Jump to content

Antimicrobial peptides

From Wikipedia, the free encyclopedia

This is an old revision of this page, as edited by Sdlitvin (talk | contribs) at 20:29, 17 November 2011 (Added new study on Hemophilus). The present address (URL) is a permanent link to this revision, which may differ significantly from the current revision.

Antimicrobial peptides (also called host defense peptides) are an evolutionarily conserved component of the innate immune response and are found among all classes of life. Fundamental differences exist between prokaryotic and eukaryotic cells that may represent targets for antimicrobial peptides. These peptides are potent, broad spectrum antibiotics which demonstrate potential as novel therapeutic agents. Antimicrobial peptides have been demonstrated to kill Gram negative and Gram positive bacteria (including strains that are resistant to conventional antibiotics), mycobacteria (including Mycobacterium tuberculosis), enveloped viruses, fungi and even transformed or cancerous cells [1] Unlike the majority of conventional antibiotics it appears as though antimicrobial peptides may also have the ability to enhance immunity by functioning as immunomodulators.

Structure

Various structures of antimicrobial peptides

Antimicrobial peptides are a unique and diverse group of molecules, which are divided into subgroups on the basis of their amino acid composition and structure.[2] Antimicrobial peptides are generally between 12 and 50 amino acids. These peptides include two or more positively charged residues provided by arginine, lysine or, in acidic environments, histidine, and a large proportion (generally >50%) of hydrophobic residues.[3][4][5] The secondary structures of these molecules follow 4 themes, including i) α-helical, ii) β-stranded due to the presence of 2 or more disulfide bonds, iii) β-hairpin or loop due to the presence of a single disulfide bond and/or cyclization of the peptide chain, and iv) extended.[6] Many of these peptides are unstructured in free solution, and fold into their final configuration upon partitioning into biological membranes. It contains hydrophilic amino acid residues aligned along one side and hydrophobic amino acid residues aligned along the opposite side of a helical molecule.[2] This amphipathicity of the antimicrobial peptides allows to partition into the membrane lipid bilayer. The ability to associate with membranes is a definitive feature of antimicrobial peptides[7] although membrane permeabilisation is not necessary. These peptides have a variety of antimicrobial activities ranging from membrane permeabilization to action on a range of cytoplasmic targets.

Type characteristic AMPs
Anionic peptides rich in glutamic and aspartic acids Maximin H5 from amphibians, Dermcidin from humans
Linear cationic α-helical peptides lack in cysteine Cecropins, andropin, moricin, ceratotoxin and melittin from insects, Magainin, dermaseptin, bombinin, brevinin-1,esculentins and buforin II from amphibians, CAP18 from rabbits, LL37 from humans
Catioinic peptide enriched for specific amino acid rich in proline, arginine, phenylalanine, glycine, tryptophan abaecin, apidaecins from honeybees, prophenin from pigs, indolicidin from cattle.
Anionic and cationic peptides that contain cysteine and form disulfide bonds contain 1~3 disulfide bond 1 bond:brevinins, 2 bonds:protegrin from pig, tachyplesins from horseshoe crabs, 3 bonds:defensins from humans, more than 3:drosomycin in fruit flies
The modes of action by Antimicrobial peptides

Antimicrobial Activities

The modes of action by which antimicrobial peptides kill bacteria is varied and includes disrupting membranes, interfering with metabolism, and targeting cytoplasmic components. The initial contact between the peptide and the target organism is electrostatic, as most bacterial surface are anionic, or hydrophobic, such as in the antimicrobial peptide Piscidin. Their amino acid composition, amphipathicity, cationic charge and size allow them to attach to and insert into membrane bilayers to form pores by ‘barrel-stave’, ‘carpet’ or ‘toroidal-pore’ mechanisms. Or allow them to penetrate into the cell to bind intracellular molecules which are crucial to cell living.[8] Intracellular binding model includes inhibition of cell wall synthesis, alteration of cytoplasmic membrane, activation of autolysin, inhibition of DNA, RNA, and protein synthesis, and inhibition of certain enzymes. However, in many cases, the exact mechanism of killing is not known. One emerging technique for the study of such mechanisms is dual polarisation interferometry,.[9][10] In contrast to many conventional antibiotics these peptides appear to be bacteriocidal (bacteria killer) instead of bacteriostatic (bacteria growth inhibitor). In general the antimicrobial activity of these peptides is determined by measuring the minimal inhibitory concentration (MIC), which is the lowest concentration of drug that inhibits bacterial growth.[11]

