Hostname: page-component-8448b6f56d-qsmjn Total loading time: 0 Render date: 2024-04-16T21:32:26.989Z Has data issue: false hasContentIssue false

Specificity and promiscuity in membrane helix interactions

Published online by Cambridge University Press:  17 March 2009

Mark A. Lemmon
Affiliation:
Department of Pharmacology, New York University Medical Center, 550 First Avenue, New York, NY 10016, USA
Donald M. Engelman
Affiliation:
Department of Molecular Biophysics & Biochemistry, Yale University, 266 Whitney Avenue, New Haven, CT 06520, USA

Extract

The membrane-spanning portions of many integral membrane proteins consist of one or a number of transmembrane α-helices, which are expected to be independently stable on thermodynamic grounds. Side-by-side interactions between these transmembrane α-helices are important in the folding and assembly of such integral membrane proteins and their complexes. In considering the contribution of these helix–helix interactions to membrane protein folding and oligomerization, a distinction between the energetics and specificity should be recognized. A number of contributions to the energetics of transmembrane helix association within the lipid bilayer will be relatively non-specific, including those resulting from charge–charge interactions and lipid–packing effects. Specificity (and part of the energy) in transmembrane α-helix association, however, appears to rely mainly upon a detailed stereochemical fit between sets of dynamically accessible states of particular helices. In some cases, these interactions are mediated in part by prosthetic groups.

