Hostname: page-component-7479d7b7d-pfhbr Total loading time: 0 Render date: 2024-07-08T19:34:21.801Z Has data issue: false hasContentIssue false

Structure and function of channel-forming peptaibols

Published online by Cambridge University Press:  17 March 2009

M. S. P. Sansom
Affiliation:
Laboratory of Molecular Biophysics, the Rex Richards Building, University of Oxford, South Parks Road, Oxford OX13QU

Extract

Transport of ions through channels is fundamental to a number of physiological processes, especially the electrical properties of excitable cells (Hille, 1992). To understand this process at a molecular level requires atomic resolution structures of channel proteins.

Type
Research Article
Copyright
Copyright © Cambridge University Press 1993

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Aléman, C., Subirana, J. A. & Perez, J. J. (1992). A molecular mechanical study of the structure of poly(α-aminoisobutyric acid). Biopolymers 32, 621631.CrossRefGoogle ScholarPubMed
Åqvist, J. & Warshel, A. (1989). Energetics of ion permeation through membrane channels: solvation of Na+ by gramicidin A. Biophys. J. 56, 171182.CrossRefGoogle Scholar
Archer, S. H. & Cafiso, D. S. (1991). Voltage-dependent conductance for alamethicin in phospholipid vesicles: a test for the mechanism of gating. Biophys J. 60, 380388.CrossRefGoogle ScholarPubMed
Archer, S. J., Ellena, J. F. & Cafiso, D. S. (1991). Dynamics and aggregation of the peptide ion channel alamethicin: measurement using spin-labeled peptides. Biophys. J. 60, 389398.CrossRefGoogle ScholarPubMed
Balaram, P. (1992). Non-standard amino acids in peptide design and protein engineering. Curr. Opin. Struct. Biol. 2, 845851.CrossRefGoogle Scholar
Balaram, P., Sukumar, M., Krishna, K., Mellor, I. R. & Sansom, M. S. P. (1992). The properties of ion channels formed by zervamicins. Eur. Biophys. J. 21, 117128.CrossRefGoogle ScholarPubMed
Balaram, P., Mellor, I. R., Kerr, I. D., Krishna, K., Sukumar, S. & Sansom, M. S. P. (1993). Ion channels formed by MS18, a synthetic amphipathic peptide. (Manuscript in preparation.)Google Scholar
Balasubramanian, T. M., Kendrick, N. C. E.Taylor, M., Marshall, G. R., Hall, J. E., Vodyanoy, I. & Reusser, F. (1981). Synthesis and characterisation of the major component of alamethicin. J. Amer. Chem. Soc. 103, 61276132.CrossRefGoogle Scholar
Ball, F. G. & Sansom, M. S. P. (1989). Single channel gating mechanisms: model identification and parameter estimation. Proc. Roy. Soc. Land. B. 236, 385416.Google ScholarPubMed
Banerjee, U., Tsui, F.-P.Balasubramanian, T. N., Marshall, G. R. & Chan, S. I. (1983). Structure of alamethicin in solution: one- and two-dimensional 1H NMR studies at 500 MHz. J. Mol. Biol. 165, 757775.CrossRefGoogle Scholar
Banerjee, U., Zidovetski, R., Birge, R. R. & Chan, S. I. (1985). Interaction of alamethicin with lecithin bilayers: A 31P and 2H NMR study. Biochem. 24, 76217627.CrossRefGoogle ScholarPubMed
Barlow, D. J. & Thornton, J. M. (1988). Helix geometry in proteins. J. Mol. Biol. 201, 601619.CrossRefGoogle ScholarPubMed
Baumann, G. & Mueller, P. (1974). A molecular model of membrane excitability. J. Supramol. Struct. 2, 538557.CrossRefGoogle ScholarPubMed
Bazzo, R., Tappin, M. J., Pastore, A., Harvey, T. S., Carver, J. A. & Campbell, I. D. (1988). The structure of melittin: a 1H-NMR study in methanol. Eur. Biochem.J. 173 139146.CrossRefGoogle ScholarPubMed
Beschiaschvili, G. & Seelig, J. (1990). Melittin binding to mixed phophatidyl-glycerol/phosphatidylcholine membranes. Biochem. 29, 5258.CrossRefGoogle ScholarPubMed
Bezrukov, S. M. & Vodyanoy, I. (1993). Probing alamethicin channels with water soluble polymers: effect on conductance of channel states. Biophys. J. 64, 1625.CrossRefGoogle ScholarPubMed
