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...

Potassium Channels Blockers from the Venom of Androctonus mauretanicus mauretanicus

DOI: 10.1155/2012/103608

Full-Text   Cite this paper   Add to My Lib

Abstract:

K+ channels selectively transport K+ ions across cell membranes and play a key role in regulating the physiology of excitable and nonexcitable cells. Their activation allows the cell to repolarize after action potential firing and reduces excitability, whereas channel inhibition increases excitability. In eukaryotes, the pharmacology and pore topology of several structural classes of K+ channels have been well characterized in the past two decades. This information has come about through the extensive use of scorpion toxins. We have participated in the isolation and in the characterization of several structurally distinct families of scorpion toxin peptides exhibiting different K+ channel blocking functions. In particular, the venom from the Moroccan scorpion Androctonus mauretanicus mauretanicus provided several high-affinity blockers selective for diverse K+ channels ( S K C a , K v 4.x, and K v 1.x K+ channel families). In this paper, we summarize our work on these toxin/channel interactions. 1. The Scorpion Venom Content Scorpion venoms are very complex mixtures of molecules, constituting a diverse, naturally occurring peptide library, with most peptides displaying different kinds of biological activity [1, 2]. These peptides can specifically bind to a variety of pharmacological targets, in particular ion channels, resulting in neurotoxic effects. Toxins modulating Na+, K+, Ca++, and Cl? currents have been described in scorpion venoms [2]. Toxins that are highly lethal for mammals modify voltage-gated Na+ (Nav) currents in excitable cells and are referred to as “Nav channel long-chain toxin.” These toxins are single-chain, small, basic peptides (60- to 75-amino-acid residue chain generally folded by four disulfide bridges). They have been described as α- or β-toxins due to their binding site on Nav channels as well as to their pharmacological effects [1, 3]. α-Toxins bind in a voltage-dependent manner on the voltage sensor of the Nav channel domain IV and inhibit the inactivation phase of the action potential. β-Toxins act on the channel activation phase by binding to extracellular loops located preferentially on the voltage sensor of the Nav channel domain II (but also occasionally of domain III) [1, 3–6]. Another class of scorpion toxins has also been widely studied, even if these toxins are devoid of serious lethal effect. They block different K+ channel subtypes (some of them in the picomolar range) and are so-called “K+ channel toxins” [1, 2, 7]. They are usually shorter than Nav channel toxins but are structurally closely related to them.