Immunomodulatory Activities

In addition to killing bacteria directly they have been demonstrated to have a number of immunomodulatory functions that may be involved in the clearance of infection, including the ability to alter host gene expression, act as chemokines and/or induce chemokine production, inhibiting lipopolysaccharide induced pro-inflammatory cytokine production, promoting wound healing, and modulating the responses of dendritic cells and cells of the adaptive immune response. Animal models indicate that host defence peptides are crucial for both prevention and clearance of infection. It appears as though many peptides initially isolated as and termed “antimicrobial peptides” have been shown to have more significant alternative functions in vivo (e.g. hepcidin[12]).

Determination of AMP action

Several methods have been used to determine the mechanisms of antimicrobial peptide activity.[8]

Methods Applications
Microscopy to visualize the effects of antimicrobial peptides on microbial cells.
Fluorescent dyes to measure antimicrobial peptides to permeabilize membrane vesicles.
Ion channel formation to assess the formation and stability of an antimicrobial-peptide-induced pore.
Circular dichroism and orientated circular dichroism to measure the orientation and secondary structure of an antimicrobial peptide bound to a lipid bilayer
Dual Polarization Interferometry to measure the different mechanisms of antimocrobial peptides
Solid-state NMR spectroscopy to measure the secondary structure, orientation and penetration of antimicrobial peptides into lipid bilayers in the biologically relevant LIQUID-CRYSTALLINE STATE
Neutron diffraction to measure the diffraction patterns of peptide-induced pores within membranes in oriented multilayers or liquids

Therapeutic Potential

These peptides are excellent candidates for development as novel therapeutic agents and complements to conventional antibiotic therapy because in contrast to conventional antibiotics they do not appear to induce antibiotic resistance while they generally have a broad range of activity, are bacteriocidal as opposed to bacteriostatic and require a short contact time to induce killing. A number of naturally occurring peptides and their derivatives have been developed as novel anti-infective therapies for conditions as diverse as oral mucositis, lung infections associated with cystic fibrosis (CF), cancer,[13] and topical skin infections. Pexiganan has been shown to be useful to treat infection related diabetic foot ulcer.

Models of Antimicrobial Peptides

Computer simulations have recently[when?] provided atomistic resolution pictures of how antimicrobial peptides interact with membranes.[14][15] Biophysical studies utilizing solid-state NMR experiments have provided an atomic-level resolution explanation of membrane disruption by antimicrobial peptides.[16][17]

Selectivity of antimicrobial peptides

In the competition of bacterial cells and host cells with the antimicrobial peptides, antimicrobial peptides will preferentially interact with the bacterial cell to the mammalian cells, which enables them to kill microorganisms without being significantly toxic to mammalian cells.[18] Selectivity is a very important feature of the antimicrobial peptides and it can guarantee their function as antibiotics in host defense systems.

Factors that determine the selectivity of antimicrobial peptides

There are some factors that are closely related to the selectivity property of antimicrobial peptides, among which the cationic property contributes most. Since the surface of the bacterial membranes is more negatively charged than mammalian cells, antimicrobial peptides will show different affinities towards the bacterial membranes and mammalian cell membranes.[19]

In addition, there are also other factors that will affect the selectivity. It’s well known that cholesterol is normally widely distributed in the mammalian cell membranes as a membrane stabilizing agents but absent in bacterial cell membranes; and the presence of these cholesterols will also generally reduce the activities of the antimicrobial peptides, due either to stabilization of the lipid bilayer or to interactions between cholesterol and the peptide. So the cholesterol in mammalian cells will protect the cells from attack by the antimicrobial peptides.[20]

Besides, the transmembrane potential is well-known to affect peptide-lipid interactions.[21] There's an inside-negative transmembrane potential existing from the outer leaflet to the inner leaflet of the cell membranes and this inside-negative transmembrane potential will facilitate membrane permeabilization probably by facilitating the insertion of positively charged peptides into membranes. By comparison, the transmembrane potential of bacterial cells is more negative than that of normal mammalian cells, so bacterial membrane will be prone to be attacked by the positively charged antimicrobial peptides.

Similarly, it is also believed that increasing ionic strength,[20] which in general reduces the activity of most antimicrobial peptides, contributes partially to the selectivity of the antimicrobial peptides by weakening the electrostatic interactions required for the initial interaction.