Bodo, B., Rebuffat, S., Hajji, M. E. & Davoust, D. (1985). Structure of trichorzianine A IIIC, an antifungal peptide from Trichoderma harzianum. J. Amer. Chem. Soc. 107, 60116017.CrossRefGoogle Scholar
Bodo, B., Rebuffat, S. & Hajji, M. E. (1988). Helical conformation of trichorzianines in solution. Peptides 1988. Berlin: Walter de Gruyter & Co.Google Scholar
Bocusz, S., Boxer, A. & Busath, D. D. (1992). An SS1-SS2 β-barrel structure for the voltage-activated potassium channel. Prot. Engineer. 5, 285293.CrossRefGoogle Scholar
Boheim, G. (1974). Statistical analysis of alamethicin channels in black lipid membranes. J. Membr. Biol. 19, 277303.CrossRefGoogle Scholar
Boheim, G., Hanke, W. & Jung, G. (1983). Alamethicin pore formation: voltage-dependent flip-flop of α-helix dipoles. Biophys. Struct. Mech. 9, 181191.CrossRefGoogle Scholar
Boheim, G., Gelfert, S., Jung, G. & Menestrina, G. (1987). α-Helical ion channels reconstituted into planar bilayers. In Ion Transport Through Membranes (ed. Yagi, K. and Pullman, B.), pp. 131145. Academic Press.Google Scholar
Breed, J., Kerr, I. D., Sankararamakrishnan, R. & Sansom, M. S. P. (1993). Molecular modelling of alamethicin and related peptaibol channels via simulated annealing. (Manuscript in preparation.)Google Scholar
Brooks, B. R., Bruccoleri, R. E., Olafson, B. D., States, D. J., Swaminathan, S. & Karplus, M. (1983). Charmm: a program for macromolecular energy, minimization, and dynamics calculations. J. Comp. Chem. 4, 187217.CrossRefGoogle Scholar
Brückner, H. & Graf, H. (1983). Paracelsin, a peptide antibiotic containing α-amino isobutyric acid, isolated from Trichoderma reesei Simmons. Experientia 39, 528530.Google Scholar
Brückner, H., Graf, H. & Bokel, M. (1984). Paracelsin: characterization by NMR spectroscopy and circular dichoism, and hemolytic properties of a peptaibol antibiotic from the cellulolytically active mold Trichoderma reesei. Experientia 40, 11891197.Google Scholar
Brünger, A. T., Krukowski, A. & Erickson, J. W. (1990). Slow-cooling protocols for crystallographic refinement by simulated annealing. Acta Cryst. A46: 585593.Google Scholar
Brumfeld, V. & Miller, I. R. (1990). Electric field dependence of alamethicin channels. Biochim. Biophys. Acta 1024, 4953.CrossRefGoogle ScholarPubMed
Burgess, A. W. & Leach, S. J. (1973). An obligatory alpha-helical amino acid residue. Biopolymers 12 25992605.CrossRefGoogle ScholarPubMed
Cafiso, D. S. (1991). Lipid bilayers: membrane-protein electrostatic interactions. Curr. Opin. Struct. Biol. 1, 185190.CrossRefGoogle Scholar
Campbell, I. D. (1988). The structure and dynamics of membrane spanning helices by high resolution NMR and molecular dynamics. In Transport Through Membranes: Carriers, Channels and Pumps (ed. Pullman, B. et al. ), pp. 91101. Dordrecht: Kluwer Academic Publishers.CrossRefGoogle Scholar
Cascio, M. & Wallace, B. A. (1988). Conformation of alamethicin in phospholipids vesicles: implications for insertion models. Proteins: Struc. Func. Genet. 4, 8998.CrossRefGoogle ScholarPubMed
Chen, D. & Eisenberg, R. (1993). Charges, currents, and potentials in ionic channel of one conformation. Biophys. J. 64, 14051421.CrossRefGoogle ScholarPubMed
Davis, M. E., Madura, J. D., Luty, B. A. & McCammon, J. A. (1991). Electrostatics and diffusion of molecules in solution: simulations with the University of Houston Brownian Dynamics program. Comp. Phys. Comm. 62, 182197.CrossRefGoogle Scholar
Dempsey, C. E. (1988). pH dependence of hydrogen exchange from backbone peptide amides of melittin in methanol. Biochem. 27, 68936901.CrossRefGoogle Scholar
Dempsey, C. E. (1990). The actions of melittin on membranes. Biochim. Biophys. Acta. 1031, 143161.CrossRefGoogle ScholarPubMed
Dempsey, C. E. (1992). Quantitation of the effects of an internal proline residue on individual hydrogen bond stabilities in an α-helix: pH-dependent amide exchange in melittin and [ala-14]melittin. Biophys. J. 31, 47054712.Google Scholar
