BIP30>3.0.CO;2-M. [DOI] [PubMed] [Google Scholar]
  • 38.Dennison SR, Wallace J, Harris F, Phoenix DA. Amphiphilic α -Helical Antimicrobial Peptides and Their Structure/Function Relationships. Protein Pept Lett. 2005;12:31–39. doi: 10.2174/0929866053406084. [DOI] [PubMed] [Google Scholar]
  • 39.Brogen KA. Antimicrobial peptides: pore formers or metabolic inhibitors in bacteria? Nat Rev Microbiol. 2005;3:238–250. doi: 10.1038/nrmicro1098. [DOI] [PubMed] [Google Scholar]
  • 40.Oren Z, Shai Mode of action of linear amphipathic α-helical antimicrobial peptides. Biopolymers. 1998;47:451–463. doi: 10.1002/(SICI)1097-0282(1998)47:6<451::AID-BIP4>3.0.CO;2-F. [DOI] [PubMed] [Google Scholar]
  • 41.Huang HW, Chen FY, Lee MT. Molecular mechanism of peptide induced pores in. membranes. Phys Rev Lett. 2004;92:198304, 1–4. doi: 10.1103/PhysRevLett.92.198304. [DOI] [PubMed] [Google Scholar]
  • 42.Westerhoff HV, Juretic D, Hendler RW, Zasloff M. Magainins and the disruption of membrane-linked free-energy transduction. Proc Natl Acad Sci. 1989;86:6597–6601. doi: 10.1073/pnas.86.17.6597. [DOI] [PMC free article] [PubMed] [Google Scholar]; Matsuzaki K. Magainins as paradigm for the mode of action of pore forming polypeptides. Biochim Biophys Acta. 1998;1376:391–400. doi: 10.1016/s0304-4157(98)00014-8. [DOI] [PubMed] [Google Scholar]
  • 43.Zasloff M. Magainins, a class of antimicrobial peptides from Xenopus skin: isolation, characterization of two active forms, and partial DNA sequence of a precursor. Proc Natl Acad Sci USA. 1987;84:5449–5453. doi: 10.1073/pnas.84.15.5449. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Giovannini MG, Poulter L, Gibson BW, Williams DH. Biosynthesis and degradation of peptides derived from Xenopus laevis prohormones. Biochem J. 1987;243:113–120. doi: 10.1042/bj2430113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Zasloff M, Martin B, Chen HC. Antimicrobial activity of synthetic magainin peptides and several analogues. Proc Natl Acad Sci. 1988;85:910–913. doi: 10.1073/pnas.85.3.910. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Chen HC, Brown JH, Morell JL, Huang CM. Synthetic magainin analogues with improved antimicrobial activity. FEBS Lett. 1988;236:462–466. doi: 10.1016/0014-5793(88)80077-2. [DOI] [PubMed] [Google Scholar]
  • 47.Matsuzaki K, Sugishita K, Harada M, Fujii N, Miyajima K. Interactions of an antimicrobial peptide, magainin 2, with outer and inner membranes of Gram-negative bacteria. Biochim Biophys Acta. 1997;1327:119–130. doi: 10.1016/s0005-2736(97)00051-5. [DOI] [PubMed] [Google Scholar]
  • 48.Cuervo JH, Rodriguez B, Houghten RA. The Magainins: sequence factors relevant to increased antimicrobial activity and decreased hemolytic activity. Pept Res. 1988;1:81–86. [PubMed] [Google Scholar]
  • 49.Besalle R, Haas H, Goria A, Shalit I, Fridkin M. Augmentation of the antibacterial activity of magainin by positive-charge chain extension. Antimicrob Agents Chemother. 1992;36:313–317. doi: 10.1128/aac.36.2.313. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Maloy WL, Kari UP. Structure-Activity Studies on Magainins and Other Host Defense Peptides. Biopolymers. 1995;37:105–122. doi: 10.1002/bip.360370206. [DOI] [PubMed] [Google Scholar]
  • 51.Moore Andrew. The Big and Small of Drug Discovery. EMBO Rep. 2003;4:114–117. doi: 10.1038/sj.embor.embor748. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Press Release, MacroChem, October 3, 2007 (http://www.residentandstaff.com/press_release.asp?id=10168311).