References

[1]  M. F. Martin-Eauclaire and F. Couraud, “Scorpion neurotoxins: effects and mechanisms,” in Handbook Neurotoxicology, L. W. Chang and R. S. Dyer, Eds., pp. 683–716, Marcel Dekker, New York, NY, USA, 1995.
[2]  L. D. Possani, E. Merino, M. Corona, F. Bolivar, and B. Becerril, “Peptides and genes coding for scorpion toxins that affect ion-channels,” Biochimie, vol. 82, no. 9-10, pp. 861–868, 2000.
[3]  R. C. Rodríguez De La Vega and L. D. Possani, “Overview of scorpion toxins specific for Na+ channels and related peptides: biodiversity, structure-function relationships and evolution,” Toxicon, vol. 46, no. 8, pp. 831–844, 2005.
[4]  S. Cestèle and W. A. Catterall, “Molecular mechanisms of neurotoxin action on voltage-gated sodium channels,” Biochimie, vol. 82, no. 9-10, pp. 883–892, 2000.
[5]  F. Bosmans, M. F. Martin-Eauclaire, and K. J. Swartz, “Deconstructing voltage sensor function and pharmacology in sodium channels,” Nature, vol. 456, no. 7219, pp. 202–208, 2008.
[6]  F. Bosmans, M. Puopolo, M. F. Martin-Eauclaire, B. P. Bean, and K. J. Swartz, “Functional properties and toxin pharmacology of a dorsal root ganglion sodium channel viewed through its voltage sensors,” Journal of General Physiology, vol. 138, no. 1, pp. 59–72, 2011.
[7]  J. Tytgat, K. G. Chandy, M. L. Garcia et al., “A unified nomenclature for short-chain peptides isolated from scorpion venoms: α-KTx molecular subfamilies,” Trends in Pharmacological Sciences, vol. 20, no. 11, pp. 444–447, 1999.
[8]  F. Bontems, C. Roumestand, B. Gilquin, A. Ménez, and F. Toma, “Refined structure of charybdotoxin: common motifs in scorpion toxins and insect defensins,” Science, vol. 254, no. 5037, pp. 1521–1523, 1991.
[9]  O. Pongs, Regulation of Excitability by Potassium Channels. Inhibitory Regulation of Excitatory Neurotransmission, vol. 44 of Results and Problems in Cell Differentiation, Springer, Heidelberg, Germany, 2007.
[10]  G. A. Gutman, K. G. Chandy, S. Grissmer et al., “International Union of Pharmacology. LIII. Nomenclature and molecular relationships of voltage-gated potassium channels,” Pharmacological Reviews, vol. 57, no. 4, pp. 473–508, 2005.
[11]  S. B. Long, E. B. Campbell, and R. MacKinnon, “Crystal structure of a mammalian voltage-dependent Shaker family K+ channel,” Science, vol. 309, no. 5736, pp. 897–903, 2005.
[12]  C. S. Park and C. Miller, “Interaction of charybdotoxin with permeant ions inside the pore of a k+ channel,” Neuron, vol. 9, no. 2, pp. 307–313, 1992.
[13]  A. Lange, K. Giller, S. Hornig et al., “Toxin-induced conformational changes in a potassium channel revealed by solid-state NMR,” Nature, vol. 440, no. 7086, pp. 959–962, 2006.
[14]  K. N. Srinivasan, V. Sivaraja, I. Huys et al., “kappa-Hefutoxin1, a novel toxin from the scorpion Heterometrus fulvipes with unique structure and function. Importance of the functional diad in potassium channel selectivity,” Journal of Biological Chemistry, vol. 277, no. 33, pp. 30040–30047, 2002.
[15]  E. Diego-García, Y. Abdel-Mottaleb, E. F. Schwartz, R. C. R. De La Vega, J. Tytgat, and L. D. Possani, “Cytolytic and k+ channel blocking activities of β-KTx and scorpine-like peptides purified from scorpion venoms,” Cellular and Molecular Life Sciences, vol. 65, no. 1, pp. 187–200, 2008.
[16]  M. Corona, G. B. Gurrola, E. Merino et al., “A large number of novel Ergtoxin-like genes and ERG k+-channels blocking peptides from scorpions of the genus Centruroides,” FEBS Letters, vol. 532, no. 1-2, pp. 121–126, 2002.