Dempsey, C. E. & Butler, G. S. (1992). Helical structure and orientation of melittin in dispersed phospholipid membranes from amide exchange analysis in situ. Biochem. 31, 1197311977.CrossRefGoogle ScholarPubMed
Dempsey, C. E., Bazzo, R., Harvey, T. S., Syperek, I., Boheim, G. & Campbell, I. D. (1991). Contribution of praline-14 to the structure and actions of melittin. FEBS Lett. 281, 240244.CrossRefGoogle Scholar
Duclohier, H., Molle, G. & Spach, G. (1989). The influence of the trichozianin C-terminal residues on the ion channel conductance in lipid bilayers. Biochim. Biophys. Acta 987, 133136.CrossRefGoogle ScholarPubMed
Duclohier, H., Molle, G., Dugast, J. Y. & Spach, G. (1992). Prolines are not essential residues in the ‘barrel-stave’ model for ion channels induced by alamethicin analogues. Biophys. J. 63, 868873.CrossRefGoogle Scholar
Durrell, S. R. & Guy, H. R. (1992). Atomic scale structure and functional models of voltage-gated potassium channels. Biophys. J. 62, 238250.CrossRefGoogle Scholar
Edholm, O. & Jähnig, F. (1988). The structure of a membrane-spanning polypeptide studied by molecular dynamics. Biophys. Chem. 30, 279292.CrossRefGoogle ScholarPubMed
Edholm, O. & Jähnig, F. (1992). Modeling of the structure of bacteriorhodopsin: a molecular dynamics study. J. Mol. Biol. 226, 837850.Google Scholar
Edmonds, D. T. (1989). A kinetic role for ionizable sites in membrane channel proteins. Eur. Biophys. J. 17, 113119.CrossRefGoogle ScholarPubMed
Esposito, G., Carver, J. A., Boyd, J. & Campbell, I. D. (1987). High resolution 1H NMR study of the solution structure of alamethicin. Biochem. 26, 10431050.CrossRefGoogle ScholarPubMed
Fox, R. O. & Richards, F. M. (1982). A voltage-gated ion channel model inferred from the crystal structure of alamethicin at 1·5 A resolution. Nature 300, 325330.CrossRefGoogle ScholarPubMed
Fraternali, F. (1990). Restrained and unrestrained molecular dynamics simulations in the NVT ensemble of alamethicin. Biopolymers 30, 10831099.CrossRefGoogle ScholarPubMed
Frey, S. & Tamm, L. K. (1991). Orientation of melittin in phospholipid bilayers: a polarized attenuated total reflection infrared study. Biophys. J. 60, 922930.CrossRefGoogle ScholarPubMed
Furois-Corbin, S. & Pullman, A. (1988). Conformation and pairing properties of the N-terminal fragments of trichorzianine and alamethicin: a theoretical study. Biochim. Biophys. Acta 944, 399413.CrossRefGoogle Scholar
Gordon, L. G. M. & Haydon, D. A. (1972). The unit conductance channel of alamethicin. Biochim. Biophys. Acta 255, 10141018.CrossRefGoogle Scholar
Gordon, L. G. M. & Haydon, D. A. (1975). Potential-dependent conductances in lipid membranes containing alamethicin. Phil. Trans. Roy. Soc. Lond. B. 270, 433447.Google ScholarPubMed
Guy, H. R. & Conti, F. (1990). Pursuing the structure and function of voltage-gated channels. Trends Neurosci. 13, 201206.Google ScholarPubMed
Hall, J. E., Vodyanoy, I., Balasubramanian, T. M. & Marshall, G. R. (1984). Alamethicin: a rich model for channel behaviour. Biophys. J. 45, 233247.CrossRefGoogle Scholar
Hanke, W. & Boheim, G. (1980). The lowest conductance state of the alamethicin pore. Biochim. Biophys. Acta 596, 456462.CrossRefGoogle ScholarPubMed
Hanke, W., Methfessel, C., Wilmsen, H. U., Katz, E., Jung, G. & Boheim, G. (1983). Melittin and a chemically modified trichotoxin form alamethicin-type multistate pores. Biochim. Biophys. Acta 727, 108114.CrossRefGoogle Scholar
Haris, P. I. & Chapman, D. (1988). Fourier transform infrared spectra of the polypeptide alamethicin and a possible structural similarity with bacteriorhodopsin. Biochim. Biophys. Acta 943, 375380.CrossRefGoogle Scholar
Harvey, S. C. (1989). Treatment of electrostatic effects in macromolecular modelling. Proteins: Struct. Func. Genet. 5, 7892.CrossRefGoogle Scholar