  • 53.Fuchs PC, Barry AL, Brown SD. In Vitro Antimicrobial Acitivity of MSI-78, a Magainin Analog. Antimicrob Agents Chemother. 1998;42:1213–1216. doi: 10.1128/aac.42.5.1213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Ge Y, MacDonald DL, Holroyd KH, Thornsberry C, Wexler H, Zasloff M. In Vitro Antibacterial Properties of Pexiganan, an Analog of Magainin. Antimicrob Agents Chemother. 1999;43:782–788. doi: 10.1128/aac.43.4.782. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Navon-Venezia S, Feder R, Gaidukov L, Carmeli Y, Mor A. Antibacterial Properties of Dermaseptin S4 Derivatives with In Vivo Activity. Antimicrob Agents Chemother. 2002;46:689–694. doi: 10.1128/AAC.46.3.689-694.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Radzishevsky IS, Rotem S, Bourdetsky D, Navon-Venezia S, Carmeli Y, Mor A. Improved antimicrobial peptides based on acyl-lysine oligomers. Nat Biotechnol. 2007;25:657–659. doi: 10.1038/nbt1309. [DOI] [PubMed] [Google Scholar]
  • 57.Eren T, Som A, Rennie JR, Nelson CF, Urgina Y, Nüsslein K, Coughlin EB, Tew GN. Antibacterial and Hemolytic Activities of Quaternary Pyridinium Functionalized Polynorbornenes. Macromol Chem and Phy. 2008;209:516–524. [Google Scholar]
  • 58.Lamb HM, Wiseman LR. Pexiganan Acetate. Drugs. 1998;56:1047–1052. doi: 10.2165/00003495-199856060-00011. [DOI] [PubMed] [Google Scholar]
  • 59.Ramamoorthy A, Thennarasu S, Lee DK, Tan A, Maloy L. Solid-state NMR Investigations of the Membrane-Disrupting Mechanism of Antimicrobial Peptides MSI-78 and MSI-594. Biophysical J. 2006;91:206–216. doi: 10.1529/biophysj.105.073890. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Matzuzaki K, Harada M, Handa T, Fumakoshi S, Fujii N, Yajima H, Miyajima K. Magainin 1 induced leakage of entrapped calcein out of negatively-charged lipid vesicles. Biochim Biophys Acta. 1989;981:130–134. doi: 10.1016/0005-2736(89)90090-4. [DOI] [PubMed] [Google Scholar]; Matsuzaki K, Sugishita K, Ishibe N, Ueha M, Nakata S, Miyajima K, Epand RM. Relationship of membrane curvature to the formation of pores by magainin 2. Biochemistry. 1998;37:11856–11863. doi: 10.1021/bi980539y. [DOI] [PubMed] [Google Scholar]; Matsuzaki K, Sugishita K, Fujii N, Miyajima K. Molecular basis for membrane selectivity of an antimicrobial peptide, magainin 2. Biochemistry. 1995;34:3423–3429. doi: 10.1021/bi00010a034. [DOI] [PubMed] [Google Scholar]; Matsuzaki K, Murase O, Fujii N, Miyajima K. An antimicrobial peptide, magainin 2, induced rapid flip-flop of phospholipids coupled with pore formation and peptide translocation. Biochemistry. 1996;35:11361–11368. doi: 10.1021/bi960016v. [DOI] [PubMed] [Google Scholar]
  • 61.Williams RW, Starman R, Taylor KM, Cable K, Beeler T, Zasloff M, Covell D. Raman spectroscopy of synthetic antimicrobial frog peptides magainin 2a and PGLa. Biochemistry. 1990;29:4490–4496. doi: 10.1021/bi00470a031. [DOI] [PubMed] [Google Scholar]
  • 62.Ludtke SJ, He K, Wu Y, Huang HW. Cooperative membrane insertion of magainin correlated woth its cytolytic activity. Biochim Biophys Acta. 1994;1190:181–184. doi: 10.1016/0005-2736(94)90050-7. [DOI] [PubMed] [Google Scholar]