Molecular Basis of Cell Selectivity of Antimicrobial Peptides

Mechanism of the selectivity

The cell membranes of bacteria are rich in acidic phospholipids, such as phosphatidylglycerol and cardiolipin.[18][22] These phospholipid headgroups are heavily negatively charged. Therefore, the outmost leaflets of the bilayer which is exposed to the outside of the bacterial membranes are more attractive to the attack of the positively charged antimicrobial peptides. So the interaction between the positive charges of antimicrobial peptides and the negatively charged bacterial membranes is mainly the electrostatic interactions, which is the major driving force for cellular association. Besides, since antimicrobial peptides form structures with a positively charged face as well as a hydrophobic face, there are also some hydrophobic interactions between the hydrophobic regions of the antimicrobial peptides and the zwitterionic phospholipids (electrically neutral) surface of the bacterial membranes, which act only as a minor effect in this case.

In contrast, the outer part of the membranes of the plants and mammals is mainly composed of lipid without any net charges since most of the lipids with negatively charged headgroups are principally sequestered into the inner leaflet of the plasma membranes.[19] Thus in the case of mammals cells, the outer surfaces of the membranes are usually made of zwitterionic phosphatidylcholine and sphingomyelin, even though a small portion of the membranes outer surfaces contain some negatively charged gangliosides. So the hydrophobic interaction between the hydrophobic face of amphipathic antimicrobial peptides and the zwitterionic phospholipids on the cell surface of mammalian cell membranes plays a major role in the formation of peptide-cell binding.[23] However, the hydrophobic interaction is relatively weak when compared to the electrostatic interaction, thus, the antimicrobial peptides will preferentially interact with the bacterial membranes.

Dual polarisation interferometry has been used in vitro to study and quantify the association to headgroup, insertion into the bilayer, pore formation and eventual disruption of the membrane.[24][25]

Methods to control the selectivity of antimicrobial peptides

A lot of efforts have been tried to control the cell selectivities. For example, Katsumi tried to modify and optimize the physicochemical parameters of the peptides to control the selectivities, including net charge, helicity, hydrophobicity per residue (H), hydrophobic moment (μ) and the angle subtended by the positively charged polar helix face (Φ).[21] Besides, other methods like the introduction of D-amino acids and fluorinated amino acids in the hydrophobic face is believed to break the secondary structures and thus to reduce hydrophobic interaction that’s necessary for interaction with mammalian cells. Wan L Z, el[citation needed] also found that Pro→Nlys substitution in Pro-containing β-turn antimicrobial peptides is a promising strategy for the design of new short bacterial cell-selective antimicrobial peptides with intracellular mechanisms of action. Nadezhda V el[citation needed] suggested that direct attachment of magainin to the substrate surface decreased nonspecific cell binding as well as led to improved detection limit for bacterial cells such as Salmonella and E. coli.

Bacterial Resistance

Bacteria use various resistance strategies to avoid antimicrobial peptide killing.[8] Some microorganisms alter net surface charges. Staphylococcus aureus transports D-alanine from the cytoplasm to the surface teichoic acid to reduce the net negative charge by introducing basic amino groups.[26] S. aureus also modifies its anionic membranes via MprF with L-lysine, increasing the positive net charge.[26] The interaction of antimicrobial peptides with membrane targets can be limited by capsule polysaccharide of Klebsiella pneumoniae.[27] Alterations occur in Lipid A. Salmonella species reduce the fluidity of their outer membrane by increasing hydrophobic interactions between an increased number of Lipid A acyl tails by adding myristate to Lipid A with 2-hydroxymyristate and forming hepta-acylated Lipid A by adding palmitate. The increased hydrophobic moment is though to retard or abolish antimicrobial peptide insertion and pore formation. The residues undergo alteration in membrane proteins. In some Gram-negative bacteria, alteration in the production of outer membrane proteins correlates with resistance to killing by antimicrobial peptides.[28] Nontypeable Hemophilus influenzae transports AMPs into the interior of the cell, where they are degraded. And H. influenzae remodels its membranes to make it appear as if the bacterium has already been successfully attacked by AMPs, protecting it from being attacked by more AMPs. [29] ATP-binding cassette transporters import antimicrobial peptides and the resistance-nodulation cell-division efflux pump exports antimicrobial peptides.[30] Both transporters have been associated with antimicrobial peptide resistance. Bacteria produce proteolytic enzymes,which may degrade antimicrobial peptides leading to their resistance.[31]

Examples

Examples of antimicrobal peptides include magainins, alamethicin, hydramacin-1, pexiganan or MSI-78, and other MSI peptides like MSI-843 and MSI-594, polyphemusin, human antimicrobial peptide, LL-37, defensins and protegrins.