Hille, B. (1992). Ionic Channels of Excitable Membranes (2nd edn). Sunderland, Mass.: Sinauer Associates Inc.Google Scholar
Hol, W. G. L. (1985). Effects of the α-helix dipole upon the functioning and structures of proteins and peptides. Adv. Biophys. 19, 133165.CrossRefGoogle ScholarPubMed
Hol, W. G., Von Duijen, P. T. & Berendsen, H. J. C. (1978). The α-helix dipole and the properties of proteins. Nature 273, 443446.CrossRefGoogle ScholarPubMed
Hol, W. G. L., Halie, L. N. & Sander, C. (1981). Dipoles of the alpha;-helix and β-sheet: their role in protein folding. Nature 294, 532536.CrossRefGoogle ScholarPubMed
Honig, B. H., Hubbell, W. L. & Flewelling, R. F. (1986). Electrostatic interactions in membranes and proteins. Ann. Rev. Biophys. Chem. 15, 163193.CrossRefGoogle ScholarPubMed
Huang, H. W. & Wu, Y. (1991). Lipid-alamethicin interactions influence alamethicin orientation. Biophys. J. 60, 10791087.CrossRefGoogle ScholarPubMed
Inouye, M. (1974). A three-dimensional molecular assembly model of a lipoprotein from the Escherichia coli outer membrane. Proc. Natl. Acad. Sci. USA 71, 23962400.CrossRefGoogle ScholarPubMed
Jung, G., König, W. A., Leibfritz, D., Ooka, T., Janko, K. & Boheim, G. (1976). Structural and membrane modifying properties of suzukacillin, a peptide antibiotic related to alamethicin. Part A. Sequence and conformation. Biochim. Biophys. Acta 433, 164181.CrossRefGoogle Scholar
Karle, I. L. (1992). Folding, aggregation and molecular recognition in peptides. Acta Cryst. B 48, 341356.CrossRefGoogle ScholarPubMed
Karle, I. L. & Balaram, P. (1990). Structural characteristics of α-helical peptide molecules containing Aib residues. Biochem. 29, 67476756.CrossRefGoogle ScholarPubMed
Karle, I. L., Flippen-Andersen, J., Sukumar, M. & Balaram, P. (1987). Conformation of a 16-residue zervamicin IIA analog peptide containing 3 different structural features: 310-helix, α-helix and β-bend ribbon. Proc. Natl. Acad. Sci. USA 84, 50875091.CrossRefGoogle Scholar
Karle, I. L., Flippen-Anderson, J. L.Sukumar, M. & Balaram, P. (1988). Monoclinic polymorph of Boc-Trp-Ile-Ala-Aib-Ile-Val-Aib-Leu-Pro-OMe (anhydrous). Int. J. Pep. Prot. Res. 31, 567576.CrossRefGoogle ScholarPubMed
Karle, I. L., Flippen-Andersen, J.Agarwalla, S. & Balaram, P. (1991). Crystal structure of Leu-zervamicin, a membrane ion channel peptide. Implications for gating mechanisms. Proc. Natl. Acad. Sci. USA 88, 53075311.CrossRefGoogle Scholar
Karle, I. L., Flippen-Andersen, J., Agarwalla, S. & Balaram, P. (1992). Implications for a ion channel in Leu-zervamicin: crystal structure of polymorph B. In Structure and Function, Vol. 2. Proteins (ed. Sarma, R. H. and Sarma, M. H.) Adenine Press.Google Scholar
Kelsh, L. P., Ellena, J. F. & Cafiso, D. S. (1992). Determination of the molecular dynamics of alamethicin using 13C NMR: implications for the mechanism of gating of a voltage-dependent channel. Biochem. 31, 51365144.CrossRefGoogle ScholarPubMed
Kerr, I. D., Sankararamakrishnan, R. & Sansom, M. S. P. (1993). A simulated annealing approach to molecular modelling of ion channels: Application to δ-toxin. (Manuscript in preparation.)Google Scholar
Kraulis, P. J. (1991). Molscript: a program to produce both detailed and schematic plots of protein structures. J. Appl. Cryst. 24, 946950.CrossRefGoogle Scholar
Krishna, K., Sukumar, M. & Balaram, P. (1990). Structural chemistry and membrane modifying activity of the fungal polypeptides zervamicins, antiamoebins and efrapeptins. Pure Appl. Chem. 62, 14171420.CrossRefGoogle Scholar
Latorre, R. & Alverez, O. (1981). Voltage-dependent channels in planar lipid bilayer membranes. Physiol. Rev. 61, 77150.CrossRefGoogle ScholarPubMed
Läuger, P. (1976). Diffusion-limited ion flow through pores. Biochim. Biophys. Acta 455, 493509.CrossRefGoogle ScholarPubMed