  • 63.Shanmugam G, Polavarapu PL, Gopinath D, Jayakumar R. The Structure of Antimicrobial Pexiganan Peptide in Solution Probed by Fourier Transform Infrared Absorption, Vibrational Circular Dichroism, and Electronic Circular Dichroism Spectroscopy. Biopolymers. 2005;80:636–642. doi: 10.1002/bip.20132. [DOI] [PubMed] [Google Scholar]
  • 64.Henzler Wildman KA, Lee DK, Ramamoorthy A. Mechanism of lipid bilayer disruption by the human antimicrobial peptide, LL-37. Biochemistry. 2003;42:6558. doi: 10.1021/bi0273563. [DOI] [PubMed] [Google Scholar]; Henzler Wildman KA, Martinez GV, Brown MF, Ramamoorthy A. Perturbation of the hydrophobic core of lipid bilayers by the human antimicrobial peptide LL-37. Biochemistry. 2004;43:8459–8469. doi: 10.1021/bi036284s. [DOI] [PubMed] [Google Scholar]
  • 65.Porcelli F, Verardi R, Shi L, Henzler-Wildman KA, Ramamoorthy A, Veglia G. NMR Structure of the Cathelicidin-Derived Human Antimicrobial Peptide LL-37 in Dodecylphosphocholine Micelles. Biochemistry. 2008;47:5565–5572. doi: 10.1021/bi702036s. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Ramamoorthy A, Lee DK, Santos JS, Henzler-Wildman KA. Nitrogen-14 Solid-State NMR Spectroscopy of Aligned Phospholipid Bilayers to Probe Peptide-Lipid Interaction and Oligomerization of Membrane Associated Peptides. J Am Chem Soc. 2008;130:11023–11029. doi: 10.1021/ja802210u. [DOI] [PubMed] [Google Scholar]
  • 67.Porcelli F, Buck-Koehntop B, Thennarasu S, Ramamoorthy A, Veglia G. Structures of the dimeric and monomeric variants of magainin antimicrobial peptides (MSI-78 and MSI-594) in micelles and bilayers by NMR spectroscopy. Biochemistry. 2006;45:5793–5799. doi: 10.1021/bi0601813. [DOI] [PubMed] [Google Scholar]
  • 68.Hallock KJ. Ph D thesis. Department of Chemistry, University of Michigan; Ann Arbor: 2002. Investigation of Membrane Disruption by Peptides Using Solid-State NMR Techniques. [Google Scholar]
  • 69.Gottler LM. Ph D thesis. Department of Chemistry, University of Michigan; Ann Arbor: 2008. Peptides as Model Systems, Antimicrobial Agents, and a Means for Protein Superassembly. [Google Scholar]
  • 70.Wakamatsu K, Takeda A, Tachi T, Matsuzaki K. Dimer structure of magainin 2 bound to phospholipid vesicles. Biopolymers. 2002;64:314–327. doi: 10.1002/bip.10198. [DOI] [PubMed] [Google Scholar]
  • 71.Gullion T, Schaefer J. Rotational-Echo Double-Resonance NMR. J Magn Reson. 1989;81:196–200. doi: 10.1016/j.jmr.2011.09.003. [DOI] [PubMed] [Google Scholar]
  • 72.Hallock KJ, Henzler Wildman KA, Lee DK, Ramamoorthy A. Sublimable solids can be used to mechanically align lipid bilayers for solid-state NMR studies. Biophys J. 2002;82:2499. doi: 10.1016/S0006-3495(02)75592-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Thennarasu S, Lee DK, Poon A, Kawulka KE, Vederas JC, Ramamoorthy A. Membrane permeabilization, orientation, and antimicrobial mechanism of subtilosin A. Chem Phys Lipids. 2005;137:38–51. doi: 10.1016/j.chemphyslip.2005.06.003. [DOI] [PubMed] [Google Scholar]; Hallock KJ, Lee Dong-Kuk, Omnaas John, Mosberg HI, Ramamoorthy A. Membrane composition determines pardaxin's mechanism of lipid bilayer disruption. Biophys J. 2002;83:1004–1013. doi: 10.1016/S0006-3495(02)75226-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Ramamoorthy A, Wei Y, Lee DK. PISEMA Solid-State NMR Spectroscopy. Ann Rep NMR Spectrosc. 2004;52:1–52. [Google Scholar]