See also

References

  1. ^ Template:Moore
  2. ^ a b Yeaman & Yount 2003
  3. ^ Papagianni 2003
  4. ^ Sitaram & Nagaraj 2002
  5. ^ Dürr, Sudheendra & Ramamoorthy 2006
  6. ^ Dhople, Krukemeyer & Ramamoorthy 2006
  7. ^ Hancock & Rozek 2002
  8. ^ a b c Brogden 2005 harvnb error: multiple targets (2×): CITEREFBrogden2005 (help)
  9. ^ Daniel J Hirst, Tzong-Hsien Lee, Marcus J Swann, Sharon Unabia, Yoonkyung Park, Kyung-Soo Hahm and Marie Isabel Aguilar, The effect of acyl chain structure and bilayer phase state on the binding and insertion of HPA3 onto a supported lipid bilayer, European Biophysics Journal
  10. ^ Tzong-Hsien Lee, Christine Heng, Marcus J. Swann, John D. Gehman, Frances Separovic, Marie-Isabel Aguilar, Real time quantitative analysis of lipid disordering by aurein 1.2 during membrane adsorption, destabilisation and lysis, . Biochimica et Biophysica Acta (BBA) - Biomembranes, Volume 1798, Issue 10, October 2010, Pages 1977-1986.
  11. ^ Amsterdam 1996
  12. ^ Hunter et al. 2002
  13. ^ Hoskin & Ramamoorthy 2008
  14. ^ Mátyus, Kandt & Tieleman 2007
  15. ^ Langham, Ahmad & Kaznessis 2008
  16. ^ Hallock, Lee & Ramamoorthy 2003
  17. ^ Wildman, Lee & Ramamoorthy 2003
  18. ^ a b Matsuzaki 2008
  19. ^ a b Hancock & Sahl 2006
  20. ^ a b Zasloff 2002
  21. ^ a b Matsuzaki et al. 1995
  22. ^ Chou et al. 2008
  23. ^ Tennessen 2005
  24. ^ Lanlan Yu, Lin Guo, Jeak Ling Ding, Bow Ho, Si-Shen Feng, Jonathan Popplewell, Marcus Swann, Thorsten Wohland. Interaction of an artificial antimicrobial peptide with lipid membranes. Biochimica et Biophysica Acta (BBA) - Biomembranes, Volume 1788, Issue 2, February 2009, Pages 333-344, Available online 25 October 2008
  25. ^ Tzong-Hsien Lee, Kristopher Hall, Adam Mechler, Lisandra Martin, Jonathan Popplewell, Gerry Ronan, Marie-Isabel Aguilar Molecular Imaging and Orientational Changes of Antimicrobial Peptides in Membranes American Peptide Society (2007) Peptides for Youth. Eds. Emanuel Escher, William D. Lubell, Susan Del Valle
  26. ^ a b Peschel et al. 1999
  27. ^ Campos et al. 2004
  28. ^ China et al. 1994
  29. ^ Catherine L. Shelton, Forrest K. Raffel, Wandy L. Beatty, Sara M. Johnson, Kevin M. Mason. Sap Transporter Mediated Import and Subsequent Degradation of Antimicrobial Peptides in Haemophilus. PLoS Pathogens, 3 November 2011, doi:10.1371/journal.ppat.1002360, available online
  30. ^ Nikaido 1996
  31. ^ Whitelock et al. 1996
  • Amsterdam, D. (1996), "Susceptibility testing of Antimicrobials in liquid media", in Lorian, V. (ed.), Antibiotics in Laboratory Medicine (Fourth ed.), Baltimore, Md.: Williams and Wilkins, pp. 52–111
  • Brogden, K.A. (2005), "Antimicrobial peptides: pore formers or metabolic inhibitors in bacteria?", Nature Reviews Microbiology, 3 (3): 238–250., doi:10.1038/nrmicro1098, PMID 15703760 {{citation}}: Unknown parameter |month= ignored (help)
  • Campos, M.A.; Vargas, M.A.; Regueiro, V.; Llompart, C.M.; Alberti, S.; Bengoechea, J.A. (2004), Weiser, J.N. (ed.), "Capsule Polysaccharide Mediates Bacterial Resistance to Antimicrobial Peptides", Infection and Immunity, 72 (12): 7107–7114, doi:10.1128/IAI.72.12.7107-7114.2004, PMC 529140, PMID 15557634 {{citation}}: |access-date= requires |url= (help); Unknown parameter |month= ignored (help)
  • China, B.; N'guyen, B.T.; De Bruyere, M.; Cornelis, G.R. (1994), "Role of YadA in resistance of Yersinia enterocolitica to phagocytosis by human polymorphonuclear leukocytes", Infection and immunity, 62 (4): 1275–1281, PMC 186269, PMID 8132334 {{citation}}: |access-date= requires |url= (help); Unknown parameter |month= ignored (help)
  • Chou, Hung-Ta; Kuo, Tsun-Yung; Chiang, Jung-Chun; Pei, Min-Ju; Yang, Wei-Ter; Yu, Hui-Chun; Lin, Shih-Bin; Chen, Wei-Jung (2008), "Design and synthesis of cationic antimicrobial peptides with improved activity and selectivity against Vibrio spp", International Journal of Antimicrobial Agents, 32 (2): 130–138, doi:10.1016/j.ijantimicag.2008.04.003, PMID 18586467 {{citation}}: Unknown parameter |month= ignored (help)
  • Dhople, V.; Krukemeyer, A.; Ramamoorthy, A. (2006), "The human beta-defensin-3, an antibacterial peptide with multiple biological functions", Biochimica et Biophysica Acta – Biomembranes, 1758 (9): 1499–1512, doi:10.1016/j.bbamem.2006.07.007, PMID 16978580 {{citation}}: Unknown parameter |month= ignored (help)
  • Dürr, U.H.N.; Sudheendra, U.S.; Ramamoorthy, A. (2006), "LL-37, the only human member of the cathelicidin family of antimicrobial peptides", Biochimica et Biophysica Acta – Biomembranes, 1758 (9): 1408–1425, doi:10.1016/j.bbamem.2006.03.030, PMID 16716248 {{citation}}: Unknown parameter |month= ignored (help)
  • Giuliani, A.; Pirri, G.; Nicoletto, S. (2007), "Antimicrobial peptides: an overview of a promising class of therapeutics", Cent. Eur. J. Biol., 2 (1): 1–33, doi:10.2478/s11535-007-0010-5 {{citation}}: Unknown parameter |month= ignored (help)
  • Hallock, K.J.; Lee, D.K.; Ramamoorthy, A. (2003), "MSI-78, an analogue of the magainin antimicrobial peptides, disrupts lipid bilayer structure via positive curvature strain", Biophys. J., 84 (5): 3052–3060, doi:10.1016/S0006-3495(03)70031-9, PMC 1302867, PMID 12719236 {{citation}}: Unknown parameter |month= ignored (help)
  • Hancock, Robert E.W.; Rozek, A. (2002), "Role of membranes in the activities of antimicrobial cationic peptides.", FEMS Microbiology Letters, 206 (2): 143–149, doi:10.1111/j.1574-6968.2002.tb11000.x, PMID 11814654
  • Hancock, Robert E.W.; Sahl, Hans-Georg (2006), "Antimicrobial and host-defense peptides as new anti-infective therapeutic strategies", Nature Biotechnology, 24 (12): 1551–1557, doi:10.1038/nbt1267, PMID 17160061
  • Hoskin, D.W.; Ramamoorthy, A. (2008), "Studies on anticancer activities of antimicrobial peptides", Biochimica et Biophysica Acta – Biomembranes, 1778 (2): 357–375, doi:10.1016/j.bbamem.2007.11.008, PMC 2238813, PMID 18078805 {{citation}}: Unknown parameter |month= ignored (help)
  • Hunter, H.N.; Fulton, D.B.; Ganz, T.; Vogel, H.J. (2002), "The solution structure of human hepcidin, a peptide hormone with antimicrobial activity that is involved in iron uptake and hereditary hemochromatosis.", J Biol Chem, 277 (40): 37597–37603, doi:10.1074/jbc.M205305200, PMID 12138110 {{citation}}: Unknown parameter |month= ignored (help)CS1 maint: unflagged free DOI (link)
  • Langham, Allison A.; Ahmad, Abdallah Sayyed; Kaznessis, Yiannis N. (2008), "On the Nature of Antimicrobial Activity: A Model for Protegrin-1 Pores", J. Am. Chem. Soc., 130 (13): 4338–4346, doi:10.1021/ja0780380, PMC 2474802, PMID 18335931 {{citation}}: Unknown parameter |month= ignored (help)
  • Matsuzaki, Katsumi (2008), "Control of cell selectivity of antimicrobial peptides", Biochimica et Biophysica Acta – Biomembranes, 1788 (8): 1687–92, doi:10.1016/j.bbamem.2008.09.013, PMID 18952049