Lear, J. D., Wasserman, Z. R. & DeGrado, W. F. (1988). Synthetic amphiphilic peptide models for protein ion channels. Science 240, 11771181.CrossRefGoogle ScholarPubMed
Le Bars, M., Bachet, B. & Mornon, J. P. (1988). Structure of the helical 19 peptide (trichorzianine AIIIC). Modelling of transmembrane channels. Zeit. Kristallogr. 185, 588 (abstract).Google Scholar
Li, J., Carroll, J. & Ellar, D. J. (1991). Crystal structure of insecticidal δ-endotoxin from Bacillus thuringeinsis at 2·5 Å resolution. Nature 353, 815821.CrossRefGoogle Scholar
Madura, J. F. & McCammon, J. A. (1989). Brownian dynamics simulation of diffusional encounters between triose phosphate isomerase and D-glyceraldehyde phosphate. J. Phys. Chem. 93, 72857287.CrossRefGoogle Scholar
Mak, D. & Webb, W. W. (1993). Two distinct sets of ion channels are formed by neutral alamethicin F-50 polypeptides. Biophys. J. 64, A95.Google Scholar
Marshall, G. D., Hodgkin, E. E., Langs, D. A., Smith, G. D., Zabrocki, J. & Leplawy, M. T. (1990). Factors governing helical preference of peptides containing multiple α,α-dialkyl amino acids. Proc. Natl. Acad. Sci. USA 87, 487491.CrossRefGoogle ScholarPubMed
Mathew, M. K. & Balaram, P. (1983 a). A helix dipole model for alamethicin and related transmembrane channels. FEBS Lett. 157, 15.CrossRefGoogle Scholar
Mathew, M. K. & Balaram, P. (1983 b). Alamethicin and related channel forming polypeptides. Mol. Cell. Biochem. 50, 4764.CrossRefGoogle ScholarPubMed
Matsuzaki, K., Nakai, S., Handa, T., Takaishi, Y., Fujita, T. & Miyajima, K. (1989). Hypelcin A, an α-aminoisobutyric acid containing antibiotic peptide, induced permeability change of phosphatidylcholine bilayers. Biochem 28, 93929398.CrossRefGoogle ScholarPubMed
Mellor, I. R., Thomas, D. H. & Sansom, M. S. P. (1988). Properties of ion channels formed by Staphylococcus aureus δ-toxin. Biochim. Biophys. Acta 942, 280294.Google Scholar
Menestrina, G., Vogues, K. P., Jung, G. & Boheim, G. (1986). Voltage-dependent channel formation by rods of helical peptides. J. Membr. Biol. 93, 111132.CrossRefGoogle Scholar
Miick, S. M., Martinez, G. V., Fiori, W. R., Todd, A. P. & Millhauser, G. I. (1992). Short alanine-based peptides may form 310-helices and not α-helices in aqueous solution. Nature 359, 653655.CrossRefGoogle Scholar
Milik, M. & Skolnick, J. (1993). Insertion of peptide chains into lipid membranes: an off-lattice Monte Carlo dynamics model. Proteins: Struct. Func. Genet. 15, 1025.CrossRefGoogle ScholarPubMed
Molle, G., Dugast, J. Y., Duclohier, H. & Spach, G. (1988). Conductance properties of des-Aib-Leu-des-Pheol-Phe-alamethicin in planar lipid bilayers. Biochim. Biophys. Acta 938, 310314.CrossRefGoogle Scholar
Molle, G., Duclohier, H., Dugast, J. Y., & Spach, G. (1989). Design and conformation of non-Aib synthetic peptides enjoying alamethicin-like ionophore activity. Biopolymers 28, 273283.CrossRefGoogle ScholarPubMed
Molle, G., Duclohier, H., Julien, S. & Spach, G. (1991). Synthetic analogues of alamethicin: effect of C-terminal residue substitutions and chain length on the ion channel lifetimes. Biochim. Biophys. Acta 1064, 365369.CrossRefGoogle ScholarPubMed
Montal, M. O., Iwamoto, T., Tomich, J. M. & Montal, M. (1993). Design, synthesis and functional characterization of a pentameric channel protein that mimics the presumed pore structure of the nicotinic cholinergic receptor. FEBS Lett. 320, 261266.CrossRefGoogle ScholarPubMed
Moore, W. J. (1972). Physical Chemistry (5th edn.). London: Longman.Google Scholar
Nilges, M. & Brünger, A. T. (1991). Automated modeling of coiled coils: application to the GCN4 dimerization region. Prot. Engineer. 4, 649659.CrossRefGoogle Scholar
Nilges, M. & Brünger, A. T. (1993). Successful prediction of the coiled coil geometry of the GCN4 leucine zipper domain by simulated annealing: comparison to the X-ray structure. Proteins: Struct. Func. Genet. 15, 133146.CrossRefGoogle Scholar
Nilges, M., Clore, G. M. & Gronenborn, A. M. (1988). Determination of three-dimensional structures of proteins from interproton distance data by dynamical simulated annealing from a random array of atoms. Circumventing problems associated with folding. FEBS Lett. 239, 129136.CrossRefGoogle ScholarPubMed
Oiki, S., Madison, V. & Montal, M. (1990). Bundles of amphipathic transmembrane α-helices as a structural motif for ion-conducting channel proteins: studies on sodium channels and acetylcholine receptors. Proteins: Struct. Func. Genet. 8, 226236.CrossRefGoogle ScholarPubMed
Opsahl, L. R. & Webb, W. W. (1993). Alamethicin channel conductances substates. Biophys. J. 64, A95.Google Scholar
Otoda, K., Kimura, S. & Imanishi, Y. (1992). Interaction of melittin derivatives with lipid bilayer membrane. Biophys. Biochim. Acta 1112, 16.CrossRefGoogle ScholarPubMed
Padley, R. C., Cook, J. C. & Rinehart, K. L. (1977). High resolution and field desorption mass spectrometry studies and revised structures of alamethicins I and II. J. Amer. Chem. Soc. 99, 84698483.Google Scholar
Parker, M. W., Pattus, F., Tucker, A. D. & Tsernoglou, D. (1989). Structure of the membrane-pore-forming fragment of colicin-A. Nature 337, 9396.CrossRefGoogle ScholarPubMed
Pastore, A., Harvey, T. S., Dempsey, C. E. & Campbell, I. D. (1989). The dynamic properties of melittin in solution – investigation by NMR and molecular dynamics. Eur. Biophys. J. 16, 363367.CrossRefGoogle ScholarPubMed
Polinsky, A., Goodman, M., Williams, K. A. & Deber, C. M. (1992). Minimum energy conformations of praline-containing helices. Biopolymers 32, 399406.CrossRefGoogle Scholar
Popot, J. L. & Engelman, D. M. (1990). Membrane protein folding and oligomerization: the two-state model. Biochem. 29, 40344037.Google Scholar
Rebuffat, S., Hajii, M. E., Hennig, P., Davoust, D. & Bodo, B. (1989). Isolation, sequence and conformation of seven trichorzianines from Trichoderma harzianum. Int.. J. Peptide Protein Res. 34, 200210.CrossRefGoogle Scholar
Rebuffat, S., Prigent, Y., Auvin-Guette, C. & Bodo, B. (1991). Tricholongins BI and BII, 19-residue peptaibols from Trichoderma longibrachiatum: solution structure from two-dimensional NMR spectroscopy. Eur. J. Biochem. 210, 661674.CrossRefGoogle Scholar
Rebuffat, S., Conraux, L., Massias, M., Auvin-Guette, C. & Bodo, B. (1993 a). Sequence and conformation of the 20-residue peptaibols, saturnisporins SA II and SA IV. Int. J. Peptide Protein Res. 41, 7484.CrossRefGoogle Scholar
Rebuffat, S., Duclohier, H., Auvin-Guette, C., Molle, G., Spach, G. & Bodo, B. (1993 b). Membrane-modifying properties of the pore-forming peptaibols saturnisporin SA IV and harzianin HA V. FEMS Microbiol. Immunol. 105, 151160.CrossRefGoogle Scholar
Rebuffat, S., Duclohier, H., Goulard, C., Hlima, S., Auvin-Guette, C., Molle, G., Spach, G. & Bodo, B. (1993 c). Les Harzianines HC, peptides antibiotiques de 14 residus formant des canax voltage-dependants. Cinquième Réunion de Group d' Etudes des Interactions Molécules Membranes, Rouen.Google Scholar
Rinehart, K. L., Gaudioso, L. A., Moore, M. L., Pandey, R. C., Cok, J. C., Brber, M., Sedgwick, D., Bordoli, R. S., Tyler, A. N. & Green, B. N. (1981). Structures of eleven zervamicin and two emerimicin peptide antibiotics studied by fast atom bombardment mass spectroscopy. J. Amer. Chem. Soc. 103, 65176520.CrossRefGoogle Scholar
Rizzo, V., Stankowski, S. & Schwarz, G. (1987). Alamethicin incorporation in lipid bilayers: a thermodynamic analysis. Biochem 26, 27512759.CrossRefGoogle Scholar
Roux, B. & Karplus, M. (1991). Ion transport in a model gramicidin channel: structure and thermodynamics. Biophys. J. 59, 961981.CrossRefGoogle Scholar
Sankararamakrishnan, R. & Vishveshwara, S. (1992). Geometry of praline-containing alpha-helices in proteins. Int. J. Peptide Protein Res. 39, 356363.CrossRefGoogle Scholar
Sankararamakrishnan, R. & Vishveshwara, S. (1993). Characterisation of praline-containing α-helix (helix F model of bacteriorhodopsin) by molecular dynamics studies. Proteins: Struc. Func. Genet. 15, 2641.CrossRefGoogle Scholar
Sansom, M. S. P. (1991). The biophysics of peptide models of ion channels. Prog. Biophys. Mol. Biol. 55, 139236.CrossRefGoogle ScholarPubMed
Sansom, M. S. P. (1992 a). An investigation of the role of serine and threonine sidechains in ion channel proteins. Eur. Biophys. J. 21, 281298.CrossRefGoogle Scholar
Sansom, M. S. P. (1992 b). Proline residues in transmembrane helices of channel and transport proteins: a molecular modelling study. Prot. Engineer. 5, 5360.CrossRefGoogle ScholarPubMed
Sansom, M. S. P. (1993 a). Acetylcholine receptor: peering down a pore. Curr. Biol. 3, 239241.CrossRefGoogle Scholar
Sansom, M. S. P. (1993 b). Alamethicin and related peptaibols – model ion channels. Eur. Biophys. J. 22, 105124.CrossRefGoogle ScholarPubMed
Sansom, M. S. P., Kerr, I. D. & Mellor, I. R. (1991). Ion channels formed by amphipathic helical peptides: a molecular modelling study. Eur. Biophys. J. 20, 229240.CrossRefGoogle ScholarPubMed
Sansom, M. S. P., Balaram, P. & Karle, I. L. (1993). Molecular modelling of ion channels formed by zervamicin-IIB. Eur. Biophys. J. 21, 369383.CrossRefGoogle ScholarPubMed
Sansom, M. S. P. & Kerr, I. D. (1993). Influenza virus M2 protein: a molecular modelling study of the ion channel. Prot. Engineer. 6, 6574.CrossRefGoogle ScholarPubMed
Schwarz, G. (1987). Basic kinetics of binding and incorporation with supramolecular aggregates. Biophys. Chem. 26, 163169.CrossRefGoogle ScholarPubMed
Schwarz, G. & Savko, P. (1982 a). Dielectric probe of the electric-field-sensitive peptide alamethicin. Bioelectromagnetics 3, 2528.CrossRefGoogle ScholarPubMed
Schwarz, G. & Savko, P. (1982 b). Structural and dipolar properties of the voltage-dependent pore former alamethicin in octanol/dioxane. Biophys. J. 39, 211219.CrossRefGoogle ScholarPubMed
Schwarz, G., Savko, P. & Jung, G. (1983). Solvent-dependent structural features of the membrane active peptide trichotoxin A40 as reflected in its dielectric dispersion. Biochim. Biophys. Acta 728, 419428.CrossRefGoogle ScholarPubMed
Schwarz, G., Stankowski, S. & Rizzo, V. (1986). Thermodynamic analysis of incorporation and aggregation in a membrane: application to the pore-forming peptide alamethicin. Biochim. Biophys. Acta 861, 141151.CrossRefGoogle Scholar
Schwarz, G., Gerke, H., Rizzo, V. & Stankowski, S. (1987). Incorporation kinetics in a membrane, studied with the pore-forming peptide alamethicin. Biophys. J. 52, 685692.CrossRefGoogle Scholar
Smith, S. O. & Peerson, O. B. (1992). Solid-state NMR approaches for studying membrane protein structure. Ann. Rev. Biophys. Biomol. Struct. 21, 2547.CrossRefGoogle ScholarPubMed
Stankowski, S. & Schwarz, G. (1989). Lipid dependence of peptide-membrane interactions: bilayer affinity and aggregation of the peptide alamethicin. FEBS Lett. 250, 556560.CrossRefGoogle ScholarPubMed
Stankowski, S. & Schwarz, U. D. & Schwarz, G. (1988). Voltage-dependent pore activity of the peptide alamethicin correlated with incorporation in the membrane: salt and cholesterol effects. Biochim. Biophys, Acta 941, 1118.CrossRefGoogle ScholarPubMed
Stein, W. D. (1990). Channels, Carriers and Pumps. San Diego: Academic Press.Google Scholar
Stroud, R. M., McCarthy, M. P. & Shuster, M. (1990). Nicotinic acetylcholine receptor superfamily of ligand gated ion channels. Biochem. 50, 1100911023.CrossRefGoogle Scholar
Terwilliger, T. C. & Eisenberg, D. (1982). The structure of melittin: II – interpretation of the structure. J. Biol. Chem. 257, 60166022.CrossRefGoogle ScholarPubMed
Terwilliger, T. C., Weissman, L. & Eisenberg, D. (1982). The structure of melittin in the form I crystals and its implications for melittin's lytic and surface activity. Biophys. J. 37, 353361.CrossRefGoogle Scholar
Toniolo, C. & Benedotti, E. (1991). The polypeptide 310-helix. Trends Biochem. Sci. 16, 350353.CrossRefGoogle ScholarPubMed
Tosteson, M. T. & Tosteson, D. C. (1981). The sting: melittin forms channels in lipid bilayers. Biophys. J. 36, 109116.CrossRefGoogle ScholarPubMed
Tosteson, M. T., Alvarez, O. & Tosteson, D. C. (1985). Peptides as promoters of ion-permeable channels. Regul. Peptides (Suppl. 4) 8, 3945.CrossRefGoogle Scholar
Tosteson, M. T., Levy, J. J., Caporale, L. H., Rosenblatt, M. & Tosteson, D. C. (1987). Solid phase synthesis of melittin: purification and functional characterization. Biochem. 26, 66276631.CrossRefGoogle ScholarPubMed
Tosteson, M. T., Alvarez, O., Hubbell, W., Bieganski, R. M., Attenbach, C., Caporales, L. H., Levy, J. J., Nutt, R. F., Rosenblatt, M. & Tosteson, D. C. (1990). Primary structure of peptides and ion channels – role of amino acid side chains in voltage gating of melittin channels. Biophys. J. 58, 13671375.CrossRefGoogle ScholarPubMed
Toyoshima, C. & Unwin, N. (1988). Ion channel of acetylcholine receptor reconstructed from images of post-synaptic membranes. Nature 336, 247250.CrossRefGoogle Scholar
Treutlein, H. R., Lemmon, M. A., Engelmann, D. M. & Brunger, A. T. (1992). The glycophorin A transmembrane domain dimer: sequence-specific propensity for a right-handed supercoil of helices. Biochem. 31, 1272612733.CrossRefGoogle ScholarPubMed
Unwin, N. (1989). The structure of ion channels in membranes of excitable cells. Neuron 3, 665676.CrossRefGoogle ScholarPubMed
Unwin, N. (1993). Nicotinic acetylcholine receptor at 9 Å resolution. J. Mol. Biol. 229, 11011124.CrossRefGoogle ScholarPubMed
Vishveshwara, S. S. & Vishveshwara, S. (1993). Effect of constraints by threonine on proline containing α-helix – a molecular dynamics approach. Biophys. Chem. 46, 7789.CrossRefGoogle Scholar
Vodyanoy, I., Hall, J. E. & Balasubramanian, T. M. (1983). Alamethicin-induced current-voltage curve asymmetry in lipid bilayers. Biophys. J. 42, 7182.CrossRefGoogle ScholarPubMed
Vogel, H. (1987). Comparison of the conformation and orientation of alamethicin and melittin in lipid membranes. Biochem. 26, 45624572.CrossRefGoogle ScholarPubMed
Vogel, H. (1992). Structure and dynamics of polypeptides and proteins in lipid membranes. Quart. Rev. Biophys. 25, 433458.CrossRefGoogle ScholarPubMed
Vogel, H., Nilsson, L., Rigler, R., Meder, S., Boheim, G., Beck, W., Kurth, H. H. & Jung, G. (1993). Structural fluctuations between two onformational states of a transmembrane helical peptide are related to its channel-forming properties in planar lipid membranes. Eur. Biochem. J. 272, 305313.CrossRefGoogle Scholar
Wada, A. (1976). The α-helix as an electric macro-dipole. Adv. Biophys. 9, 163.Google Scholar
Woolley, G. A. & Wallace, B. A. (1992). Model ion channels: gramicidin and alamethicin. J. Membr. Biol. 129, 109136.Google ScholarPubMed
Woolley, G. A. & Wallace, B. A. (1993). Temperature-dependence of the interaction of alamethicin helices in membranes. Biochem. (in press.)CrossRefGoogle Scholar
Wu, Y., Huang, H. W. & Olah, G. A. (1993). Method of oriented circular dichroism. Biophys. J. 57, 797806.CrossRefGoogle Scholar
Yee, A. A. & O'Neil, J. D. J. (1992). Uniform 15N labeling of a fungal peptide: the structure and dynamics of an alamethicin by 15N and 1H NMR spectroscopy. Biochem. 31, 31353143.CrossRefGoogle Scholar
Yun, R. H., Anderson, A. & Hermans, J. (1992). Proline in α-helix: stability and conformation studied by dynamics simulation. Proteins: Struct. Func. Genet. 10, 219228.CrossRefGoogle Scholar