  • 75.Bechinger B, Zasloff M, Opella SJ. Structure and orientation of the antibiotic peptide magainin in membranes by solid-state Nuclear Magnetic Resonance Spectroscopy. Prot Sci. 1993;2:2077–2084. doi: 10.1002/pro.5560021208. [DOI] [PMC free article] [PubMed] [Google Scholar]; Bechinger B, Lohner K. Detergent-like actions of linear amphipathic cationic antimicrobial peptides. BBA Biomembranes. 2006;1758:1529–1539. doi: 10.1016/j.bbamem.2006.07.001. [DOI] [PubMed] [Google Scholar]
  • 76.Lee DK, Santos JS, Ramamoorthy A. Application of one-dimensional dipolar-shift solid-state NMR spectroscopy to study the backbone conformation of membrane-associated peptides in phospholipid bilayers. J Phys Chem B. 1999;103:8383. [Google Scholar]
  • 77.Tremouilhac P, Strandberg E, Wadhwani P, Ulrich AS. Synergistic transmembrane alignment of the antimicrobial heterodimer PGLa/magainin. J Biol Chem. 2006;281:32089–32094. doi: 10.1074/jbc.M604759200. [DOI] [PubMed] [Google Scholar]; Glaser RW, Sachse C, Dürr UHN, Wadhwani P, Afonin S, Strandberg E, Ulrich AS. Concentration-dependent realignment of the antimicrobial peptide PGLa in lipid membranes observed by solid-state 19F-NMR. Biophys J. 2005;88:3392–3397. doi: 10.1529/biophysj.104.056424. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Tremouilhac P, Strandberg E, Wadhwani P, Ulrich AS. Conditions affecting the re-alignment of the antimicrobial peptide PGLa in membranes as monitored by solid-state 2H NMR. BBA Biomembranes. 2006;1758:1330–1342. doi: 10.1016/j.bbamem.2006.02.029. [DOI] [PubMed] [Google Scholar]
  • 79.Gottler LM, Lee HY, Shelburne CE, Ramamoorthy A, Marsh ENG. Using Fluorous Amino Acids to Modulate the Biological Activity of an Antimicrobial Peptide. ChemBioChem. 2008;9:370–373. doi: 10.1002/cbic.200700643. [DOI] [PubMed] [Google Scholar]
  • 80.Wenk MR, Seelig J. Magainin 2 Amide Interaction with Lipid Membranes: Calorimetric Detection of Peptide Binding and Pore Formation. Biochemistry. 1998;37:3909–3916. doi: 10.1021/bi972615n. [DOI] [PubMed] [Google Scholar]
  • 81.Hallock KJ, Lee DK, Ramamoorthy A. MSI-78, an analogue of the magainin antimicrobial peptides, disrupts lipid bilayer structure via positive curvature strain. Biophys J. 2003;84:3052–3060. doi: 10.1016/S0006-3495(03)70031-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Dvinskikh SV, Dürr UHN, Yamamoto K, Ramamoorthy A. A high-resolution solid-state NMR approach for the structural studies of bicelles. J Am Chem Soc. 2006;128:6326–7. doi: 10.1021/ja061153a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Kandasamy SK, Larson RG. Binding and insertion of alpha-helical antimicrobial peptides in POPC bilayers studied by molecular dynamics simulations. Chem Phys Lipids. 2004;132:113–132. doi: 10.1016/j.chemphyslip.2004.09.011. [DOI] [PubMed] [Google Scholar]