[17]  N. Oukkache, M. F. Martin-Eauclaire, F. Chgoury et al., “Identification des paramètres influen?ant la prise en charge de l'envenimation scorpionique au Maroc,” in Toxines et Cancer, Rencontres en Toxinologie, pp. 295–300, LAVOISIER, 2006.
[18]  J. P. Rosso and H. Rochat, “Characterization of ten proteins from the venom of the Moroccan scorpion Androctonus mauretanicus mauretanicus, six of which are toxic to the mouse,” Toxicon, vol. 23, no. 1, pp. 113–125, 1985.
[19]  H. Zerrouk, P. E. Bougis, B. Ceard, A. Benslimane, and M. F. Martin-Eauclaire, “Analysis by high-performance liquid chromatography of Androctonus mauretanicus mauretanicus (black scorpion) venom,” Toxicon, vol. 29, no. 8, pp. 951–960, 1991.
[20]  M. Crest, G. Jacquet, M. Gola et al., “Kaliotoxin, a novel peptidyl inhibitor of neuronal BK-type Ca2+-activated k+ channels characterized from Androctonus mauretanicus mauretanicus venom,” Journal of Biological Chemistry, vol. 267, no. 3, pp. 1640–1647, 1992.
[21]  J. M. Sabatier, H. Zerrouk, H. Darbon et al., “P05, a new leiurotoxin I-like scorpion toxin: synthesis and structure-activity relationships of the α-amidated analog, a ligand of Ca2+-activated k+ channels with increased affinity,” Biochemistry, vol. 32, no. 11, pp. 2763–2770, 1993.
[22]  H. Vacher, M. Alami, M. Crest, L. D. Possani, P. E. Bougis, and M. F. Martin-Eauclaire, “Expanding the scorpion toxin α-KTX 15 family with AmmTX3 from Androctonus mauretanicus,” European Journal of Biochemistry, vol. 269, no. 24, pp. 6037–6041, 2002.
[23]  N. Oukkache, J. P. Rosso, M. Alami et al., “New analysis of the toxic compounds from the Androctonus mauretanicus mauretanicus scorpion venom,” Toxicon, vol. 51, no. 5, pp. 835–852, 2008.
[24]  J. Aiyar, J. M. Withka, J. P. Rizzi et al., “Topology of the pore-region of a k+ channel revealed by the NMR-derived structures of scorpion toxins,” Neuron, vol. 15, no. 5, pp. 1169–1181, 1995.
[25]  J. C. Fontecilla-Camps, C. Habersetzer-Rochat, and H. Rochat, “Orthorhombic crystals and three-dimensional structure of the potent toxin II from the scorpion Androctonus australis Hector,” Proceedings of the National Academy of Sciences of the United States of America, vol. 85, no. 20, pp. 7443–7447, 1988.
[26]  H. Zerrouk, F. Laraba-Djebari, V. Fremont et al., “Characterization of PO15 a new peptide ligand of the apamin-sensitive Ca2+ activated k+ channel,” International Journal of Peptide and Protein Research, vol. 48, no. 6, pp. 514–521, 1996.
[27]  Y. Abdel-Mottaleb, E. Clynen, A. Jalali et al., “The first potassium channel toxin from the venom of the Iranian scorpion Odonthobuthus doriae,” FEBS Letters, vol. 580, no. 26, pp. 6254–6258, 2006.
[28]  H. Darbon, E. Blanc, and J. M. Sabatier, “Three-dimensional structure of scorpion toxins: toward a new model of interaction with potassium channels,” Perspectives in Drug Discovery and Design, vol. 16, pp. 41–60, 1999.
[29]  P. Auguste, M. Hugues, B. Grave et al., “Leiurotoxin I (scyllatoxin), a peptide ligand for Ca2+-activated k+ channels. Chemical synthesis, radiolabeling, and receptor characterization,” Journal of Biological Chemistry, vol. 265, no. 8, pp. 4753–4759, 1990.
[30]  G. G. Chicchi, G. Gimenez-Callego, E. Ber, M. L. Garcia, R. Winquist, and M. A. Cascieri, “Purification and characterization of a unique, potent inhibitor of apamin binding from Leiurus quinquestriatus hebraeus venom,” Journal of Biological Chemistry, vol. 263, no. 21, pp. 10192–10197, 1988.