Type
Research Article
Copyright
Copyright © Cambridge University Press 1994

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

REFERENCES

Adair, B. D. (1993). An investigation into the effects of packing on the folding of integral membrane proteins. Ph.D. Thesis, Yale University, New Haven, CT., U.S.A.Google Scholar
Adair, B. D. & Engelman, D. M. (1994). Glycophorin A helical transmembrane domains dimerize in phospholipid bilayers: A fluorescence energy transfer study. Biochemistry.(In the Press.)CrossRefGoogle Scholar
Allen, L. C. (1975). A model for the hydrogen bond. Proc. natn. Acad. Sci. U.S.A. 72, 47014705.CrossRefGoogle Scholar
Allen, J. P., Feher, G., Yeates, T. O., Komiya, H. & Rees, D. C. (1987). Structure of the reaction center from Rhodobacter sphaeroides R-26: the protein subunits. Proc. natn. Acad. Sci. U.S.A. 84, 61626166.CrossRefGoogle ScholarPubMed
Ames, P. & Parkinson, J. S. (1988). Transmembrane signaling by bacterial chemoreceptors: E. coli transducers with locked output. Cell 55, 817826.CrossRefGoogle ScholarPubMed
Aoki, D., Lee, N., Yamaguchi, N., Dubois, C. & Fukuda, M. N. (1992). Golgi retention of a trans-Golgi membrane protein, galactosyltransferase, requires cysteine and histidine residues within the membrane-anchoring domain. Proc. natn. Acad. Sci. U.S.A. 89, 43194323.CrossRefGoogle ScholarPubMed
Argos, P., Rao, J. K. M. & Hargrave, P. A. (1982). Structural prediction of membranebound proteins. Eur.J. Biochem. 128, 565575.CrossRefGoogle ScholarPubMed
Armstrong, J. & Patel, S. (1991). The Golgi sorting domain of coronavirus Ei protein. J. Cell. Sci. 98, 567575.CrossRefGoogle Scholar
Bargmann, C. I., Hung, M.-C. & Weinberg, R. A. (1986 a). Multiple independent activations of the neu oncogene by a point mutation altering the transmembrane domain of p185. Cell 45, 649657.CrossRefGoogle ScholarPubMed
Bargmann, C. I., Hung, M.-C. & Weinberg, R. A. (1986 b). The neu oncogene encodes an epidermal growth factor receptor-related protein. Nature, Lond. 319, 226230.CrossRefGoogle ScholarPubMed
Bargmann, C. I. & Weinberg, R. A. (1988 a). Oncogenic activation of the new-encoded receptor protein by point mutation and deletion. EMBO J. 7, 20432052.CrossRefGoogle ScholarPubMed
Bargmann, C. I. & Weinberg, R. A. (1988 b). Increased tyrosine kinase activity associated with the protein encoded by the activated neu oncogene. Proc. natn. Acad. Sci. U.S.A. 85, 53945398.CrossRefGoogle ScholarPubMed
Barrow, G. M. & Yerger, E. H. (1954). The dimerization of acetic acid in carbon tetrachloride and chloroform. J. Am. chem. Soc. 76, 52485249.CrossRefGoogle Scholar
Barsukov, I. L., Nolde, D. E., Lomize, A. L. & Arseniev, A. S. (1992). Threedimensional structure of proteolytic fragment 163–231 of bacterioopsin determined from nuclear magnetic resonance in solution. Eur. J. Biochem. 206, 665672.CrossRefGoogle ScholarPubMed
Ben-Efraim, I., Bach, D. & Shai, Y. (1993). Spectroscopic and functional characterization of the putative transmembrane segment of the min K potassium channel. Biochemistry 32, 23712377.CrossRefGoogle Scholar
Bernstein, H. H., Poritz, M. A., Strub, K., Hoben, P. J., Brenner, S. & Walter, P. (1989). Model for signal sequence recognition from amino-acid sequence of 54K subunit of signal recognition particle. Nature, Lond. 340, 482486.CrossRefGoogle ScholarPubMed
Betz, H. (1990). Homology and analogy in transmembrane channel design: Lessons from synaptic membrane proteins. Biochemistry 29, 35913599.CrossRefGoogle ScholarPubMed
Bibi, E. & Kaback, H. R. (1990). In vivo expression of the lac Y gene in two segments leads to functional lac permease. Proc. natn. Acad. Sci. U.S.A. 87, 43254329.CrossRefGoogle Scholar
Blond-Elguindi, S., Cwiria, S. E., Dower, W. J., Lipshutz, R. J., Sprang, S. R., Sambrook, J. F. & Gething, M.-J. H. (1993). Affinity panning of a library of peptides displayed on bacteriophages reveals the binding specificity of BiP. Cell 75, 717728.CrossRefGoogle ScholarPubMed
Blum, J. H., Stevens, T. L. & Defranco, A. L. (1993). Role of the μ immunoglobulin heavy chain transmembrane and cytoplasmic domains in B-cell antigen receptor expression and signal transduction. J. biol. Chem. 268, 2723627245.CrossRefGoogle ScholarPubMed
Bonifacino, J. S., Suzuki, C. K., Lippincott-Schwartz, J., Weissman, A. M. & Klausner, R. D. (1989). Pre-Golgi degradation of newly synthesized T-cell antigen receptor chains: Intrinsic sensitivity and the role of subunit assembly. J. Cell Biol. 109, 7383.CrossRefGoogle ScholarPubMed
Bonifacino, J. S., Suzuki, C. K. & Klausner, R. D. (1990 a). A peptide sequence confers retention and rapid degradation in the endoplasmic reticulum. Science, N. Y. 247, 7982.CrossRefGoogle ScholarPubMed
Bonifacino, J. S., Cosson, P. & Klausner, R. D. (1990 b). Co-localized transmembrane determinants for ER degradation and subunit assembly explain the intracellular fate of TCR chains. Cell, 63, 503513.CrossRefGoogle Scholar
Bonifacino, J. S., Cosson, P., Shah, N. & Klausner, R. D. (1991). Role of potentially charged transmembrane residues in targeting proteins for retention and degradation within the endoplasmic reticulum. EMBO J. 10, 27832793.CrossRefGoogle ScholarPubMed
Bormann, B.-J. & Engelman, D. M. (1992). Intramembrane helix-helix association in oligomerization and transmembrane signaling. A. Rev. Biophys. Biomol. Struct. 21, 223242.CrossRefGoogle ScholarPubMed
Bormann, B.-J., Knowles, W. J. & Marchesi, V. T. (1989). Synthetic peptides mimic the assembly of transmembrane glycoproteins. J. biol. Chem. 264, 40334037.CrossRefGoogle ScholarPubMed
Bresler, S. E. (1958). Structure, molecular forces and aggregation reactions of macromolecules of complex polymers. Discuss. Faraday Soc. 25, 158166.CrossRefGoogle Scholar
Bretscher, M. S. (1971). Major human erythrocyte glycoprotein spans the cell membrane. Nature New Biol., Lond. 231, 229232.CrossRefGoogle ScholarPubMed
Bretscher, M. S. & Munro, S. (1993). Cholesterol and the Golgi apparatus. Science, N.Y. 261, 12801281.CrossRefGoogle ScholarPubMed
Burke, J., Pettitt, J. M., Schachter, H., Sarkar, M. & Gleeson, P. A. (1992). The transmembrane and flanking sequences of beta 1, 2-N-acetylglucosaminyltransferase I specify medial-Golgi localization. J. biol. Chem. 267, 2443324440.CrossRefGoogle ScholarPubMed
Cao, H., Bangalore, L., Bormann, B.-J. & Stern, D. F. (1992 a). A subdomain in the transmembrane domain is necessary for p185neu*activation. EMBO J. 11, 923932.CrossRefGoogle Scholar
Cao, H., Bangalore, L., Dompe, C., Bormann, B.-J. & Stern, D. F. (1992 b). An extra cysteine proximal to the transmembrane domain induces differential cross-linking of p185neu and p185 neu*. J. biol. