全部 标题 作者
关键词 摘要

OALib Journal期刊
ISSN: 2333-9721
费用:99美元

查看量下载量

相关文章

更多...

Host-Defense Peptides with Therapeutic Potential from Skin Secretions of Frogs from the Family Pipidae

DOI: 10.3390/ph7010058

Keywords: frog skin, magainin, PGLa, caerulein-precursor fragment, xenopsin-precursor-fragment, hymenochirin

Full-Text   Cite this paper   Add to My Lib

Abstract:

Skin secretions from frogs belonging to the genera Xenopus, Silurana, Hymenochirus, and Pseudhymenochirus in the family Pipidae are a rich source of host-defense peptides with varying degrees of antimicrobial activities and cytotoxicities to mammalian cells. Magainin, peptide glycine-leucine-amide (PGLa), caerulein-precursor fragment (CPF), and xenopsin-precursor fragment (XPF) peptides have been isolated from norepinephrine-stimulated skin secretions from several species of Xenopus and Silurana. Hymenochirins and pseudhymenochirins have been isolated from Hymenochirus boettgeri and Pseudhymenochirus merlini. A major obstacle to the development of these peptides as anti-infective agents is their hemolytic activities against human erythrocytes. Analogs of the magainins, CPF peptides and hymenochirin-1B with increased antimicrobial potencies and low cytotoxicities have been developed that are active (MIC < 5 μM) against multidrug-resistant clinical isolates of Staphylococcus aureus, Escherichia coli, Acinetobacter baumannii, Stenotrophomonas maltophilia and Klebsiella pneumoniae. Despite this, the therapeutic potential of frog skin peptides as anti-infective agents has not been realized so that alternative clinical applications as anti-cancer, anti-viral, anti-diabetic, or immunomodulatory drugs are being explored.

References

[1]  Savard, P.; Perl, T.M. A call for action: Managing the emergence of multidrug-resistant Enterobacteriaceae in the acute care settings. Curr. Opin. Infect. Dis. 2012, 25, 371–377, doi:10.1097/QCO.0b013e3283558c17.
[2]  Conlon, J.M. The contribution of skin antimicrobial peptides to the system of innate immunity in anurans. Cell Tissue Res. 2011, 343, 201–212, doi:10.1007/s00441-010-1014-4.
[3]  Conlon, J.M. Structural diversity and species distribution of host-defense peptides in frog skin secretions. Cell. Mol. Life Sci. 2011, 68, 2303–2315, doi:10.1007/s00018-011-0720-8.
[4]  Yeung, A.T.Y.; Gellatly, S.L.; Hancock, R.E. Multifunctional cationic host defence peptides and their clinical applications. Cell. Mol. Life Sci. 2011, 68, 2161–2176, doi:10.1007/s00018-011-0710-x.
[5]  Gammill, W.M.; Fites, J.S.; Rollins-Smith, L.A. Norepinephrine depletion of antimicrobial peptides from the skin glands of Xenopus laevis. Dev. Comp. Immunol. 2012, 37, 19–27, doi:10.1016/j.dci.2011.12.012.
[6]  Almeida, P.F.; Pokorny, A. Mechanisms of antimicrobial, cytolytic, and cell-penetrating peptides: From kinetics to thermodynamics. Biochemistry 2009, 48, 8083–8093, doi:10.1021/bi900914g.
[7]  Huang, Y.; Huang, J.; Chen, Y. Alpha-helical cationic antimicrobial peptides: Relationships of structure and function. Protein Cell 2010, 1, 143–152, doi:10.1007/s13238-010-0004-3.
[8]  Tennessen, J.A.; Woodhams, D.C.; Chaurand, P.; Reinert, L.K.; Billheimer, D.; Shyr, Y.; Caprioli, R.M.; Blouin, M.S.; Rollins-Smith, L.A. Variations in the expressed antimicrobial peptide repertoire of northern leopard frog (Rana pipiens) populations suggest intraspecies differences in resistance to pathogens. Dev. Comp. Immunol. 2009, 33, 1247–1257, doi:10.1016/j.dci.2009.07.004.
[9]  Conlon, J.M.; Al-Ghaferi, N.; Abraham, B.; Leprince, J. Strategies for transformation of naturally-occurring amphibian antimicrobial peptides into therapeutically valuable anti-infective agents. Methods 2007, 42, 349–357, doi:10.1016/j.ymeth.2007.01.004.
[10]  Jiang, Z.; Vasil, A.I.; Hale, J.D.; Hancock, R.E.; Vasil, M.L.; Hodges, R.S. Effects of net charge and the number of positively charged residues on the biological activity of amphipathic alpha-helical cationic antimicrobial peptides. Biopolymers 2008, 90, 369–383, doi:10.1002/bip.20911.
[11]  Matsuzaki, K. Control of cell selectivity of antimicrobial peptides. Biochim. Biophys. Acta 2009, 1788, 1687–1692, doi:10.1016/j.bbamem.2008.09.013.
[12]  Frost, D.R. Amphibian Species of the World: An Online Reference, Version 5.6. 2013, Electronic Database. American Museum of Natural History: New York, USA. Available online: http://research.amnh.org/herpetology/amphibia/index.php (accessed on 13 January 2014).
[13]  Frost, D.R.; Grant, T.; Faivovich, J.; Bain, R.H.; Haas, A.; Haddad, C.F.B.; de Sá, R.O.; Channing, A.; Wilkinson, M.; Donnellan, S.C.; et al. The amphibian tree of life. Bull. Am. Mus. Nat. Hist. 2006, 297, 1–370, doi:10.1206/0003-0090(2006)297[0001:TATOL]2.0.CO;2.
[14]  Evans, B.J.; Kelley, D.B.; Tinsley, R.C.; Melnick, D.J.; Cannatella, D.C. A mitochondrial DNA phylogeny of African clawed frogs: Phylogeography and implications for polyploid evolution. Mol. Phylogenet. Evol. 2004, 33, 197–213, doi:10.1016/j.ympev.2004.04.018.
[15]  Evans, B.J. Genome evolution and speciation genetics of clawed frogs (Xenopus and Silurana). Front. Biosci. 2008, 13, 4687–4706, doi:10.2741/3033.
[16]  Irisarri, I.; Vences, M.; San Mauro, D.; Glaw, F.; Zardoya, R. Reversal to air-driven sound production revealed by a molecular phylogeny of tongueless frogs, family Pipidae. BMC Evol. Biol. 2011, 11, doi:10.1186/1471-2148-11-114.
[17]  Bewick, A.J.; Chain, F.J.; Heled, J.; Evans, B.J. The pipid root. Syst. Biol. 2012, 61, 913–926, doi:10.1093/sysbio/sys039.
[18]  Roelants, K.; Bossuyt, F. Archaeobatrachian paraphyly and pangaean diversification of crown- group frogs. Syst. Biol. 2005, 54, 111–126, doi:10.1080/10635150590905894.
[19]  Roelants, K.; Gower, D.J.; Wilkinson, M.; Loader, S.P.; Biju, S.D.; Guillaume, K.; Moriau, L.; Bossuyt, F. Global patterns of diversification in the history of modern amphibians. Proc. Natl. Acad. Sci. USA 2007, 104, 887–892, doi:10.1073/pnas.0608378104.
[20]  Báez, A.M. The Fossil Record of the Pipidae. In The Biology of Xenopus; Tinsley, R.C., Kobel, H.R., Eds.; Clarendon Press: Oxford, UK, 1996; pp. 329–347.
[21]  Kobel, H.R.; Loumont, C.; Tinsley, R.C. The Extant Species. In The Biology of Xenopus; Tinsley, R.C., Kobel, H.R., Eds.; Clarendon Press: Oxford, UK, 1996; pp. 9–33.
[22]  Evans, B.J.; Greenbaum, E.; Kusamba, C.; Carter, T.F.; Tobias, M.L.; Mendel, S.A.; Kelley, D.B. Description of a new octoploid frog species (Anura: Pipidae: Xenopus) from the Democratic Republic of the Congo, with a discussion of the biogeography of African clawed frogs in the Albertine Rift. J. Zool. 2011, 283, 276–290, doi:10.1111/j.1469-7998.2010.00769.x.
[23]  Tymowska, J.; Fischberg, M. A comparison of the karyotype, constitutive heterochromatin, and nucleolar organizer regions of the new tetraploid species Xenopus epitropicalis Fischberg and Picard with those of Xenopus tropicalis Gray (Anura, Pipidae). Cytogenet. Cell Genet. 1982, 34, 49–157.
[24]  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, doi:10.1073/pnas.84.15.5449.
[25]  Gibson, B.W.; Poulter, L.; Williams, D.H.; Maggio, J.E. Novel peptide fragments originating from PGLa and the caerulein and xenopsin precursors from Xenopus laevis. J. Biol. Chem. 1986, 261, 5341–5349.
[26]  Soravia, E.; Martini, G.; Zasloff, M. Antimicrobial properties of peptides from Xenopus granular gland secretions. FEBS Lett. 1988, 228, 337–340, doi:10.1016/0014-5793(88)80027-9.
[27]  Hunt, L.T.; Barker, W.C. Relationship of promagainin to three other prohormones from the skin of Xenopus laevis: A different perspective. FEBS Lett. 1988, 233, 282–288, doi:10.1016/0014-5793(88)80443-5.
[28]  Conlon, J.M.; Al-Ghaferi, N.; Ahmed, E.; Meetani, M.A.; Leprince, J.; Nielsen, P.F. Orthologs of magainin, PGLa, procaerulein-derived, and proxenopsin-derived peptides from skin secretions of the octoploid frog Xenopus amieti (Pipidae). Peptides 2010, 31, 989–994, doi:10.1016/j.peptides.2010.03.002.
[29]  Mechkarska, M.; Ahmed, E.; Coquet, L.; Leprince, J.; Jouenne, T.; Vaudry, H.; King, J.D.; Takada, K.; Conlon, J.M. Genome duplications within the Xenopodinae do not increase the multiplicity of antimicrobial peptides in Silurana paratropicalis and Xenopus andrei skin secretions. Comp. Biochem. Physiol. D Genomics Proteomics 2011, 6, 206–212, doi:10.1016/j.cbd.2011.03.003.
[30]  Mechkarska, M.; Ahmed, E.; Coquet, L.; Leprince, J.; Jouenne, T.; Vaudry, H.; King, J.D.; Conlon, J.M. Antimicrobial peptides with therapeutic potential from skin secretions of the Marsabit clawed frog Xenopus borealis (Pipidae). Comp. Biochem. Physiol. C Toxicol. Pharmacol. 2010, 152, 467–472, doi:10.1016/j.cbpc.2010.07.007.
[31]  Conlon, J.M.; Mechkarska, M.; Ahmed, E.; Leprince, J.; Vaudry, H.; King, J.D.; Takada, K. Purification and properties of antimicrobial peptides from skin secretions of the Eritrea clawed frog Xenopus clivii (Pipidae). Comp. Biochem. Physiol. C Toxicol. Pharmacol. 2011, 153, 350–354, doi:10.1016/j.cbpc.2010.12.007.
[32]  King, J.D.; Mechkarska, M.; Coquet, L.; Leprince, J.; Jouenne, T.; Vaudry, H.; Takada, K.; Conlon, J.M. Host-defense peptides from skin secretions of the tetraploid frogs Xenopus petersii and Xenopus pygmaeus, and the octoploid frog Xenopus lenduensis (Pipidae). Peptides 2012, 33, 35–43, doi:10.1016/j.peptides.2011.11.015.
[33]  Mechkarska, M.; Ahmed, E.; Coquet, L.; Leprince, J.; Jouenne, T.; Vaudry, H.; King, J.D.; Conlon, J.M. Peptidomic analysis of skin secretions demonstrates that the allopatric populations of Xenopus muelleri (Pipidae) are not conspecific. Peptides 2011, 32, 1502–1508, doi:10.1016/j.peptides.2011.05.025.
[34]  King, J.D.; Mechkarska, M.; Meetani, M.A.; Conlon, J.M. Peptidomic analysis of skin secretions provides insight into the taxonomic status of the African clawed frogs Xenopus victorianus and Xenopus laevis sudanensis (Pipidae). Comp. Biochem. Physiol. D Genomics Proteomics 2013, 8, 250–254, doi:10.1016/j.cbd.2013.07.001.
[35]  Mechkarska, M.; Meetani, M.; Michalak, P.; Vaksman, Z.; Takada, K.; Conlon, J.M. Hybridization between the tetraploid African clawed frogs Xenopus laevis and Xenopus muelleri (Pipidae) increases the multiplicity of antimicrobial peptides in the skin secretions of female offspring. Comp. Biochem. Physiol. D Genomics Proteomics 2012, 7, 285–291, doi:10.1016/j.cbd.2012.05.002.
[36]  Mechkarska, M.; Prajeep, M.; Leprince, J.; Vaudry, H.; Meetani, M.A.; Evans, B.J.; Conlon, J.M. A comparison of host-defense peptides in skin secretions of female Xenopus laevis × Xenopus borealis and X. borealis × X. laevis F1 hybrids. Peptides 2013, 45, 1–8, doi:10.1016/j.peptides.2013.04.008.
[37]  Ali, M.F.; Soto, A.; Knoop, F.C.; Conlon, J.M. Antimicrobial peptides isolated from skin secretions of the diploid frog, Xenopus tropicalis (Pipidae). Biochim. Biophys. Acta 2001, 1550, 81–89.
[38]  Roelants, K.; Fry, B.G.; Ye, L.; Stijlemans, B.; Brys, L.; Kok, P.; Clynen, E.; Schoofs, L.; Cornelis, P.; Bossuyt, F. Origin and functional diversification of an amphibian defense peptide arsenal. PLoS Genet. 2013, 9, e1003662, doi:10.1371/journal.pgen.1003662.
[39]  Conlon, J.M.; Mechkarska, M.; Prajeep, M.; Sonnevend, A.; Coquet, L.; Leprince, J.; Jouenne, T.; Vaudry, H.; King, J.D. Host-defense peptides in skin secretions of the tetraploid frog Silurana epitropicalis with potent activity against methicillin-resistant Staphylococcus aureus (MRSA). Peptides 2012, 37, 113–119, doi:10.1016/j.peptides.2012.07.005.
[40]  Mechkarska, M.; Prajeep, M.; Coquet, L.; Leprince, J.; Jouenne, T.; Vaudry, H.; King, J.D.; Conlon, J.M. The hymenochirins: A family of antimicrobial peptides from the Congo dwarf clawed frog Hymenochirus boettgeri (Pipidae). Peptides 2012, 35, 269–275, doi:10.1016/j.peptides.2012.03.029.
[41]  Conlon, J.M.; Prajeep, M.; Mechkarska, M.; Coquet, L.; Leprince, J.; Jouenne, T.; Vaudry, H.; King, J.D. Characterization of the host-defense peptides from skin secretions of Merlin’s clawed frog Pseudhymenochirus merlini: Insights into phylogenetic relationships among the Pipidae. Comp. Biochem. Physiol. D Genomics Proteomics 2013, 8, 352–357, doi:10.1016/j.cbd.2013.10.002.
[42]  Helmerhorst, E.J.; Reijnders, M.; van’t Hof, W.; Veerman, C.; Nieuw-Amerongen, A.V. A critical comparison of the hemolytic and fungicidal activities of cationic antimicrobial peptides. FEBS Lett. 1999, 449, 105–110, doi:10.1016/S0014-5793(99)00411-1.
[43]  Imura, Y.; Choda, N.; Matsuzaki, K. Dagainin 2 in action: Distinct modes of membrane permeabilization in living bacterial and mammalian cells. Biophys. J. 2008, 95, 5757–5765, doi:10.1529/biophysj.108.133488.
[44]  Tamba, Y.; Ariyama, H.; Levadny, V.; Yamazaki, M. Kinetic pathway of antimicrobial peptide magainin 2-induced pore formation in lipid membranes. J. Phys. Chem. B 2010, 114, 12018–12026, doi:10.1021/jp104527y.
[45]  Epand, R.F.; Maloy, W.L.; Ramamoorthy, A.; Epand, R.M. Probing the “charge cluster mechanism” in amphipathic helical cationic antimicrobial peptides. Biochemistry 2010, 49, 4076–4084, doi:10.1021/bi100378m.
[46]  Zasloff, M.; Martin, B.; Chen, H.C. Antimicrobial activity of synthetic magainin peptides and several analogues. Proc. Natl. Acad. Sci. USA 1988, 85, 910–913, doi:10.1073/pnas.85.3.910.
[47]  Cuervo, J.H.; Rodriguez, B.; Houghten, R.A. The magainins: Sequence factors relevant to increased antimicrobial activity and decreased hemolytic activity. Pept. Res. 1988, 1, 81–86.
[48]  Shin, S.Y.; Kang, J.H.; Lee, M.K.; Kim, S.Y.; Kim, Y.; Hahm, K.S. Cecropin A—Magainin 2 hybrid peptides having potent antimicrobial activity with low hemolytic effect. Biochem. Mol. Biol. Int. 1998, 44, 1119–1126.
[49]  Fuchs, P.C.; Barry, A.L.; Brown, S.D. In vitro antimicrobial activity of MSI-78, a magainin analog. Antimicrob. Agents Chemother. 1998, 42, 1213–1216.
[50]  Ge, Y.; MacDonald, D.L.; Holroyd, K.J.; Thornsberry, C.; Wexler, H.; Zasloff, M. In vitro antibacterial properties of pexiganan, an analog of magainin. Antimicrob. Agents Chemother. 1999, 43, 782–788.
[51]  Ge, Y.; MacDonald, D.; Henry, M.M.; Hait, H.I.; Nelson, K.A.; Lipsky, B.A.; Zasloff, M.A.; Holroyd, K.J. In vitro susceptibility to pexiganan of bacteria isolated from infected diabetic foot ulcers. Diagn. Microbiol. Infect. Dis. 1999, 35, 45–53, doi:10.1016/S0732-8893(99)00056-5.
[52]  Lipsky, B.A.; Holroyd, K.J.; Zasloff, M. Topical versus systemic antimicrobial therapy for treating mildly infected diabetic foot ulcers: A randomized, controlled, double-blinded, multicenter trial of pexiganan cream. Clin. Infect. Dis. 2008, 47, 1537–1545, doi:10.1086/593185.
[53]  Iwahori, A.; Hirota, Y.; Sampe, R.; Miyano, S.; Takahashi, N.; Sasatsu, M.; Kondo, I.; Numao, N. On the antibacterial activity of normal and reversed magainin 2 analogs against Helicobacter pylori. Biol. Pharm. Bull. 1997, 20, 805–808, doi:10.1248/bpb.20.805.
[54]  Macias, E.A.; Rana, F.; Blazyk, J.; Modrzakowski, M.C. Bactericidal activity of magainin 2: Use of lipopolysaccharide mutants. Can. J. Microbiol. 1990, 36, 582–584, doi:10.1139/m90-102.
[55]  Genco, C.A.; Maloy, W.L.; Kari, U.P.; Motley, M. Antimicrobial activity of magainin analogues against anaerobic oral pathogens. Int. J. Antimicrob. Agents 2003, 21, 75–78, doi:10.1016/S0924-8579(02)00275-3.
[56]  Schuster, F.L.; Jacob, L.S. Effects of magainins on ameba and cyst stages of Acanthamoeba polyphaga. Antimicrob. Agents Chemother. 1992, 36, 1263–1271, doi:10.1128/AAC.36.6.1263.
[57]  Helmerhorst, E.J.; Reijnders, I.M.; van’t Hof, W.; Simoons-Smit, I.; Veerman, E.C.; Amerongen, A.V. Amphotericin B- and fluconazole-resistant Candida spp., Aspergillus fumigatus, and other newly emerging pathogenic fungi are susceptible to basic antifungal peptides. Antimicrob. Agents Chemother. 1999, 43, 702–704.
[58]  Westerhoff, H.V.; Zasloff, M.; Rosner, J.L.; Hendler, R.W.; de Waal, A.; Vaz Gomes, A.; Jongsma, P.M.; Riethorst, A.; Jureti?, D. Functional synergism of the magainins PGLa and magainin-2 in Escherichia coli, tumor cells and liposomes. Eur. J. Biochem. 1995, 228, 257–264, doi:10.1111/j.1432-1033.1995.00257.x.
[59]  Lohner, K.; Prossnigg, F. Biological activity and structural aspects of PGLa interaction with membrane mimetic systems. Biochim. Biophys. Acta 2009, 1788, 1656–1666, doi:10.1016/j.bbamem.2009.05.012.
[60]  Strandberg, E.; Zerweck, J.; Wadhwani, P.; Ulrich, A.S. Synergistic insertion of antimicrobial magainin-family peptides in membranes depends on the lipid spontaneous curvature. Biophys. J. 2013, 104, L9–L11, doi:10.1016/j.bpj.2013.01.047.
[61]  Conlon, J.M.; Sonnevend, A.; Pál, T.; Vila-Farrés, X. Efficacy of six frog skin-derived antimicrobial peptides against colistin-resistant strains of the Acinetobacter baumannii group. Int. J. Antimicrob. Agents 2012, 39, 317–320, doi:10.1016/j.ijantimicag.2011.12.005.
[62]  Conlon, J.M.; Galadari, S.; Raza, H.; Condamine, E. Design of potent, non-toxic antimicrobial agents based upon the naturally occurring frog skin peptides, ascaphin-8 and peptide XT-7. Chem. Biol. Drug Des. 2008, 72, 58–64, doi:10.1111/j.1747-0285.2008.00671.x.
[63]  Subasinghage, A.P.; Conlon, J.M.; Hewage, C.M. Development of potent anti-infective agents from Silurana tropicalis: Conformational analysis of the amphipathic, alpha-helical antimicrobial peptide XT-7 and its non-haemolytic analogue [G4K]XT-7. Biochim. Biophys. Acta 2010, 1804, 1020–1028, doi:10.1016/j.bbapap.2010.01.015.
[64]  Moore, K.S.; Bevins, C.L.; Brasseur, M.M.; Tomassini, N.; Turner, K.; Eck, H.; Zasloff, M. Antimicrobial peptides in the stomach of Xenopus laevis. J. Biol. Chem. 1991, 266, 19851–19857.
[65]  Mechkarska, M.; Prajeep, M.; Radosavljevic, G.D.; Jovanovic, I.P.; Al Baloushi, A.; Sonnevend, A.; Lukic, M.L.; Conlon, J.M. An analog of the host-defense peptide hymenochirin-1B with potent broad-spectrum activity against multidrug-resistant bacteria and immunomodulatory properties. Peptides 2013, 50, 153–159, doi:10.1016/j.peptides.2013.10.015.
[66]  Mader, J.S.; Hoskin, D.W. Cationic antimicrobial peptides as novel cytotoxic agents for cancer treatment. Expert Opin. Investig. Drugs 2006, 15, 933–946, doi:10.1517/13543784.15.8.933.
[67]  Ohsaki, Y.; Gazdar, A.F.; Chen, H.C.; Johnson, B.E. Antitumor activity of magainin analogues against human lung cancer cell lines. Cancer Res. 1992, 52, 3534–3538.
[68]  Lehmann, J.; Retz, M.; Sidhu, S.S.; Suttmann, H.; Sell, M.; Paulsen, F.; Harder, J; Unteregger, G.; St?ckle, M. Antitumor activity of the antimicrobial peptide magainin II against bladder cancer cell lines. Eur. Urol. 2006, 50, 141–147, doi:10.1016/j.eururo.2005.12.043.
[69]  Cruciani, R.A.; Barker, J.L.; Zasloff, M.; Chen, H.C.; Colamonici, O. Antibiotic magainins exert cytolytic activity against transformed cell lines through channel formation. Proc. Natl. Acad. Sci. USA 1991, 88, 3792–3796, doi:10.1073/pnas.88.9.3792.
[70]  Kosza?ka, P.; Kamysz, E.; Wejda, M.; Kamysz, W.; Bigda, J. Antitumor activity of antimicrobial peptides against U937 histiocytic cell line. Acta. Biochim. Pol. 2011, 58, 111–117.
[71]  Baker, M.A.; Maloy, W.L.; Zasloff, M.; Jacob, L.S. Anticancer efficacy of magainin 2 and analogue peptides. Anticancer efficacy of magainin 2 and analogue peptides. Cancer Res. 1993, 53, 3052–3057.
[72]  Miyazaki, Y.; Aoki, M.; Yano, Y.; Matsuzaki, K. Interaction of antimicrobial peptide magainin 2 with gangliosides as a target for human cell binding. Biochemistry 2012, 51, 10229–10235, doi:10.1021/bi301470h.
[73]  Attoub, S.; Arafat, H.; Mechkarska, M.; Conlon, J.M. Anti-tumor activities of the host-defense peptide hymenochirin-1B. Regul. Pept. 2013, 115, 141–149.
[74]  Albiol Matanic, V.C.; Castilla, V. Antiviral activity of antimicrobial cationic peptides against Junin virus and herpes simplex virus. Int. J. Antimicrob. Agents 2004, 23, 382–389, doi:10.1016/j.ijantimicag.2003.07.022.
[75]  Chinchar, V.G.; Bryan, L.; Silphadaung, U.; Noga, E.; Wade, D.; Rollins-Smith, L. Inactivation of viruses infecting ectothermic animals by amphibian and piscine antimicrobial peptides. Virology 2004, 323, 268–275, doi:10.1016/j.virol.2004.02.029.
[76]  Dean, R.E.; O’Brien, L.M.; Thwaite, J.E.; Fox, M.A.; Atkins, H.; Ulaeto, D.O. A carpet-based mechanism for direct antimicrobial peptide activity against vaccinia virus membranes. Peptides 2010, 31, 1966–1972, doi:10.1016/j.peptides.2010.07.028.
[77]  Srinivasan, D.; Mechkarska, M.; Abdel-Wahab, Y.H.; Flatt, P.R.; Conlon, J.M. Caerulein precursor fragment (CPF) peptides from the skin secretions of Xenopus laevis and Silurana epitropicalis are potent insulin-releasing agents. Biochimie 2013, 95, 429–435, doi:10.1016/j.biochi.2012.10.026.
[78]  Ojo, O.O.; Conlon, J.M.; Flatt, P.R.; Abdel-Wahab, Y.H. Frog skin peptides (tigerinin-1R, magainin-AM1, -AM2, CPF-AM1, and PGla-AM1) stimulate secretion of glucagon-like peptide 1 (GLP-1) by GLUTag cells. Biochem. Biophys. Res. Commun. 2013, 431, 14–18, doi:10.1016/j.bbrc.2012.12.116.
[79]  Pantic, J.M.; Mechkarska, M.; Lukic, M.L.; Conlon, J.M. Effects of tigerinin peptides on cytokine production by mouse peritoneal macrophages and spleen cells and by human peripheral blood mononuclear cells. Biochimie 2014. in press.
[80]  Kotsaki, A.; Giamarellos-Bourboulis, E.J. Emerging drugs for the treatment of sepsis. Expert Opin. Emerg. Drugs 2012, 17, 379–391, doi:10.1517/14728214.2012.697151.

Full-Text

comments powered by Disqus

Contact Us

service@oalib.com

QQ:3279437679

WhatsApp +8615387084133