[31]  C. Labbe-Jullie, C. Granier, F. Albericio et al., “Binding and toxicity of apamin. Characterization of the active site,” European Journal of Biochemistry, vol. 196, no. 3, pp. 639–645, 1991.
[32]  C. Devaux, M. Knibiehler, M. L. Defendini et al., “Recombinant and chemical derivatives of apamin. Implication of post-transcriptional C-terminal amidation of apamin in biological activity,” European Journal of Biochemistry, vol. 231, no. 3, pp. 544–550, 1995.
[33]  M. Cui, J. Shen, J. M. Briggs et al., “Brownian dynamics simulations of the recognition of the scorpion toxin P05 with the small-conductance calcium-activated potassium channels,” Journal of Molecular Biology, vol. 318, no. 2, pp. 417–428, 2002.
[34]  C. Brugnara, C. C. Armsby, L. De Franceschi, M. Crest, M. F. Martin Euclaire, and S. L. Alper, “Ca2+-activated k+ channels of human and rabbit erythrocytes display distinctive patterns of inhibition by venom peptide toxins,” Journal of Membrane Biology, vol. 147, no. 1, pp. 71–82, 1995.
[35]  E. Honore, E. Guillemare, F. Lesage, J. Barhanin, and M. Lazdunski, “Injection of a k+ channel (Kv1.3) cRNA in fertilized eggs leads to functional expression in cultured myotomal muscle cells from Xenopus embryos,” FEBS Letters, vol. 348, no. 3, pp. 259–262, 1994.
[36]  M. D. Cahalan and K. G. Chandy, “The functional network of ion channels in T lymphocytes,” Immunological Reviews, vol. 231, no. 1, pp. 59–87, 2009.
[37]  C. Beeton, J. Barbaria, P. Giraud et al., “Selective blocking of voltage-gated k+ channels improves experimental autoimmune encephalomyelitis and inhibits T cell activation,” Journal of Immunology, vol. 166, no. 2, pp. 936–944, 2001.
[38]  P. Valverde, T. Kawai, and M. A. Taubman, “Selective blockade of voltage-gated potassium channels reduces inflammatory bone resorption in experimental periodontal disease,” Journal of Bone and Mineral Research, vol. 19, no. 1, pp. 155–164, 2004.
[39]  C. Mourre, M. N. Chernova, M. F. Martin-Eauclaire et al., “Distribution in rat brain of binding sites of kaliotoxin, a blocker of Kv1.1 and Kv1.3 α-subunits,” Journal of Pharmacology and Experimental Therapeutics, vol. 291, no. 3, pp. 943–952, 1999.
[40]  S. Kourrich, C. Mourre, and B. Soumireu-Mourat, “Kaliotoxin, a Kv1.1 and Kv1.3 channel blocker, improves associative learning in rats,” Behavioural Brain Research, vol. 120, no. 1, pp. 35–46, 2001.
[41]  R. Romi, M. Crest, M. Gola et al., “Synthesis and characterization of kaliotoxin. Is the 26-32 sequence essential for potassium channel recognition?” Journal of Biological Chemistry, vol. 268, no. 35, pp. 26302–26309, 1993.
[42]  G. M. Lipkind and H. A. Fozzard, “A model of scorpion toxin binding to voltage-gated k+ channels,” Journal of Membrane Biology, vol. 158, no. 3, pp. 187–196, 1997.
[43]  C. Ader, R. Schneider, S. Hornig et al., “A structural link between inactivation and block of a k+ channel,” Nature Structural and Molecular Biology, vol. 15, no. 6, pp. 605–612, 2008.
[44]  C. Ader, O. Pongs, S. Becker, and M. Baldus, “Protein dynamics detected in a membrane-embedded potassium channel using two-dimensional solid-state NMR spectroscopy,” Biochimica et Biophysica Acta, vol. 1798, no. 2, pp. 286–290, 2010.
[45]  U. Zachariae, R. Schneider, P. Velisetty et al., “The molecular mechanism of toxin-induced conformational changes in a potassium channel: relation to C-type inactivation,” Structure, vol. 16, no. 5, pp. 747–754, 2008.
