Hostname: page-component-848d4c4894-nmvwc Total loading time: 0 Render date: 2024-06-16T23:59:17.815Z Has data issue: false hasContentIssue false

Interactions of surfactants with lipid membranes

Published online by Cambridge University Press:  11 December 2008

Heiko Heerklotz*
Affiliation:
Leslie Dan Faculty of Pharmacy, University of Toronto, Toronto, ON, CanadaM5S 3M2
*
*H. Heerklotz, Leslie Dan Faculty of Pharmacy, University of Toronto, 144 College St., Toronto, ON, CanadaM5S 3M2. Tel.: +1-416-978-1188; Fax: +1-416-978-8511; Email: heiko.heerklotz@utoronto.ca

Abstract

Surfactants are surface-active, amphiphilic compounds that are water-soluble in the micro- to millimolar range, and self-assemble to form micelles or other aggregates above a critical concentration. This definition comprises synthetic detergents as well as amphiphilic peptides and lipopeptides, bile salts and many other compounds. This paper reviews the biophysics of the interactions of surfactants with membranes of insoluble, naturally occurring lipids. It discusses structural, thermodynamic and kinetic aspects of membrane–water partitioning, changes in membrane properties induced by surfactants, membrane solubilisation to micelles and other phases formed by lipid–surfactant systems. Each section defines and derives key parameters, mentions experimental methods for their measurement and compiles and discusses published data. Additionally, a brief overview is given of surfactant-like effects in biological systems, technical applications of surfactants that involve membrane interactions, and surfactant-based protocols to study biological membranes.

Type
Review Article
Copyright
Copyright © 2008 Cambridge University Press

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

Ahyayauch, H., Larijani, B., Alonso, A. & Goni, F. M. (2006). Detergent solubilization of phosphatidylcholine bilayers in the fluid state: influence of the acyl chain structure. Biochimica et Biophysica Acta – Biomembranes 1758, 190196.CrossRefGoogle ScholarPubMed
Ahyayauch, H., Requero, M. A., Alonso, A., Bennouna, M. & Goni, F. M. (2002). Surfactant effects of chlorpromazine and imipramine on lipid bilayers containing sphingomyelin and cholesterol. Journal of Colloid and Interface Science 256, 284289.CrossRefGoogle ScholarPubMed
Almgren, M. (2000). Mixed micelles and other structures in the solubilization of bilayer lipid membranes by surfactants. Biochimica et Biophysica Acta 1508, 146163.CrossRefGoogle ScholarPubMed
Almgren, M. (2007). Vesicle Transformations resulting from curvature tuning in systems with micellar, lamellar, and bicontinuous cubic phases. Journal of Dispersion Science and Technology 28, 4354.CrossRefGoogle Scholar
Almog, S., Kushnir, T., Nir, S. & Lichtenberg, D. (1986). Kinetic and structural aspects of reconstitution of phosphatidylcholine vesicles by dilution of phosphatidylcholine–sodium cholate mixed micelles. Biochemistry 25, 25972605.CrossRefGoogle ScholarPubMed
Almog, S. & Lichtenberg, D. (1988). Effect of calcium on kinetic and structural aspects of dilution-induced micellar to lamellar phase transformation in phosphatidylcholine–cholate mixtures. Biochemistry 27, 873880.CrossRefGoogle ScholarPubMed
Alonso, A. & Goni, F. M. (2000). Detergents in biomembrane studies. Biochimica et Biophysica Acta 1508, 1256.Google Scholar
Alonso, A., Saez, R. & Goni, F. M. (1982). The interaction of detergents with phospholipid vesicles: a spectrofluorimetric study. FEBS Letters 137, 141145.CrossRefGoogle ScholarPubMed
Alonso, A., Urbaneja, M. A., Goni, F. M., Carmona, F. G., Cánovas, F. G. & Gómez-Fernández, J. C. (1987). Kinetic studies on the interaction of phosphatidylcholine liposomes with Triton X-100. Biochimica et Biophysica Acta 902, 237246.CrossRefGoogle ScholarPubMed
Alonso, A., Villena, A. & Goni, F. M. (1981). Lysis and reassembly of sonicated lecithin vesicles in the presence of Triton X-100. FEBS Letters 123, 200204.CrossRefGoogle ScholarPubMed
Alqawi, O. & Georges, E. (2003). The multidrug resistance protein ABCC1 drug-binding domains show selective sensitivity to mild detergents. Biochemical and Biophysical Research Communications 303, 11351141.CrossRefGoogle ScholarPubMed
Alves, I. D., Salgado, G. F. J., Salamon, Z., Brown, M. F., Tollin, G. & Hruby, V. J. (2005). Phosphatidylethanolamine enhances rhodopsin photoactivation and transducin binding in a solid supported lipid bilayer as determined using plasmon-waveguide resonance spectroscopy. Biophysical Journal 88, 198210.CrossRefGoogle Scholar
Andelman, D., Kozlov, M. M. & Helfrich, W. (1994). Phase-transitions between vesicles and micelles driven by competing curvatures. Europhysics Letters 25, 231236.CrossRefGoogle Scholar
Angelov, B., Ollivon, M. & Angelova, A. (1999). X-ray diffraction study of the effect of the detergent octyl glucoside on the structure of lamellar and nonlamellar lipid/water phases of use for membrane protein reconstitution. Langmuir 15, 82258234.CrossRefGoogle Scholar
Apel-Paz, M., Doncel, G. F. & Vanderlick, T. K. (2005). Impact of membrane cholesterol content on the resistance of vesicles to surfactant attack. Langmuir 21, 98439849.CrossRefGoogle ScholarPubMed
Arnulphi, C., Sot, J., García-Pacios, M., Arrondo, J.-L. R., Alonso, A. & Goñi, F. M. (2007). Triton X-100 partitioning into sphingomyelin bilayers at subsolubilizing detergent concentrations. Effect of lipid phase and a comparison with dipalmitoylphosphatidylcholine. Biophysical Journal 93, 35043514.CrossRefGoogle Scholar
Baker, B. M. & Murphy, K. P. (1998). Prediction of binding energetics from structure using empirical parametrization. Methods Enzymology 295, 294314.CrossRefGoogle Scholar
Bakht, O., Pathak, P. & London, E. (2007). Effect of the structure of lipids favoring disordered domain formation on the stability of cholesterol-containing ordered domains (lipid rafts): identification of multiple raft-stabilization mechanisms. Biophysical Journal 93, 43074318.CrossRefGoogle ScholarPubMed
Barriocanal, L., Taylor, K. M. G. & Buckton, G. (2005). Bilayer to micelle transition of DMPC and alcohol ethoxylate surfactants as studied by isoperibol calorimetry. Journal of Pharmaceutical Sciences 94, 17471755.CrossRefGoogle ScholarPubMed
Bechinger, B. & Lohner, K. (2006). Detergent-like actions of linear amphipathic cationic antimicrobial peptides. Biochimica et Biophysica Acta – Biomembranes 1758, 15291539.CrossRefGoogle ScholarPubMed
Beck, A., Tsamaloukas, A., Jurcevic, P. & Heerklotz, H. (2008). Additive action of two or more solutes on lipid membranes. Langmuir 24, 88338840.CrossRefGoogle ScholarPubMed
Beschiaschvili, G. & Seelig, J. (1992). Peptide binding to lipid bilayers. Nonclassical hydrophobic effect and membrane-induced pK shifts. Biochemistry 31, 1004410053.CrossRefGoogle ScholarPubMed
Beven, L. & Wroblewski, H. (1997). Effect of natural amphipathic peptides on viability, membrane potential, cell shape and motility of mollicutes. Research in Microbiology 148, 163175.CrossRefGoogle ScholarPubMed
Bhakoo, M. & McElhaney, R. N. (1988). The effect of variations in growth temperature, fatty acid composition and cholesterol content on the lipid polar head-group composition of Acholeplasma laidlawii B membranes. Biochimica et Biophysica Acta 945, 307314.CrossRefGoogle ScholarPubMed
Bhamidipati, S. P. & Hamilton, J. A. (1995). Interactions of lyso 1-palmitoylphosphatidylcholine with phospholipids: a 13C and 31P NMR study. Biochemistry 34, 56665677.CrossRefGoogle Scholar
Binder, H. & Klose, G. (2002). Lyotropic phase behavior and structure of mixed lipid (POPC)–detergent (C12En, n=2, 4, 6) assemblies: insights from hydration-tuning infrared spectroscopy. Journal of Physical Chemistry B 106, 1099111001.CrossRefGoogle Scholar
Binford, J. S. Jr., & Palm, W. H. (1994). Absorption of surfactants by membranes: erythrocytes versus synthetic vesicles. Biophysical Journal 66, 20242028.CrossRefGoogle ScholarPubMed
Botelho, A. V., Huber, T., Sakmar, T. P. & Brown, M. F. (2006). Curvature and hydrophobic forces drive oligomerization and modulate activity of rhodopsin in membranes. Biophysical Journal 91, 44644477.CrossRefGoogle ScholarPubMed
Cantor, R. S. (1999). The influence of membrane lateral pressures on simple geometric models of protein conformational equilibria. Chemistry and Physics of Lipids 101, 4556.CrossRefGoogle ScholarPubMed
Carion-Taravella, B., Lesieur, S., Chopineau, J., Lesieur, P. & Ollivon, M. (2002). Phase behavior of mixed aqueous dispersions of dipalmitoylphosphatidylcholine and dodecyl glycosides: a differential scanning calorimetry and X-ray diffraction investigation. Langmuir 18, 325335.CrossRefGoogle Scholar
Castillo, J. A., Pinazo, A., Carilla, J., Infante, M. R., Alsina, M. A., Haro, I. & Clapés, P. (2004). Interaction of antimicrobial arginine-based cationic surfactants with liposomes and lipid monolayers. Langmuir 20, 33793387.CrossRefGoogle ScholarPubMed
Ceu Rei, M., Coutinho, P. J. G., Castanheira, E. M. S. & Real Oliveira, M. E. C. D. (2004). C12E7–DPPC mixed systems studied by pyrene fluorescence emission. Progress in Colloid and Polymer Science 123, 8387.Google Scholar
Cevc, G. (2004). Lipid vesicles and other colloids as drug carriers on the skin. Advanced Drug Delivery Reviews 56, 675711.CrossRefGoogle ScholarPubMed
Cevc, G. & Marsh, D. (1985). Phospholipid Bilayers. John Wiley and Sons, New York.Google Scholar
Chernomordik, L. V., Leikina, E., Frolov, V., Bronk, P. & Zimmerberg, J. (1997). An early stage of membrane fusion mediated by the low pH conformation of influenza hemagglutinin depends upon membrane lipids. Journal of Cell Biology 136, 8193.CrossRefGoogle Scholar
Chernomordik, L. V., Vogel, S. S., Sokoloff, A., Onaran, H. O., Leikina, E. A. & Zimmerberg, J. (1993). Lysolipids reversibly inhibit Ca(2+)-, GTP- and pH-dependent fusion of biological membranes. FEBS Letters 318, 7176.CrossRefGoogle ScholarPubMed
Chopineau, J., Lesieur, S., Carion-Taravella, B. & Ollivon, M. (1998). Self-evolving microstructured systems upon enzymatic catalysis. Biochimie 80, 421435.CrossRefGoogle ScholarPubMed
Cocera, M., Lopez, O., Coderch, L., Parra, J. L. & de la Maza, A. (1999). Influence of the level of ceramides on the permeability of stratum corneum lipid liposomes caused by a C12-betaine/sodium dodecyl sulfate mixture. International Journal of Pharmaceutics 183, 165173.CrossRefGoogle ScholarPubMed
Cocera, M., Lopez, O., Coderch, L., Parra, J. L. & De La Maza, A. (2003). Permeability investigations of phospholipid liposomes by adding cholesterol. Colloids and Surfaces A: Physicochemical and Engineering Aspects 221, 917.CrossRefGoogle Scholar
Cocera, M., Lopez, O., Estelrich, J., Parra, J. L. & De la Maza, A. (2002). Adsorption of sodium lauryl ether sulfate on liposomes by means of a fluorescent probe: effect of the ethylene oxide groups. Langmuir 18, 82508254.CrossRefGoogle Scholar
Cocera, M., Lopez, O., Pons, R., Amenitsch, H. & De La Maza, A. (2004). Effect of the electrostatic charge on the mechanism inducing liposome solubilization: a kinetic study by synchrotron radiation SAXS. Langmuir 20, 30743079.CrossRefGoogle ScholarPubMed
Corless, J. M., McCaslin, D. R. & Scott, B. L. (1982). Two-dimensional rhodopsin crystals from disk membranes of frog retinal rod outer segments. Proceedings of the National Academy of Sciences USA 79, 11161120.CrossRefGoogle ScholarPubMed
Cullis, P. R. & de Kruijff, B. (1979). Lipid polymorphism and the functional roles of lipids in biological membranes. Biochimica et Biophysica Acta 559, 399420.CrossRefGoogle ScholarPubMed
Cullis, P. R., Hope, M. J., Bally, M. B., Madden, T. D., Mayer, L. D. & Fenske, D. B. (1997). Influence of pH gradients on the transbilayer transport of drugs, lipids, peptides and metal ions into large unilamellar vesicles. Biochimica et Biophysica Acta – Reviews on Biomembranes 1331, 187211.CrossRefGoogle ScholarPubMed
da Graca-Miguel, M., Eidelman, O., Ollivon, M. & Walter, A. (1989). Temperature dependence of the vesicle–micelle transition of egg phosphatidylcholine and octyl glucoside. Biochemistry 28, 89218928.CrossRefGoogle ScholarPubMed
Davies, S. M., Epand, R. M., Kraayenhof, R. & Cornell, R. B. (2001). Regulation of CTP: phosphocholine cytidylyltransferase activity by the physical properties of lipid membranes: an important role for stored curvature strain energy. Biochemistry 40, 1052210531.CrossRefGoogle ScholarPubMed
de la Maza, A., Coderch, L., Gonzalez, P. & Parra, J. L. (1998a). Subsolubilizing alterations caused by alkyl glucosides in phosphatidylcholine liposomes. Journal of Controlled Release 52, 159168.CrossRefGoogle ScholarPubMed
de la Maza, A., Lopez, O., Coderch, L. & Parra, J. L. (1998b). Solubilization of phosphatidylcholine liposomes by the amphoteric surfactant dodecyl betaine. Chemistry and Physics of Lipids 94, 7179.CrossRefGoogle ScholarPubMed
de la Maza, A. & Parra, J. L. (1994a). Structural phase transitions involved in the interaction of phospholipid bilayers with octyl glucoside. European Journal of Biochemistry 226, 10291038.CrossRefGoogle ScholarPubMed
de la Maza, A. & Parra, J. L. (1994b). Vesicle–micelle structural transition of phosphatidylcholine bilayers and Triton X-100. Biochemical Journal 303, 907914.CrossRefGoogle ScholarPubMed
de la Maza, A. & Parra, J. L. (1997). Solubilizing effects caused by the nonionic surfactant dodecylmaltoside in phosphatidylcholine liposomes. Biophysical Journal 72, 16681675.CrossRefGoogle ScholarPubMed
Demina, T., Grozdova, I., Krylova, O., Zhirnov, A., Istratov, V., Frey, H., Kautz, H. & Melik-Nubarov, N. (2005). Relationship between the structure of amphiphilic copolymers and their ability to disturb lipid bilayers. Biochemistry 44, 40424054.CrossRefGoogle ScholarPubMed
Duzgunes, N., Straubinger, R. M., Baldwin, P. A., Friend, D. S. & Papahadjopoulos, D. (1985). Proton-induced fusion of oleic acid–phosphatidylethanolamine liposomes. Biochemistry 24, 30913098.CrossRefGoogle ScholarPubMed
Edidin, M. (2003). The state of lipid rafts: from model membranes to cells. Annual Review of Biophysics and Biomolecular Structures 16, 257283.CrossRefGoogle Scholar
Edwards, K. & Almgren, M. (1990). Kinetics of surfactant-induced leakage and growth of unilamellar vesicles. Progress in Colloid and Polymer Science 82, 190197.CrossRefGoogle Scholar
Edwards, K. & Almgren, M. (1991). Solubilization of lecithin vesicles by C12E8 – structural transitions and temperature effects. Journal of Colloid and Interface Science 147, 121.CrossRefGoogle Scholar
Edwards, K., Almgren, M., Bellare, J. & Brown, W. (1989). Effects of Triton X-100 on sonicated lecithin vesicles. Langmuir 5, 473478.CrossRefGoogle Scholar
Elamrani, K. & Blume, A. (1982). Incorporation kinetics of lysolecithin into lecithin vesicles. Kinetics of lysolecithin-induced vesicle fusion. Biochemistry 21, 521526.CrossRefGoogle ScholarPubMed
Encinas, M. & Lissi, E. (1982). Evaluation of partition constants in compartmentalized systems from fluorescence quenching data. Chemical Physics Letters 91, 5557.CrossRefGoogle Scholar
Epand, R. M. (1998a). Lipid polymorphism and membrane properties. In Current Topics in Membranes, vol. 44. San Diego, CA: Academic Press.Google Scholar
Epand, R. M. (1998b). Lipid polymorphism and protein–lipid interactions. Biochimica et Biophysica Acta – Reviews on Biomembranes 1376, 353368.CrossRefGoogle ScholarPubMed
Epand, R. M. & Epand, R. F. (1994). Calorimetric detection of curvature strain in phospholipid bilayers. Biophysical Journal 66, 14501456.CrossRefGoogle ScholarPubMed
Epand, R. M., Epand, R. F. & Lancaster, C. R. (1988). Modulation of the bilayer to hexagonal phase transition of phosphatidylethanolamines by acylglycerols. Biochimica et Biophysica Acta 945, 161166.CrossRefGoogle ScholarPubMed
Farge, E. (1994). Scale-dependent elastic response of closed phospholipid-bilayers to transmembrane molecular pumping activity – a key for exo-endocytosis physiological process. Nuovo Cimento Della Societa Italiana Di Fisica D-Condensed Matter Atomic Molecular and Chemical Physics Fluids Plasmas Biophysics 16, 14571470.Google Scholar
Francis, G., Kerem, Z., Makkar, H. P. S. & Becker, K. (2002). The biological action of saponins in animal systems: a review. British Journal of Nutrition 88, 587605.CrossRefGoogle ScholarPubMed
Fuller, N. & Rand, P. (2001). The influence of lysolipids on the intrinsic curvature and bending elasticity of phospholipid membranes. Biophysical Journal 80, 2357.Google Scholar
Funari, S. S., Madler, B. & Rapp, G. (1996). Cubic topology in surfactant and lipid mixtures. European Biophysics Journal 24, 293299.CrossRefGoogle Scholar
Funari, S. S., Nuscher, B., Rapp, G. & Beyer, K. (2001). Detergent–phospholipid mixed micelles with a crystalline phospholipid core. Proceedings of the National Academy of Sciences USA 98, 89388943.CrossRefGoogle ScholarPubMed
Galla, H. J. & Sackmann, E. (1974). Lateral diffusion in the hydrophobic region of membranes: use of pyrene excimers as optical probes. Biochimica et Biophysica Acta 339, 103115.CrossRefGoogle ScholarPubMed
Gallova, J., Devinsky, F. & Balgavy, P. (1990). Interaction of surfactants with model and biological membranes. II. Effect of N-alkyl-N,N,N-trimethylammonium ions on phosphatidylcholine bilayers as studied by spin probe ESR. Chemistry and Physics of Lipids 53, 231241.Google Scholar
Garidel, P., Hildebrand, A., Knauf, K. & Blume, A. (2007). Membranolytic activity of bile salts: influence of biological membrane properties and composition. Molecules 12, 22922326.CrossRefGoogle ScholarPubMed
Garner, A. E., Smith, D. A. & Hooper, N. M. (2008). Visualization of detergent solubilization of membranes: implications for the isolation of rafts. Biophysical Journal 94, 13261340.CrossRefGoogle ScholarPubMed
Giorgione, J., Epand, R. M., Buda, C. & Farkas, T. (1995). Role of phospholipids containing docosahexaenoyl chains in modulating the activity of protein kinase C. Proceedings of the National Academy of Sciences USA 92, 97679770.CrossRefGoogle ScholarPubMed
Glover, K. J., Whiles, J. A., Wu, G., Yu, N., Deems, R., Struppe, J. O., Stark, R. E., Komives, E. A. & Vold, R. R. (2001). Structural evaluation of phospholipid bicelles for solution-state studies of membrane-associated biomolecules. Biophysical Journal 81, 21632171.CrossRefGoogle ScholarPubMed
Gohon, Y. & Popot, J. L. (2003). Membrane protein–surfactant complexes. Current Opinion in Colloid and Interface Science 8, 1522.CrossRefGoogle Scholar
Goni, F. M., Urbaneja, M. A., Arrondo, J. L., Alonso, A., Durrani, A. A. & Chapman, D. (1986). The interaction of phosphatidylcholine bilayers with triton X-100. European Journal of Biochemistry 160, 659665.CrossRefGoogle ScholarPubMed
Gonzales-Manas, J. M., Kaschny, P. & Goni, F. M. (1994). Use of Merocyanine 540 as an optical probe in the study of membrane–surfactant interactions. Journal of Physical Chemistry 98, 1065010654.CrossRefGoogle Scholar
Gradzielski, M. (2004). Investigations of the dynamics of morphological transitions in amphiphilic systems. Current Opinion in Colloid and Interface Science 9, 256263.CrossRefGoogle Scholar
Gregory, S. M., Cavenaugh, A., Journigan, V., Pokorny, A. & Almeida, P. F. (2008). A quantitative model for the all-or-none permeabilization of phospholipid vesicles by the antimicrobial peptide cecropin A. Biophysical Journal 94, 16671680.CrossRefGoogle ScholarPubMed
Grisshammer, R., White, J. F., Trinh, L. B. & Shiloach, J. (2005). Large-scale expression and purification of a G-protein-coupled receptor for structure determination – an overview. Journal of Structural and Functional Genomics 6, 159163.CrossRefGoogle ScholarPubMed
Gruner, S. M. (1985). Intrinsic curvature hypothesis for biomembrane lipid composition: a role for nonbilayer lipids. Proceedings of the National Academy of Sciences USA 82, 36653669.CrossRefGoogle ScholarPubMed
Guggenheim, E. A. (1952). Mixtures. The Theory of the Equilibrium Properties of Some Simple Classes of Mixtures, Solutions, and Alloys. Oxford: Clarendon Press.Google Scholar
Gustafsson, J., Oradd, G. & Almgren, M. (1997). Disintegration of the lecithin lamellar phase by cationic surfactants. Langmuir 13, 69566963.CrossRefGoogle Scholar
Gutberlet, T., Dietrich, U., Klose, G. & Rapp, G. (1998). X-ray diffraction study of the lamellar–hexagonal phase transition in phospholipid/surfactant mixtures. Journal of Colloid and Interface Science 203, 317327.CrossRefGoogle ScholarPubMed
Gutberlet, T., Kiselev, M., Heerklotz, H. & Klose, G. (2000). SANS study of mixed POPC/C12En aggregates. Physica B 276, 381383.CrossRefGoogle Scholar
Hagerstrand, H., Holmstrom, T. H., Bobrowska-Hagerstrand, M., Eriksson, J. E. & Isomaa, B. (1998). Amphiphile-induced phosphatidylserine exposure in human erythrocytes. Molecular Membrane Biology 15, 8995.CrossRefGoogle ScholarPubMed
Hagerstrand, H., Kralj-Iglic, V., Fosnaric, M., Bobrowska-Hagerstrand, M., Wrobel, A., Mrowczynska, L., Soderstrom, T. & Iglic, A. (2004). Endovesicle formation and membrane perturbation induced by polyoxyethyleneglycolalkylethers in human erythrocytes. Biochimica et Biophysica Acta – Biomembranes 1665, 191200.CrossRefGoogle ScholarPubMed
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. Biophysical Journal 84, 30523060.CrossRefGoogle ScholarPubMed
Heerklotz, H. (unpublished data).Google Scholar
Heerklotz, H. (2001). Membrane stress and permeabilization induced by asymmetric incorporation of compounds. Biophysical Journal 81, 184195.CrossRefGoogle ScholarPubMed
Heerklotz, H. (2002). Triton promotes domain formation in lipid raft mixtures. Biophysical Journal 83, 26932701.CrossRefGoogle ScholarPubMed
Heerklotz, H., Binder, H. & Lantzsch, G. (1994a). Determination of the partition coefficients of the nonionic detergent C12E7 between lipid-detergent mixed membranes and water by means of laurdan fluorescence spectroscopy. Journal of Fluorescence 4, 349352.CrossRefGoogle ScholarPubMed
Heerklotz, H., Binder, H., Lantzsch, G. & Klose, G. (1994b). Membrane/water partition of oligo(ethylene oxide) dodecyl ethers and its relevance for solubilization. Biochimica et Biophysica Acta 1196, 114122.CrossRefGoogle ScholarPubMed
Heerklotz, H., Binder, H., Lantzsch, G., Klose, G. & Blume, A. (1997). Lipid/detergent interaction thermodynamics as a function of molecular shape. Journal of Physical Chemistry B 101, 639645.CrossRefGoogle Scholar
Heerklotz, H. & Epand, R. M. (2001). The enthalpy of acyl chain packing and the apparent water-accessible apolar surface area of phospholipids. Biophysical Journal 80, 271279.CrossRefGoogle ScholarPubMed
Heerklotz, H., Lantzsch, G., Binder, H., Klose, G. & Blume, A. (1995). Application of isothermal titration calorimetry for detecting lipid membrane solubilization. Chemical Physics Letters 235, 517520.CrossRefGoogle Scholar
Heerklotz, H., Lantzsch, G., Binder, H., Klose, G. & Blume, A. (1996). Thermodynamic characterization of dilute aqueous lipid/detergent mixtures of POPC and C12EO8 by means of isothermal titration calorimetry. Journal of Physical Chemistry 100, 67646774.CrossRefGoogle Scholar
Heerklotz, H. & Seelig, J. (2000a). Correlation of the membrane/water partition coefficients of detergents with the critical micelle concentration. Biophysical Journal 78, 24352440.CrossRefGoogle ScholarPubMed
Heerklotz, H. & Seelig, J. (2000b). Titration calorimetry of surfactant–membrane partitioning and membrane solubilization. Biochimica et Biophysica Acta 1508, 6985.CrossRefGoogle ScholarPubMed
Heerklotz, H. & Seelig, J. (2001). Detergent-like action of the antibiotic peptide surfactin on lipid membranes. Biophysical Journal 81, 15471554.CrossRefGoogle ScholarPubMed
Heerklotz, H. & Seelig, J. (2007). Leakage and lysis of lipid membranes induced by the lipopeptide surfactin. European Biophysics Journal 36, 305314.CrossRefGoogle ScholarPubMed
Heerklotz, H., Szadkowska, H., Anderson, T. & Seelig, J. (2003). The sensitivity of lipid domains to small perturbations demonstrated by the effect of triton. Journal of Molecular Biology 329, 793799.CrossRefGoogle ScholarPubMed
Heerklotz, H., Tsamaloukas, A., Kita-Tokarczyk, K., Strunz, P. & Gutberlet, T. (2004a). Structural, volumetric, and thermodynamic characterization of a micellar sphere-to-rod transition. Journal of the American Chemical Society 126, 1654416552.CrossRefGoogle ScholarPubMed
Heerklotz, H., Wieprecht, T. & Seelig, J. (2004b). Membrane perturbation by the lipopeptide surfactin and detergents as studied by deuterium NMR. Journal of Physical Chemistry B 108, 49094915.CrossRefGoogle Scholar
Heerklotz, H. H., Binder, H. & Epand, R. M. (1999). A ‘release’ protocol for isothermal titration calorimetry. Biophysical Journal 76, 26062613.CrossRefGoogle ScholarPubMed
Heerklotz, H. H., Binder, H. & Schmiedel, H. (1998). Excess enthalpies of mixing in phospholipid-additive membranes. Journal of Physical Chemistry B 102, 53635368.CrossRefGoogle Scholar
Helenius, A. & Simons, K. (1975). Solubilization of membranes by detergents. Biochimica et Biophysica Acta 415, 2979.CrossRefGoogle ScholarPubMed
Helfrich, W. (1973). Elastic properties of lipid bilayers – theory and possible experiments. Zeitschrift fur Naturforschung C 28, 693703.CrossRefGoogle ScholarPubMed
Hildebrand, A., Beyer, K., Neubert, R., Garidel, P. & Blume, A. (2003). Temperature dependence of the interaction of cholate and deoxycholate with fluid model membranes and their solubilization into mixed micelles. Colloids and Surfaces B: Biointerfaces 32, 335351.CrossRefGoogle Scholar
Hildebrand, A., Beyer, K., Neubert, R., Garidel, P. & Blume, A. (2004). Solubilization of negatively charged DPPC/DPPG liposomes by bile salts. Journal of Colloid and Interface Science 279, 559571.CrossRefGoogle ScholarPubMed
Hildebrand, A., Neubert, R., Garidel, P. & Blume, A. (2002). Bile salt induced solubilization of synthetic phosphatidylcholine vesicles studied by isothermal titration calorimetry. Langmuir 18, 28362847.CrossRefGoogle Scholar
Hildebrand, J. H. (1929). Solubility. XII. Regular solutions. Journal of the American Chemical Society 51, 66.CrossRefGoogle Scholar
Hjelm, R. P., Thiyagarajan, P. & Alkanonyuksel, H. (1992). Organization of phosphatidylcholine and bile-salt in rodlike mixed micelles. Journal of Physical Chemistry 96, 86538661.CrossRefGoogle Scholar
Hoyrup, P., Davidsen, J. & Jorgensen, K. (2001). Lipid membrane partitioning of lysolipids and fatty acids: effects of membrane phase structure and detergent chain length. Journal of Physical Chemistry B 105, 26492657.CrossRefGoogle Scholar
Inoue, T., Kawamura, H., Okukado, S. & Shimozawa, R. (1994a). Characterization of molecular assemblies formed in aqueous C(10)E(7)/DPPC mixture by spin-label and fluorescence probe techniques and mechanism of micelle-to-vesicle transformation. Journal of Colloid and Interface Science 168, 94102.CrossRefGoogle Scholar
Inoue, T., Miyakawa, K. & Shimozawa, R. (1986). Interaction of surfactants with vesicle membrane of dipalmitoylphosphatidylcholine – effect on gel-to-liquid-crystalline phase-transition of lipid bilayer. Chemistry and Physics of Lipids 42, 261270.CrossRefGoogle ScholarPubMed
Inoue, T., Motoyama, R., Totoki, M., Miyakawa, K. & Shimozawa, R. (1994b). Molecular aggregates formed in aqueous mixtures of Poe type nonionic surfactants and phosphatidylcholines. Journal of Colloid and Interface Science 164, 318324.CrossRefGoogle Scholar
Inoue, T., Yanagihara, S., Misono, Y. & Suzuki, M. (2001). Effect of fatty acids on phase behavior of hydrated dipalmitoylphosphatidylcholine bilayer: saturated versus unsaturated fatty acids. Chemistry and Physics of Lipids 109, 117133.CrossRefGoogle ScholarPubMed
Ipsen, J. H., Karlstrom, G., Mouritsen, O. G., Wennerstrom, H. & Zuckermann, M. J. (1987). Phase equilibria in the phosphatidylcholine–cholesterol system. Biochimica et Biophysica Acta 905, 162172.CrossRefGoogle ScholarPubMed
Ipsen, J. H., Mouritsen, O. G. & Zuckermann, M. J. (1989). Theory of thermal anomalies in the specific heat of lipid bilayers containing cholesterol. Biophysical Journal 56, 661667.CrossRefGoogle ScholarPubMed
Isomaa, B., Hagerstrand, H., Paatero, G. & Engblom, A. C. (1986). Permeability alterations and antihaemolysis induced by amphiphiles in human erythrocytes. Biochimica et Biophysica Acta 860, 510524.CrossRefGoogle ScholarPubMed
Israelachvili, J. N. (1991). Intermolecular and Surface Forces, 2nd edn. London: Academic Press.Google Scholar
Jackson, M. L., Schmidt, C. F., Lichtenberg, D., Litman, B. J. & Albert, A. D. (1982). Solubilization of phosphatidylcholine bilayers by octyl glucoside. Biochemistry 21, 45764582.CrossRefGoogle ScholarPubMed
Jacobson, K., Mouritsen, O. G. & Anderson, R. G. W. (2007). Lipid rafts: at a crossroad between cell biology and physics. Nature Cell Biology 9, 714.CrossRefGoogle Scholar
Johnsson, M. & Bergstrand, N. (2004). Phase behavior of DOPE/TritonX100 (reduced) in dilute aqueous solution: aggregate structure and pH-dependence. Colloids and Surfaces B: Biointerfaces 34, 6976.CrossRefGoogle ScholarPubMed
Johnsson, M. & Edwards, K. (2003). Liposomes, disks, and spherical micelles: aggregate structure in mixtures of gel phase phosphatidylcholines and poly(ethylene glycol)-phospholipids. Biophysical Journal 85, 38393847.CrossRefGoogle ScholarPubMed
Kadi, M., Hansson, P. & Almgren, M. (2004). Determination of isotherms for binding of surfactants to vesicles using a surfactant-selective electrode. Journal of Physical Chemistry B 108, 73447351.CrossRefGoogle Scholar
Kamp, F., Westerhoff, H. V. & Hamilton, J. A. (1993). Movement of fatty-acids, fatty-acid analogs, and bile-acids across phospholipid-bilayers – kinetics of fatty acid-mediated proton movement across small unilamellar vesicles. Biochemistry 32, 1107411086.CrossRefGoogle Scholar
Kanduser, M., Fosnaric, M., Sentjurc, M., Kralj-Iglic, V., Hagerstrand, H., Iglic, A. & Miklavcic, D. (2003). Effect of surfactant polyoxyethylene glycol (C12E8) on electroporation of cell line DC3F. Colloids and Surfaces A: Physicochemical and Engineering Aspects 214, 205217.CrossRefGoogle Scholar
Karatekin, E., Sandre, O. & Brochard-Wyart, F. (2003). Transient pores in vesicles. Polymer International 52, 486493.CrossRefGoogle Scholar
Karlovska, J., Lohner, K., Degovics, G., Lacko, I., Devinsky, F. & Balgavy, P. (2004). Effects of non-ionic surfactants N-alkyl-N,N-dimethylamine-N-oxides on the structure of a phospholipid bilayer: small-angle X-ray diffraction study. Chemistry and Physics of Lipids 129, 3141.CrossRefGoogle ScholarPubMed
Kaufmann, T. C., Engel, A. & Remigy, H. W. (2006). A novel method for detergent concentration determination. Biophysical Journal 90, 310317.CrossRefGoogle ScholarPubMed
Keller, M., Kerth, A. & Blume, A. (1997). Thermodynamics of interaction of octyl glucoside with phosphatidylcholine vesicles: partitioning and solubilization as studied by high sensitivity titration calorimetry. Biochimica et Biophysica Acta 1326, 178192.CrossRefGoogle ScholarPubMed
Keller, S., Bothe, M., Bienert, M., Dathe, M. & Blume, A. (2007). A simple fluorescence-spectroscopic membrane translocation assay. ChemBioChem 8, 546552.CrossRefGoogle ScholarPubMed
Keller, S., Heerklotz, H. & Blume, A. (2006a). Monitoring lipid membrane translocation of sodium dodecyl sulfate by isothermal titration calorimetry. Journal of the American Chemical Society 128, 12791286.CrossRefGoogle ScholarPubMed
Keller, S., Heerklotz, H., Jahnke, N. & Blume, A. (2006b). Thermodynamics of lipid membrane solubilization by sodium dodecyl sulfate. Biophysical Journal 90, 45094521.CrossRefGoogle ScholarPubMed
Keller, S., Sauer, I., Strauss, H., Gast, K., Dathe, M. & Bienert, M. (2005a). Membrane-mimetic nanocarriers formed by a dipalmitoylated cell-penetrating peptide. Angewandte Chemie – International Edition 44, 52525255.CrossRefGoogle ScholarPubMed
Keller, S., Tsamaloukas, A. & Heerklotz, H. (2005b). A quantitative model describing the selective solubilization of membrane domains. Journal of the American Chemical Society 127, 1146911476.CrossRefGoogle ScholarPubMed
Kelly, E., Prive, G. G. & Tieleman, D. P. (2005). Molecular models of lipopeptide detergents: large coiled-coils with hydrocarbon interiors. Journal of the American Chemical Society 127, 1344613447.CrossRefGoogle ScholarPubMed
Klose, G., Eisenblatter, S. & Konig, B. (1995). Ternary phase-diagram of mixtures of palmitoyl-oleoyl-phosphatidylcholine, tetraoxyethylene dodecyl ether, and heavy-water as seen by 31P and 2H NMR. Journal of Colloid and Interface Science 172, 438446.CrossRefGoogle Scholar
Klose, G., Islamov, A., Konig, B. & Cherezov, V. (1996). Structure of mixed multilayers of palmitoyloleoylphosphatidylcholine and oligo(oxyethylene glycol) monododecyl ether determined by X-ray and neutron diffraction. Langmuir 12, 409415.CrossRefGoogle Scholar
Klose, G. & Levine, Y. K. (2000). Membranes of palmitoyloleoylphosphatidylcholine and C12E4 – a lattice model simulation. Langmuir 16, 671676.CrossRefGoogle Scholar
Klose, G., Madler, B., Schafer, H. & Schneider, K. P. (1999). Structural characterization of POPC, and C12E4 in their mixed membranes at reduced hydration by solid state 2H NMR. Journal of Physical Chemistry B 103, 30223029.CrossRefGoogle Scholar
Kluge, S., Gawrisch, K. & Nuhn, P. (1987). Loss of infectivity of red clover mottle virus by lysolecithin. Acta Virology 31, 185188.Google ScholarPubMed
Koltover, I., Salditt, T., Radler, J. O. & Safinya, C. R. (1998). An inverted hexagonal phase of cationic liposome–DNA complexes related to DNA release and delivery. Science 281, 7881.CrossRefGoogle ScholarPubMed
Konig, B., Dietrich, U. & Klose, G. (1997). Hydration and structural properties of mixed lipid/surfactant model membranes. Langmuir 13, 525532.CrossRefGoogle Scholar
Kozlov, M. M., Lichtenberg, D. & Andelman, D. (1997). Shape of phospholipid/surfactant mixed micelles: cylinders or disks? Theoretical analysis. Journal of Physical Chemistry B 101, 66006606.CrossRefGoogle Scholar
Kragh-Hansen, U., le Maire, M. & Moller, J. V. (1998). The mechanism of detergent solubilization of liposomes and protein-containing membranes. Biophysical Journal 75, 29322946.CrossRefGoogle ScholarPubMed
Kresheck, G. C. & Mihelich, J. (2003). Observation of complex thermal transitions for mixed micelle solutions containing alkyldimethylphosphine oxides and phospholipids and the accompanying cloud points. Chemistry and Physics of Lipids 123, 4562.CrossRefGoogle ScholarPubMed
Kresheck, G. C. & Nimsgern, R. A. (1983). Unusual enthalpy changes which accompany the titration of dimyristoylphosphatidylcholine vesicles with Triton X-100. Chemistry and Physics of Lipids 33, 5565.CrossRefGoogle ScholarPubMed
Ladokhin, A. S., Selsted, M. E. & White, S. H. (1997). Sizing membrane pores in lipid vesicles by leakage of co-encapsulated markers: pore formation by melittin. Biophysical Journal 72, 17621766.CrossRefGoogle ScholarPubMed
Ladokhin, A. S. & White, S. H. (2001). ‘Detergent-like’ permeabilization of anionic lipid vesicles by melittin. Biochimica et Biophysica Acta 1514, 253260.CrossRefGoogle ScholarPubMed
Ladokhin, A. S., Wimley, W. C. & White, S. H. (1995). Leakage of membrane vesicle contents: determination of mechanism using fluorescence requenching. Biophysical Journal 69, 19641971.CrossRefGoogle ScholarPubMed
Landau, E. M. & Rosenbusch, J. P. (1996). Lipidic cubic phases: a novel concept for the crystallization of membrane proteins. Proceedings of the National Academy of Sciences USA 93, 1453214535.CrossRefGoogle ScholarPubMed
Lantzsch, G., Binder, H., Heerklotz, H., Wendling, M. & Klose, G. (1996). Surface areas and packing constraints in POPC/C12EOn membranes. A time-resolved fluorescence study. Biophysical Chemistry 58, 289302.CrossRefGoogle Scholar
Lasch, J. (1995). Interaction of detergents with lipid vesicles. Biochimica et Biophysica Acta 1241, 269292.CrossRefGoogle ScholarPubMed
Lasch, J., Hoffmann, J., Omelyanenko, W. G., Klibanov, A. A., Torchilin, V. P., Binder, H. & Gawrisch, K. (1990). Interaction of Triton X-100 and octyl glucoside with liposomal membranes at sublytic and lytic concentrations. Spectroscopic studies. Biochimica et Biophysica Acta 1022, 171180.CrossRefGoogle ScholarPubMed
le Maire, M., Champeil, P. & Moller, J. V. (2000). Interaction of membrane proteins and lipids with solubilizing detergents. Biochimica et Biophysica Acta 1508, 86111.CrossRefGoogle ScholarPubMed
le Maire, M., Moller, J. V. & Champeil, P. (1987). Binding of a nonionic detergent to membranes: flip-flop rate and location on the bilayer. Biochemistry 26, 48034810.CrossRefGoogle ScholarPubMed
Lemmich, J., Richter, F. & Callisen, T. H. (1998). Structural studies on equimolar suspensions of palmitic acid and 1-lyso-palmitoyl-phosphatidylcholine. Material Research Society Symposium Proceedings 489, 125130.CrossRefGoogle Scholar
Leng, J., Egelhaaf, S. U. & Cates, M. E. (2003). Kinetics of the micelle-to-vesicle transition: aqueous lecithin–bile salt mixtures. Biophysical Journal 85, 16241646.CrossRefGoogle ScholarPubMed
Lesieur, S., Grabielle-Madelmont, C., Menager, C., Cabuil, V., Dadhi, D., Pierrot, P. & Edwards, K. (2003). Evidence of surfactant-induced formation of transient pores in lipid bilayers by using magnetic-fluid-loaded liposomes. Journal of American Chemical Society 125, 52665267.CrossRefGoogle ScholarPubMed
Lichtenberg, D. (1985). Characterization of the solubilization of lipid bilayers by surfactants. Biochimica et Biophysica Acta 821, 470478.CrossRefGoogle ScholarPubMed
Lichtenberg, D. (1993). Micelles and liposomes. In Biomembranes – Physical Aspects (ed. Shinitzky, M.), pp. 6396. Weinheim: VCH.Google Scholar
Lichtenberg, D., Goni, F. M. & Heerklotz, H. (2005). Detergent-resistant membranes should not be identified with membrane rafts. Trends in Biochemical Sciences 30, 430436.CrossRefGoogle Scholar
Lichtenberg, D., Opatowski, E. & Kozlov, M. M. (2000). Phase boundaries in mixtures of membrane-forming amphiphiles and micelle-forming amphiphiles. Biochimica et Biophysica Acta 1508, 119.CrossRefGoogle ScholarPubMed
Lichtenberg, D., Robson, R. J. & Dennis, E. A. (1983). Solubilization of phospholipids by detergents. Structural and kinetic aspects. Biochimica et Biophysica Acta 737, 285304.CrossRefGoogle ScholarPubMed
London, E. (2005). How principles of domain formation in model membranes may explain ambiguities concerning lipid raft formation in cells. Biochimica et Biophysica Acta – Molecular Cell Research 1746, 203220.CrossRefGoogle ScholarPubMed
Long, M. A., Kaler, E. W. & Lee, S. P. (1994). Structural characterization of the micelle–vesicle transition in lecithin–bile salt solutions. Biophysical Journal 67, 17331742.CrossRefGoogle ScholarPubMed
Lopez, O., Cocera, M., de la Maza, A., Coderch, L. & Parra, J. L. (2000). Different stratum corneum lipid liposomes as models to evaluate the effect of the sodium dodecyl sulfate. Biochimica et Biophysica Acta 1508, 196209.CrossRefGoogle ScholarPubMed
Lopez, O., de la Maza, A., Coderch, L., Lopez-Iglesias, C., Wehrli, E. & Parra, J. L. (1998). Direct formation of mixed micelles in the solubilization of phospholipid liposomes by Triton X-100. FEBS Letters 426, 314318.CrossRefGoogle ScholarPubMed
Loudet, C., Khemtemourian, L., Aussenac, F., Gineste, S., Achard, M. F. & Dufourc, E. J. (2005). Bicelle membranes and their use for hydrophobic peptide studies by circular dichroism and solid state NMR. Biochimica et Biophysica Acta – General Subjects 1724, 315323.CrossRefGoogle ScholarPubMed
Luchette, P. A., Vetman, T. N., Prosser, R. S., Hancock, R. E., Nieh, M. P., Glinka, C. J., Krueger, S. & Katsaras, J. (2001). Morphology of fast-tumbling bicelles: a small angle neutron scattering and NMR study. Biochimica et Biophysica Acta 1513, 8394.CrossRefGoogle ScholarPubMed
Madler, B., Binder, H. & Klose, G. (1998). Compound complex formation in phospholipid membranes induced by a nonionic surfactant of the oligo(ethylene oxide) alkyl ether type: a comparative DSC and FTIR study. Journal of Colloid and Interface Science 202, 124138.CrossRefGoogle Scholar
Madler, B., Klose, G., Mops, A., Richter, W. & Tschierske, C. (1994). Thermotropic phase behaviour of the pseudobinary mixture DPPC/C12E4 at excess water. Chemistry and Physics of Lipids 71, 112.CrossRefGoogle Scholar
Majhi, P. R. & Blume, A. (2001). Thermodynamic characterization of temperature-induced micellization and demicellization of detergents studied by differential scanning calorimetry. Langmuir 17, 38443851.CrossRefGoogle Scholar
Majhi, P. R. & Blume, A. (2002). Temperature-induced micelle–vesicle transitions in DMPC–SDS and DMPC–DTAB mixtures studied by calorimetry and dynamic light scattering. Journal of Physical Chemistry B 106, 1075310763.CrossRefGoogle Scholar
Manceva, S. D., Pusztai-Carey, M., Russo, P. S. & Butko, P. (2005). A detergent-like mechanism of action of the cytolytic toxin Cyt1A from Bacillus thuringiensis var. israelensis. Biochemistry 44, 589597.CrossRefGoogle ScholarPubMed
Marcelino, J., Lima, J. L. F. C., Reis, S. & Matos, C. (2007). Assessing the effects of surfactants on the physical properties of liposome membranes. Chemistry and Physics of Lipids 146, 94103.CrossRefGoogle ScholarPubMed
Marcotte, L., Barbeau, J., Edwards, K., Karlsson, G. & Lafleur, M. (2005a). Influence of the lipid composition on the membrane affinity, and the membrane-perturbing ability of cetylpyridinium chloride. Colloids and Surfaces A: Physicochemical and Engineering Aspects 266, 5161.CrossRefGoogle Scholar
Marcotte, L., Barbeau, J. & Lafleur, M. (2005b). Permeability and thermodynamics study of quaternary ammonium surfactants – phosphocholine vesicle system. Journal of Colloid and Interface Science 292, 219227.CrossRefGoogle ScholarPubMed
Martin, I., Dubois, M. C., Saermark, T., Epand, R. M. & Ruysschaert, J. M. (1993). Lysophosphatidylcholine mediates the mode of insertion of the NH2-terminal SIV fusion peptide into the lipid bilayer. FEBS Letters 333, 325330.CrossRefGoogle ScholarPubMed
Matsuzaki, K., Murase, O., Fujii, N. & Miyajima, K. (1996). An antimicrobial peptide, magainin 2, induced rapid flip-flop of phospholipids coupled with pore formation and peptide translocation. Biochemistry 35, 1136111368.CrossRefGoogle ScholarPubMed
Matsuzaki, K., Sugishita, K., Ishibe, N., Ueha, M., Nakata, S., Miyajima, K. & Epand, R. M. (1998). Relationship of membrane curvature to the formation of pores by magainin 2. Biochemistry 37, 1185611863.CrossRefGoogle Scholar
May, S. & Ben-Shaul, A. (2001). Molecular theory of the sphere-to-rod transition and the second CMC in aqueous micellar solutions. Journal of Physical Chemistry B 105, 630640.CrossRefGoogle Scholar
Mazer, N. A., Benedek, G. B. & Carey, M. C. (1980). Quasi-elastic light-scattering-studies of aqueous biliary lipid systems – mixed micelle formation in bile-salt lecithin solutions. Biochemistry 19, 601615.CrossRefGoogle Scholar
McGregor, C. L., Chen, L., Pomroy, N. C., Hwang, P., Go, S., Chakrabartty, A. & Prive, G. G. (2003). Lipopeptide detergents designed for the structural study of membrane proteins. Nature Biotechnology 21, 171176.CrossRefGoogle ScholarPubMed
McLaughlin, S. (1989). The electrostatic properties of membranes. Annual Review of Biophysics and Biophysical Chemistry 18, 113136.CrossRefGoogle ScholarPubMed
McMullen, T. P. W., Lewis, R. N. A. H. & McElhaney, R. N. (2004). Cholesterol–phospholipid interactions, the liquid-ordered phase and lipid rafts in model and biological membranes. Current Opinion in Colloid and Interface Science 8, 459468.CrossRefGoogle Scholar
Meagher, R. J. & Hatton, T. A. (1998). Enthalpy measurements in aqueous SDS/DTAB solutions using isothermal titration microcalorimetry. Langmuir 14, 40814087.CrossRefGoogle Scholar
Meister, A. & Blume, A. (2004). Solubilization of DMPC-d54 and DMPG-d54 vesicles with octylglucoside and sodium dodecyl sulfate studied by FT-IR spectroscopy. Physical Chemistry Chemical Physics 6, 15511556.CrossRefGoogle Scholar
Meister, A., Kerth, A. & Blume, A. (2004a). The interaction of n-nonyl-beta-d-glucopyranoside and sodium dodecyl sulfate with DMPC and DMPG monolayers studied by infrared reflection absorption spectroscopy. Physical Chemistry Chemical Physics 6, 55435550.CrossRefGoogle Scholar
Meister, A., Kerth, A. & Blume, A. (2004b). Interaction of sodium dodecyl sulfate with dimyristoyl-sn-glycero-3-phosphocholine monolayers studied by infrared reflection absorption spectroscopy. A new method for the determination of surface partition coefficients. Journal of Physical Chemistry B 108, 83718378.CrossRefGoogle Scholar
Moschetta, A., Frederik, P. M., Portincasa, P., VanBerge-Henegouwen, G. P. & Van Erpecum, K. J. (2002). Incorporation of cholesterol in sphingomyelin–phosphatidylcholine vesicles has profound effects on detergent-induced phase transitions. Journal of Lipid Research 43, 10461053.CrossRefGoogle ScholarPubMed
Mui, B. L., Dobereiner, H. G., Madden, T. D. & Cullis, P. R. (1995). Influence of transbilayer area asymmetry on the morphology of large unilamellar vesicles. Biophysical Journal 69, 930941.CrossRefGoogle ScholarPubMed
Munro, S. (2003). Lipid rafts: elusive or illusive? Cell 115, 377388.CrossRefGoogle ScholarPubMed
Nagle, J. F. & Tristram-Nagle, S. (2000). Structure of lipid bilayers. Biochimica et Biophysica Acta 1469, 159195.CrossRefGoogle ScholarPubMed
Nernst, W. (1891). Verteilung eines Stoffes zwischen zwei Lösungsmitteln und zwischen Lösungsmittel und Dampfraum. Zeitschrift Physikalische Chemie 8, 110.CrossRefGoogle Scholar
Nibbering, C. P., Frederik, P. M., Van Berge-Henegouwen, G. P., Van Veen, H. A., Van Marle, J. & Van Erpecum, K. J. (2002). Different interactions of egg-yolk phosphatidylcholine and sphingomyelin with detergent bile salts. Biochimica et Biophysica Acta – Molecular and Cell Biology of Lipids 1583, 213220.CrossRefGoogle ScholarPubMed
Nicolini, C., Thiyagarajan, P. & Winter, R. (2004). Small-scale composition fluctuations and microdomain formation in lipid raft models as revealed by small-angle neutron scattering. Physical Chemistry Chemical Physics 6, 55315534.CrossRefGoogle Scholar
Nilsson, P. G., Wennerstrom, H. & Lindman, B. (1983). Structure of micellar solutions of non-ionic surfactants – nuclear magnetic-resonance self-diffusion and proton relaxation studies of poly(ethylene oxide) alkyl ethers. Journal of Physical Chemistry 87, 13771385.CrossRefGoogle Scholar
Nyholm, T. & Slotte, J. P. (2001). Comparison of Triton X-100 penetration into phosphatidylcholine and sphingomyelin mono- and bilayers. Langmuir 17, 47244730.CrossRefGoogle Scholar
Ollila, F. & Slotte, J. P. (2002). Partitioning of Triton X-100, deoxycholate and C10EO8 into bilayers composed of native and hydrogenated egg yolk sphingomyelin. Biochimica et Biophysica Acta 1564, 281288.CrossRefGoogle ScholarPubMed
Ollivon, M., Eidelman, O., Blumenthal, R. & Walter, A. (1988). Micelle–vesicle transition of egg phosphatidylcholine and octyl glucoside. Biochemistry 27, 16951703.CrossRefGoogle ScholarPubMed
Ollivon, M., Lesieur, S., Grabielle-Madelmont, C. & Paternostre, M. (2000). Vesicle reconstitution from lipid–detergent mixed micelles. Biochimica et Biophysica Acta 1508, 3450.CrossRefGoogle ScholarPubMed
Opatowski, E., Kozlov, M. M. & Lichtenberg, D. (1997a). Partitioning of octyl glucoside between octyl glucoside/phosphatidylcholine mixed aggregates and aqueous media as studied by isothermal titration calorimetry. Biophysical Journal 73, 14481457.CrossRefGoogle Scholar
Opatowski, E., Lichtenberg, D. & Kozlov, M. M. (1997b). The heat of transfer of lipid and surfactant from vesicles into micelles in mixtures of phospholipid and surfactant. Biophysical Journal 73, 14581467.CrossRefGoogle ScholarPubMed
Otten, D., Brown, M. F. & Beyer, K. (2000). Softening of membrane bilayers by detergents elucidated by deuterium NMR spectroscopy. Journal of Physical Chemistry B 104, 1211912129.CrossRefGoogle Scholar
Otten, D., Lobbecke, L. & Beyer, K. (1995). Stages of the bilayer–micelle transition in the system phosphatidylcholine-C(12)E(8) as studied by deuterium-NMR and phosphorus-NMR, light-scattering, and calorimetry. Biophysical Journal 68, 584597.CrossRefGoogle Scholar
Pantaler, E., Kamp, D. & Haest, C. W. (2000). Acceleration of phospholipid flip-flop in the erythrocyte membrane by detergents differing in polar head group and alkyl chain length. Biochimica et Biophysica Acta 1509, 397408.CrossRefGoogle ScholarPubMed
Partearroyo, M. A., Alonso, A., Goni, F. M., Tribout, M. & Paredes, S. (1996). Solubilization of phospholipid bilayers by surfactants belonging to the Triton X series: effect of polar group size. Journal of Colloid and Interface Science 178, 156159.CrossRefGoogle Scholar
Paternostre, M., Meyer, O., Grabielle-Madelmont, C., Lesieur, S., Ghanam, M. & Ollivon, M. (1995). Partition coefficient of a surfactant between aggregates and solution: application to the micelle–vesicle transition of egg phosphatidylcholine and octyl beta-d-glucopyranoside. Biophysical Journal 69, 24762488.CrossRefGoogle Scholar
Paternostre, M. T., Roux, M. & Rigaud, J. L. (1988). Mechanisms of membrane protein insertion into liposomes during reconstitution procedures involving the use of detergents. 1. Solubilization of large unilamellar liposomes (prepared by reverse-phase evaporation) by triton X-100, octyl glucoside, and sodium cholate. Biochemistry 27, 26682677.CrossRefGoogle ScholarPubMed
Patra, S. K., Alonso, A., Arrondo, J. L. R. & Goni, F. M. (1999). Liposomes containing sphingomyelin and cholesterol: detergent solubilization and infrared spectroscopic studies. Journal of Liposome Research 9, 247260.CrossRefGoogle Scholar
Patra, S. K., Alonso, A. & Goni, F. M. (1998). Detergent solubilisation of phospholipid bilayers in the gel state: the role of polar and hydrophobic forces. Biochimica et Biophysica Acta 1373, 112118.CrossRefGoogle ScholarPubMed
Pfeiffer, H., Klose, G., Heremans, K. & Glorieux, C. (2006). Thermotropic phase behaviour of the pseudobinary mixtures of DPPC/C12E5 and DMPC/C12E5 determined by differential scanning calorimetry and ultrasonic velocimetry. Chemistry and Physics of Lipids 139, 5467.CrossRefGoogle ScholarPubMed
Pike, L. J. (2006). Rafts defined: a report on the Keystone symposium on lipid rafts and cell function. Journal of Lipid Research 47, 15971598.CrossRefGoogle ScholarPubMed
Prenner, E. J., Lewis, R. N., Jelokhani-Niaraki, M., Hodges, R. S. & McElhaney, R. N. (2001). Cholesterol attenuates the interaction of the antimicrobial peptide gramicidin S with phospholipid bilayer membranes. Biochimica et Biophysica Acta 1510, 8392.CrossRefGoogle ScholarPubMed
Prive, G. G. (2007). Detergents for the stabilization and crystallization of membrane proteins. Methods 41, 388397.CrossRefGoogle ScholarPubMed
Radler, J. O., Koltover, I., Salditt, T. & Safinya, C. R. (1997). Structure of DNA–cationic liposome complexes: DNA intercalation in multilamellar membranes in distinct interhelical packing regimes. Science 275, 810814.CrossRefGoogle ScholarPubMed
Rakotomanga, M., Loiseau, P. M. & Saint-Pierre-Chazalet, M. (2004). Hexadecylphosphocholine interaction with lipid monolayers. Biochimica et Biophysica Acta – Biomembranes 1661, 212218.CrossRefGoogle ScholarPubMed
Rand, P., Fuller, N. & Chen, Z. S. (1998). Measuring spontaneous curvature and bending moduli of membrane lipid layers. Biophysical Journal 74, A27A27.Google Scholar
Rauch, C. & Farge, E. (2000). Endocytosis switch controlled by transmembrane osmotic pressure and phospholipid number asymmetry. Biophysical Journal 78, 30363047.CrossRefGoogle ScholarPubMed
Record, M. T. Jr.,, Anderson, C. F. & Lohman, T. M. (1978). Thermodynamic analysis of ion effects on the binding and conformational equilibria of proteins and nucleic acids: the roles of ion association or release, screening, and ion effects on water activity. Quarterly Reviews of Biophysics 11, 103178.CrossRefGoogle ScholarPubMed
Redlich, O. & Kister, A. (1948). Algebraic representation of thermodynamic properties and the classification of solutions. Industrial and Engineering Chemistry 40, 345348.CrossRefGoogle Scholar
Ribeiro, A. A. & Dennis, E. A. (1975). Proton magnetic resonance relaxation studies on the structure of mixed micelles of Triton X-100 and dimyristoylphosphatidylcholine. Biochemistry 14, 37463755.CrossRefGoogle ScholarPubMed
Rieber, K., Sykora, J., Olzynska, A., Jelinek, R., Cevc, G. & Hof, M. (2007). The use of solvent relaxation technique to investigate head group hydration and protein binding of simple and mixed phosphatidylcholine/surfactant bilayer membranes. Biochimica et Biophysica Acta – Biomembranes 1768, 10501058.CrossRefGoogle Scholar
Rietveld, T. G., Chupin, V. V., Koorengevel, M. C., Wienk, H. L. J., Dowhan, W. & De Kruijff, B. (1994). Regulation of lipid polymorphism is essential for the viability of phosphatidylethanolamine-deficient Escherichia coli cells. Journal of Biological Chemistry 269, 2867028675.CrossRefGoogle ScholarPubMed
Rigaud, J. L. & Levy, D. (2003). Reconstitution of membrane proteins into liposomes. Methods in Enzymology 372, 6586.CrossRefGoogle ScholarPubMed
Rosenberg, E. & Ron, E. Z. (1999). High- and low-molecular-mass microbial surfactants. Applied Microbiology and Biotechnology 52, 154162.CrossRefGoogle ScholarPubMed
Roth, Y., Opatowski, E., Lichtenberg, D. & Kozlov, M. M. (2000). Phase behavior of dilute aqueous solutions of lipid–surfactant mixtures: effects of finite size of micelles. Langmuir 16, 20522061.CrossRefGoogle Scholar
Rowe, E. S., Zhang, F., Leung, T. W., Parr, J. S. & Guy, P. T. (1998). Thermodynamics of membrane partitioning for a series on n-alcohols determined by titration calorimetry: role of hydrophobic effects. Biochemistry 37, 24302440.CrossRefGoogle ScholarPubMed
Ruiz, J., Goni, F. M. & Alonso, A. (1988). Surfactant-induced release of liposomal contents – a survey of methods and results. Biochimica et Biophysica Acta 937, 127134.CrossRefGoogle ScholarPubMed
Sadaghiani, A. S., Khan, A. & Lindman, B. (1989). Liquid crystallinity of lecithin systems. Ternary phase diagrams of lecithin–water with Triton X-100 and decanol. Journal of Colloid and Interface Science 132, 352363.CrossRefGoogle Scholar
Saez-Cirion, A., Alonso, A., Goni, F. M., McMullen, T. P. W., McElhaney, R. N. & Rivas, E. A. (2000). Equilibrium and kinetic studies of the solubilization of phospholipid–cholesterol bilayers by C12E8. The influence of the lipid phase structure. Langmuir 16, 19601968.CrossRefGoogle Scholar
Saitoh, A., Takiguchi, K., Tanaka, Y. & Hotani, H. (1998). Opening-up of liposomal membranes by talin. Proceedings of the National Academy of Sciences USA 95, 10261031.CrossRefGoogle ScholarPubMed
Sanchez, L., Martinez, V., Infante, M. R., Mitjans, M. & Vinardell, M. P. (2007). Hemolysis and antihemolysis induced by amino acid-based surfactants. Toxicology Letters 169, 177184.CrossRefGoogle ScholarPubMed
Sanders, C. R. & Landis, G. C. (1995). Reconstitution of membrane proteins into lipid-rich bilayered mixed micelles for NMR studies. Biochemistry 34, 40304040.CrossRefGoogle ScholarPubMed
Sanders, C. R. & Oxenoid, K. (2000). Customizing model membranes and samples for NMR spectroscopic studies of complex membrane proteins. Biochimica et Biophysica Acta 1508, 129145.CrossRefGoogle ScholarPubMed
Schneider, M. J. & Feller, S. E. (2001). Molecular dynamics simulations of a phospholipid–detergent mixture. Journal of Physical Chemistry B 105, 13311337.CrossRefGoogle Scholar
Schnitzer, E., Kozlov, M. M. & Lichtenberg, D. (2005). The effect of cholesterol on the solubilization of phosphatidylcholine bilayers by the non-ionic surfactant Triton X-100. Chemistry and Physics of Lipids 135, 6982.CrossRefGoogle ScholarPubMed
Schubert, R. (2003). Liposome preparation by detergent removal. Methods in Enzymology 367, 4670.CrossRefGoogle ScholarPubMed
Schubert, R., Wolburg, H., Schmidt, K. H. & Roth, H. J. (1991). Loading of preformed liposomes with high trapping efficiency by detergent-induced formation of transient membrane holes. Chemistry and Physics of Lipids 58, 121129.CrossRefGoogle Scholar
Schurholz, T. (1996). Critical dependence of the solubilization of lipid vesicles by the detergent CHAPS on the lipid composition. Functional reconstitution of the nicotinic acetylcholine receptor into preformed vesicles above the critical micellization concentration. Biophysical Chemistry 58, 8796.CrossRefGoogle ScholarPubMed
Schurtenberger, P., Mazer, N. & Känzig, W. (1985). Micelle to vesicle transition in aqueous solutions of bile salt and lecithin. Journal of Physical Chemistry 89, 10421049.CrossRefGoogle Scholar
Schwarz, S., Haest, C. W. & Deuticke, B. (1999). Extensive electroporation abolishes experimentally induced shape transformations of erythrocytes: a consequence of phospholipid symmetrization? Biochimica et Biophysica Acta 1421, 361379.CrossRefGoogle ScholarPubMed
Seelig, A. & Gerebtzoff, G. (2006). Enhancement of drug absorption by noncharged detergents through membrane and P-glycoprotein binding. Expert Opinion on Drug Metabolism and Toxicology 2, 733752.CrossRefGoogle ScholarPubMed
Seelig, J. (1997). Titration calorimetry of lipid–peptide interactions. Biochimica et Biophysica Acta 1331, 103116.CrossRefGoogle ScholarPubMed
Seelig, J. & Ganz, P. (1991). Non-classical hydrophobic effect in membrane binding equilibria. Biochemistry 30, 93549359.CrossRefGoogle Scholar
Seifert, M., Breitenstein, D., Klenz, U., Meyer, M. C. & Galla, H. J. (2007). Solubility versus electrostatics: what determines lipid/protein interaction in lung surfactant. Biophysical Journal 93, 11921203.CrossRefGoogle ScholarPubMed
Sharp, K. A., Nicholls, A., Fine, R. F. & Honig, B. (1991). Reconciling the magnitude of the microscopic and macroscopic hydrophobic effects. Science 252, 106109.CrossRefGoogle ScholarPubMed
Sheetz, M. P. & Singer, S. J. (1974). Biological membranes as bilayer couples. A molecular mechanism of drug–erythrocyte interactions. Proceedings of the National Academy of Sciences USA 71, 44574461.CrossRefGoogle ScholarPubMed
Simes, S. I., Marques, C. M., Cruz, M. E. M., Cevc, G. & Martins, M. B. F. (2004). The effect of cholate on solubilisation and permeability of simple and protein-loaded phosphatidylcholine/sodium cholate mixed aggregates designed to mediate transdermal delivery of macromolecules. European Journal of Pharmaceutics and Biopharmaceutics 58, 509519.CrossRefGoogle Scholar
Simons, K. & Ikonen, E. (1997). Functional rafts in cell membranes. Nature 387, 569572.CrossRefGoogle ScholarPubMed
Simons, K. & Vaz, W. L. (2004). Model systems, lipid rafts, and cell membranes. Annual Review of Biophysics and Biomolecular Structures 33, 269295.CrossRefGoogle ScholarPubMed
Small, D. M. (1970). Surface and bulk interactions of lipids and water with a classification of biologically active lipids based on these interactions. Federation Proceedings 29, 13201326.Google ScholarPubMed
Small, D. M. (1986). The Physical Chemistry of Lipids. New York: Plenum Press.CrossRefGoogle Scholar
Sot, J., Bagatolli, L. A., Goni, F. M. & Alonso, A. (2006). Detergent-resistant, ceramide-enriched domains in sphingomyelin/ceramide bilayers. Biophysical Journal 90, 903914.CrossRefGoogle ScholarPubMed
Sot, J., Collado, M. I., Arrondo, J. L. R., Alonso, A. & Goni, F. M. (2002). Triton X-100-resistant bilayers: effect of lipid composition and relevance to the raft phenomenon. Langmuir 18, 28282835.CrossRefGoogle Scholar
Spolar, R. S., Ha, J. H. & Record, M. T. Jr., (1989). Hydrophobic effect in protein folding and other noncovalent processes involving proteins. Proceedings of the National Academy of Sciences USA 86, 83828385.CrossRefGoogle ScholarPubMed
Staneva, G., Seigneuret, M., Koumanov, K., Trugnan, G. & Angelova, M. I. (2005). Detergents induce raft-like domains budding and fission from giant unilamellar heterogeneous vesicles: a direct microscopy observation. Chemistry and Physics of Lipids 136, 5566.CrossRefGoogle ScholarPubMed
Tamba, Y., Tanaka, T., Yahagi, T., Yamashita, Y. & Yamazaki, M. (2004). Stability of giant unilamellar vesicles and large unilamellar vesicles of liquid-ordered phase membranes in the presence of Triton X-100. Biochimica et Biophysica Acta – Biomembranes 1667, 16.CrossRefGoogle ScholarPubMed
Tan, A. M., Ziegler, A., Steinbauer, B. & Seelig, J. (2002). Thermodynamics of sodium dodecyl sulfate partitioning into lipid membranes. Biophysical Journal 83, 15471556.CrossRefGoogle ScholarPubMed
Tanford, C. (1980). The Hydrophobic Effect: Formation of Micelles and Biological Membranes, 2nd edn. New York: Wiley.Google Scholar
Taranto, M. P., Fernandez Murga, M. L., Lorca, G. & De Valdez, G. F. (2003). Bile salts and cholesterol induce changes in the lipid cell membrane of Lactobacillus reuteri. Journal of Applied Microbiology 95, 8691.CrossRefGoogle ScholarPubMed
Thurmond, R. L., Otten, D., Brown, M. F. & Beyer, K. (1994). Structure and packing of phosphatidylcholines in lamellar and hexagonal liquid–crystalline mixtures with a nonionic detergent – a wide-line deuterium and P-31 NMR-study. Journal of Physical Chemistry 98, 972983.CrossRefGoogle Scholar
Treyer, M., Walde, P. & Oberholzer, T. (2002). Permeability enhancement of lipid vesicles to nucleotides by use of sodium cholate: basic studies and application to an enzyme-catalyzed reaction occurring inside the vesicles. Langmuir 18, 10431050.CrossRefGoogle Scholar
Tsamaloukas, A., Szadkowska, H. & Heerklotz, H. (2006). Nonideal mixing in multicomponent lipid/detergent systems. Journal of Physics Condensed Matter 18, S1125S1138.CrossRefGoogle ScholarPubMed
Tsamaloukas, A., Szadkowska, H., Slotte, P. J. & Heerklotz, H. (2005). Interactions of cholesterol with lipid membranes and cyclodextrin characterized by calorimetry. Biophysical Journal 89, 11091119.CrossRefGoogle ScholarPubMed
Tsamaloukas, A. D., Keller, S. & Heerklotz, H. (2007). Uptake and release protocol for assessing membrane binding and permeation by way of isothermal titration calorimetry. Nature Protocols 2, 695704.CrossRefGoogle ScholarPubMed
Tsuzuki, W., Ue, A., Nagao, A., Endo, M. & Abe, M. (2004). Inhibitory effect of lysophosphatidylcholine on pancreatic lipase-mediated hydrolysis in lipid emulsion. Biochimica et Biophysica Acta – Molecular and Cell Biology of Lipids 1684, 17.CrossRefGoogle ScholarPubMed
Ueno, M. (1989). Partition behavior of a nonionic detergent, octyl glucoside, between membrane and water phases, and its effect on membrane permeability. Biochemistry 28, 56315634.CrossRefGoogle ScholarPubMed
Ueno, M., Hirota, N., Kashiwagi, H. & Sagasaki, S. (2003). Process of destruction of large unilamellar vesicles by a nonionic detergent, octylglucoside, and size growth factor in vesicle formation from phospholipid–detergent mixed micelles. Colloid and Polymer Science 282, 6975.CrossRefGoogle Scholar
Valpuesta, J. M., Arrondo, J. L., Barbero, M. C., Pons, M. & Goni, F. M. (1986). Membrane-surfactant interactions. The role of surfactant in mitochondrial complex III–phospholipid–Triton X-100 mixed micelles. Journal of Biological Chemistry 261, 65786584.CrossRefGoogle ScholarPubMed
Van Dam, L., Karlsson, G. & Edwards, K. (2004). Direct observation and characterization of DMPC/DHPC aggregates under conditions relevant for biological solution NMR. Biochimica et Biophysica Acta – Biomembranes 1664, 241256.CrossRefGoogle ScholarPubMed
Van Den Brink-Van Der Laan, E., Antoinette Killian, J. & De Kruijff, B. (2004). Nonbilayer lipids affect peripheral and integral membrane proteins via changes in the lateral pressure profile. Biochimica et Biophysica Acta – Biomembranes 1666, 275288.CrossRefGoogle ScholarPubMed
van Rheenen, J., Mulugeta Achame, E., Janssen, H., Calafat, J. & Jalink, K. (2005). PIP(2) signaling in lipid domains: a critical re-evaluation. EMBO Journal 24, 16641673.CrossRefGoogle Scholar
Vanderkooi, G., Senior, A. E., Capaldi, R. A. & Hayashi, H. (1972). Biological membrane structure. 3. The lattice structure of membranous cytochrome oxidase. Biochimica et Biophysica Acta 274, 3848.CrossRefGoogle ScholarPubMed
Viguera, A. R., Gonzalezmanas, J. M., Taneva, S. & Goni, F. M. (1994). Early and delayed stages in the solubilization of purple membrane by a polyoxyethylenic surfactant. Biochimica et Biophysica Acta – Biomembranes 1196, 7680.CrossRefGoogle ScholarPubMed
Vinson, P. K., Talmon, Y. & Walter, A. (1989). Vesicle–micelle transition of phosphatidylcholine and octyl glucoside elucidated by cryo-transmission electron-microscopy. Biophysical Journal 56, 669681.CrossRefGoogle ScholarPubMed
Viriyaroj, A., Kashiwagi, H. & Ueno, M. (2005). Process of destruction of large unilamellar vesicles by a zwitterionic detergent, CHAPS: partition behavior between membrane and water phases. Chemical and Pharmaceutical Bulletin 53, 11401146.CrossRefGoogle ScholarPubMed
Vogel, S. S., Leikina, E. A. & Chernomordik, L. V. (1993). Lysophosphatidylcholine reversibly arrests exocytosis and viral fusion at a stage between triggering and membrane merger. Journal of Biological Chemistry 268, 2576425768.CrossRefGoogle Scholar
Volke, F. & Pampel, A. (1995). Membrane hydration and structure on a subnanometer scale as seen by high resolution solid state nuclear magnetic resonance: POPC and POPC/C12EO4 model membranes. Biophysical Journal 68, 19601965.CrossRefGoogle ScholarPubMed
Walter, A., Kuehl, G., Barnes, K. & VanderWaerdt, G. (2000). The vesicle-to-micelle transition of phosphatidylcholine vesicles induced by nonionic detergents: effects of sodium chloride, sucrose and urea. Biochimica et Biophysica Acta 1508, 2033.CrossRefGoogle ScholarPubMed
Walter, A., Vinson, P. K., Kaplun, A. & Talmon, Y. (1991). Intermediate structures in the cholate–phosphatidylcholine vesicle micelle transition. Biophysical Journal 60, 13151325.CrossRefGoogle ScholarPubMed
Walter, A., Vinson, P. K. & Talmon, Y. (1990). Cryo-TEM reveals structural transitions of egg pc and sodium cholate mixtures. Biophysical Journal 57, A476A476.Google Scholar
Wang, H., Nieh, M. P., Hobbie, E. K., Glinka, C. J. & Katsaras, J. (2003). Kinetic pathway of the bilayered-micelle to perforated-lamellae transition. Physical Review E 67, art. no.-060902.CrossRefGoogle ScholarPubMed
Wassall, S. R., Brzustowicz, M. R., Shaikh, S. R., Cherezov, V., Caffrey, M. & Stillwell, W. (2004). Order from disorder, corralling cholesterol with chaotic lipids: the role of polyunsaturated lipids in membrane raft formation. Chemistry and Physics of Lipids 132, 7988.Google ScholarPubMed
Wei, T., Geijer, S., Lindberg, M., Berne, B. & Torma, H. (2006). Detergents with different chemical properties induce variable degree of cytotoxicity and mRNA expression of lipid-metabolizing enzymes and differentiation markers in cultured keratinocytes. Toxicology in Vitro 20, 13871394.CrossRefGoogle ScholarPubMed
Wenk, M. & Seelig, J. (1997a). Vesicle–micelle transformation of phosphatidylcholine/octyl-β-d-glucopyranoside mixtures as detected with titration calorimetry. Journal of Physical Chemistry B 101, 52245231.CrossRefGoogle Scholar
Wenk, M. R., Alt, T., Seelig, A. & Seelig, J. (1997). Octyl-beta-d-glucopyranoside partitioning into lipid bilayers: thermodynamics of binding and structural changes of the bilayer. Biophysical Journal 72, 17191731.CrossRefGoogle ScholarPubMed
Wenk, M. R. & Seelig, J. (1997b). Interaction of octyl-beta-thioglucopyranoside with lipid membranes. Biophysical Journal 73, 25652574.CrossRefGoogle ScholarPubMed
Wieprecht, T., Apostolov, O., Beyermann, M. & Seelig, J. (2000). Membrane binding and pore formation of the antibacterial peptide PGLa: thermodynamic and mechanistic aspects. Biochemistry 39, 442452.CrossRefGoogle ScholarPubMed
Wieprecht, T., Beyermann, M. & Seelig, J. (1999). Binding of antibacterial magainin peptides to electrically neutral membranes: thermodynamics and structure. Biochemistry 38, 1037710387.CrossRefGoogle ScholarPubMed
Wieprecht, T., Beyermann, M. & Seelig, J. (2002). Thermodynamics of the coil-alpha-helix transition of amphipathic peptides in a membrane environment: the role of vesicle curvature. Biophysical Chemistry 96, 191201.CrossRefGoogle Scholar
Wieslander, A., Christiansson, A., Rilfors, L. & Lindblom, G. (1980). Lipid bilayer stability in membranes. Regulation of lipid composition in Acholeplasma laidlawii as governed by molecular shape. Biochemistry 19, 36503655.CrossRefGoogle ScholarPubMed
Wieslander, A., Rilfors, L. & Lindblom, G. (1986). Metabolic changes of membrane lipid composition in Acholeplasma laidlawii by hydrocarbons, alcohols, and detergents: arguments for effects on lipid packing. Biochemistry 25, 75117517.CrossRefGoogle ScholarPubMed
Williams, A. C. & Barry, B. W. (2004). Penetration enhancers. Advanced Drug Delivery Reviews 56, 603618.CrossRefGoogle ScholarPubMed
Wimley, W. C. & White, S. H. (1993). Membrane partitioning: distinguishing effects from the hydrophobic effect. Biochemistry 32, 63076312.CrossRefGoogle ScholarPubMed
Winter, R., Erbes, J., Templer, R. H., Seddon, J. M., Syrykh, A., Warrender, N. A. & Rapp, G. (1999). Inverse bicontinuous cubic phases in fatty acid/phosphatidylcholine mixtures: the effects of pressure and lipid composition. Physical Chemistry Chemical Physics 1, 887893.CrossRefGoogle Scholar
Wrenn, S. P., Gudheti, M., Veleva, A. N., Kaler, E. W. & Lee, S. P. (2001). Characterization of model bile using fluorescence energy transfer from dehydroergosterol to dansylated lecithin. Journal of Lipid Research 42, 923934.CrossRefGoogle ScholarPubMed
Yandek, L. E., Pokorny, A., Floren, A., Knoelke, K., Langel, U. & Almeida, P. F. F. (2007). Mechanism of the cell-penetrating peptide transportan 10 permeation of lipid bilayers. Biophysical Journal 92, 24342444.CrossRefGoogle ScholarPubMed
Yang, L., Harroun, T. A., Weiss, T. M., Ding, L. & Huang, H. W. (2001). Barrel-stave model or toroidal model? A case study on melittin pores. Biophysical Journal 81, 14751485.CrossRefGoogle ScholarPubMed
Yegutkin, G. G. (1997). Effect of increasing concentrations of non-ionic detergent Triton X-100 on solubilization and structure of rat liver and adipose plasma membranes. Membrane and Cell Biology 10, 515520.Google Scholar
Yeliseev, A. A., Wong, K. K., Soubias, O. & Gawrisch, K. (2005). Expression of human peripheral cannabinoid receptor for structural studies. Protein Science 14, 26382653.CrossRefGoogle ScholarPubMed
Yu, J., Fishman, D. A. & Steck, T. L. (1973). Selective solubilization of proteins and phospholipids from red blood cell membranes by nonionic detergents. Journal of Supramolecular Structures 3, 233247.CrossRefGoogle Scholar