  • Matsuzaki, Katsumi; Ken-ichi, S.; Nobutaka, F.; Koichiro, M. (1995), "Molecular Basis for Membrane Selectivity of an Antimicrobial Peptides, Magainin 2", Biochemistry, 34 (10): 3423–3429, doi:10.1021/bi00010a034, PMID 7533538 {{citation}}: Unknown parameter |month= ignored (help)
  • Mátyus, Edit; Kandt, Christian; Tieleman, D Peter (2007), "Computer simulation of antimicrobial peptides", Curr Med Chem, 14 (26): 2789–98, doi:10.2174/092986707782360105, PMID 18045125
  • Moore, A.J.; Devine,, D.A.; Bibby,, M.C. (1994), "Preliminary experimental anticancer activity of cecropins", Peptide Research, 7 (5): 265–269, PMID 7849420{{citation}}: CS1 maint: extra punctuation (link)
  • Nikaido, H. (1996), "Multidrug efflux pumps of gram-negative bacteria", Journal of Bacteriology, 178 (20): 5853–5859, PMC 178438, PMID 8830678 {{citation}}: |access-date= requires |url= (help); Unknown parameter |month= ignored (help)
  • Papagianni, M. (2003), "Ribosomally synthesized peptides with antimicrobial properties: biosynthesis, structure, function, and applications", Biotechnol Adv, 21 (6): 465–499, doi:10.1016/S0734-9750(03)00077-6, PMID 14499150 {{citation}}: Unknown parameter |month= ignored (help)
  • Peschel, A.; Otto, M.; Jack, R.W.; Kalbacher, H.; Jung, G.; Götz, F. (1999), "Inactivation of the dlt operon in Staphylococcus aureus confers sensitivity to defensins, protegrins, and other antimicrobial peptides", Journal of Biological Chemistry, 274 (13): 8405–8410, doi:10.1074/jbc.274.13.8405, PMID 10085071 {{citation}}: |access-date= requires |url= (help); Unknown parameter |month= ignored (help)CS1 maint: unflagged free DOI (link)
  • Sitaram, N.; Nagaraj, R. (2002), "Host-defense antimicrobial peptides: importance of structure for activity", Curr Pharm Des, 8 (9): 727–742, PMID 11945168
  • Tennessen, Jacob A. (2005), "Molecular evolution of animal antimicrobial peptide: widespread moderate positive selection", Journal of Evolutionary Biology, 18 (6): 1387–1394, doi:10.1111/j.1420-9101.2005.00925.x, PMID 16313451 {{citation}}: Unknown parameter |month= ignored (help)
  • Whitelock, J.M.; Murdoch, A.D.; Iozzo, R.V.; Underwood, P.A. (1996), "The Degradation of Human Endothelial Cell-derived Perlecan and Release of Bound Basic Fibroblast Growth Factor by Stromelysin, Collagenase, Plasmin, and Heparanases", Journal of Biological Chemistry, 271 (17): 10079–10086, doi:10.1074/jbc.271.17.10079, PMID 8626565 {{citation}}: |access-date= requires |url= (help); Unknown parameter |month= ignored (help)CS1 maint: unflagged free DOI (link)
  • Wildman, K.A.H.; Lee, D.K.; Ramamoorthy, A. (2003), "Mechanism of Lipid Bilayer Disruption by the Human Antimicrobial Peptide, LL-37", Biochemistry, 42 (21): 6545–6558, doi:10.1021/bi0273563, PMID 12767238
  • Yeaman, M.R.; Yount, N.Y. (2003), "Mechanisms of antimicrobial peptide action and resistance", Pharmacological reviews, 55 (1): 27–55, doi:10.1124/pr.55.1.2, PMID 12615953 {{citation}}: |access-date= requires |url= (help); Unknown parameter |month= ignored (help)
  • Brogden, K.A. (2005), "Antimicrobial peptides: pore formers or metabolic inhibitors in bacteria?", Nature Reviews Microbiology, 3 (3): 238–250., doi:10.1038/nrmicro1098, PMID 15703760 {{citation}}: Unknown parameter |month= ignored (help)