[46]  M. Pisciotta, F. I. Coronas, C. Bloch, G. Prestipino, and L. D. Possani, “Fast k+ currents from cerebellum granular cells are completely blocked by a peptide purified from Androctonus australis Garzoni scorpion venom,” Biochimica et Biophysica Acta, vol. 1468, no. 1-2, pp. 203–212, 2000.
[47]  H. Vacher, R. Romi-Lebrun, C. Mourre et al., “A new class of scorpion toxin binding sites related to an A-type k+ channel: pharmacological characterization and localization in rat brain,” FEBS Letters, vol. 501, no. 1-3, pp. 31–36, 2001.
[48]  C. Legros, P. E. Bougis, and M. F. Martin-Eauclaire, “Characterisation of the genes encoding Aa1 isoforms from the scorpion Androctonus australis,” Toxicon, vol. 41, no. 1, pp. 115–119, 2003.
[49]  T. Chen, B. Walker, M. Zhou, and C. Shaw, “Molecular cloning of a novel putative potassium channel-blocking neurotoxin from the venom of the North African scorpion, Androctonus amoreuxi,” Peptides, vol. 26, no. 5, pp. 731–736, 2005.
[50]  M. C. Sanguinetti, J. H. Johnson, L. G. Hammerland et al., “Heteropodatoxins: peptides isolated from spider venom that block Kv4.2 potassium channels,” Molecular Pharmacology, vol. 51, no. 3, pp. 491–498, 1997.
[51]  W. A. Coetzee, Y. Amarillo, J. Chiu et al., “Molecular diversity of k+ channels,” Annals of the New York Academy of Sciences, vol. 868, pp. 233–285, 1999.
[52]  H. Vacher and J. S. Trimmer, “Diverse roles for auxiliary subunits in phosphorylation-dependent regulation of mammalian brain voltage-gated potassium channels,” Pflugers Archiv European Journal of Physiology, vol. 642, no. 5, pp. 631–643, 2011.
[53]  H. Vacher, S. Diochot, P. E. Bougis, M. F. Martin-Eauclaire, and C. Mourre, “Kv4 channels sensitive to BmTX3 in rat nervous system: autoradiographic analysis of their distribution during brain ontogenesis,” European Journal of Neuroscience, vol. 24, no. 5, pp. 1325–1340, 2006.
[54]  Y. Abdel-Mottaleb, G. Corzo, M. F. Martin-Eauclaire et al., “A common "hot spot" confers hERG blockade activity to α-scorpion toxins affecting k+ channels,” Biochemical Pharmacology, vol. 76, no. 6, pp. 805–815, 2008.
[55]  J. Amendola, A. Woodhouse, M. F. Martin-Eauclaire, and J. M. Goaillard, “Ca2+/cAMP-Sensitive co-variation of IA and IH voltage dependences tune rebound firing in dopaminergic neurons,” The Journal of Neuroscience, vol. 32, no. 6, pp. 2166–2181, 2012.
[56]  L. R. Phillips, M. Milescu, Y. Li-Smerin, J. A. Mindell, J. I. Kim, and K. J. Swartz, “Voltage-sensor activation with a tarantula toxin as cargo,” Nature, vol. 436, no. 7052, pp. 857–860, 2005.
[57]  H. Vacher, R. Romi-Lebrun, M. Crest et al., “Functional consequences of deleting the two C-terminal residues of the scorpion toxin BmTX3,” Biochimica et Biophysica Acta, vol. 1646, no. 1-2, pp. 152–156, 2003.
[58]  G. N. Tseng, K. D. Sonawane, Y. V. Korolkova et al., “Probing the outer mouth structure of the hERG channel with peptide toxin footprinting and molecular modeling,” Biophysical Journal, vol. 92, no. 10, pp. 3524–3540, 2007.
[59]  R. C. Rodríguez De La Vega, E. Merino, B. Becerril, and L. D. Possani, “Novel interactions between k+ channels and scorpion toxins,” Trends in Pharmacological Sciences, vol. 24, no. 5, pp. 222–227, 2003.
[60]  I. Huys, C. Q. Xu, C. Z. Wang et al., “BmTx3, a scorpion toxin with two putative functional faces separately active on A-type k+ and HERG currents,” Biochemical Journal, vol. 378, no. 3, pp. 745–752, 2004.

Full-Text

comments powered by Disqus

Contact Us

service@oalib.com

QQ:3279437679

WhatsApp +8615387084133