  • 84.Mecke A, Lee DK, Ramamoorthy A, Orr BG, Holl MMB. Membrane thinning due to antimicrobial peptide binding: An atomic force microscopy study of MSI-78 in lipid bilayers. Biophys J. 2005;99:4043–4050. doi: 10.1529/biophysj.105.062596. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Kulkarni MM, McMaster WR, Kamysz E, Kamysz W, Engman DM, McGwire BS. The surface-metalloprotease of the parasitic protozoan, Leishmania, protects against antimicrobial peptide- induced apoptotic killing. Mol Biol. 2006;62:1484–1497. doi: 10.1111/j.1365-2958.2006.05459.x. [DOI] [PubMed] [Google Scholar]
  • 86.Zelezetsky I, Tossi A. Alpha-helical antimicrobial peptides—Using a sequence template to guide structure–activity relationship studies. BBA Biomembranes. 2006;1758:1436–1449. doi: 10.1016/j.bbamem.2006.03.021. [DOI] [PubMed] [Google Scholar]
  • 87.Radzishevsky IS, Rotem S, Zaknoon F, Gaidukov L, Dagan A, Mor A. Effects of acyl versus aminoacyl conjugation on the properties of antimicrobial peptides. Antimicrob Agents Chemother. 2005;49:2412–2420. doi: 10.1128/AAC.49.6.2412-2420.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Thennarasu S, Lee DK, Tan A, Kari UP, Ramamoorthy A. Antimicrobial Actvity and Membrane Selective Interactions of a Synthetic Lipopeptide MSI-843. Biochim Biophys Acta. 2005;1711:49–58. doi: 10.1016/j.bbamem.2005.02.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Meng H, Kumar K. Antimicrobial activity and protease stability of peptides containing fluorinated amino acids. J Am Chem Soc. 2007;129:15615–15622. doi: 10.1021/ja075373f. [DOI] [PubMed] [Google Scholar]
  • 90.Porter EA, Weisblum B, Gellman SH. Mimicry of host-defense peptides by unnatural oligomers: antimicrobial β -peptides. J Am Chem Soc. 2002;124:7324–7330. doi: 10.1021/ja0260871. [DOI] [PubMed] [Google Scholar]
  • 91.Chongsiriwatana NP, Patch JA, Czyzewski AM, Dohm MT, Ivankin A, Gidalevitz D, Zuckerman RN, Barron AE. Peptoids that mimic the structure, function, and mechanism of helical antimicrobial peptides. Proc Natl Acad Sci. 2008;105:2794–2799. doi: 10.1073/pnas.0708254105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Tang H, Doerksen RJ, Tew GN. Synthesis of Urea Oligomers and Their Antibacterial Activity. Chem Comm. 2005;12:1537–1539. doi: 10.1039/b413679a. [DOI] [PubMed] [Google Scholar]
  • 93.Giacometti A, Cirioni O, Kamysz W, D'Amato G, Silvestri C, Licci A, Nadolski P, Riva A, Lukasiak J, Scalise G. In vitro activity of MSI-78 alone and in combination with antibiotics against bacteria responsible for bloodstream infections in neutropenic patients. Int J Antimicrobial Agents. 2005;26:235–240. doi: 10.1016/j.ijantimicag.2005.06.011. [DOI] [PubMed] [Google Scholar]
  • 94.Giacometti A, Cirioni O, Ghiselli R, Orlando F, Kamysz W, Rocchi M, D'Amato G, Mocchegiani F, Silvestri C, Lukasiak J, Saba V, Scalise G. Effects of pexiganan alone and combined with betalactams in experimental endotoxic shock. Peptides. 2005;26:207–216. doi: 10.1016/j.peptides.2004.09.012. [DOI] [PubMed] [Google Scholar]

RESOURCES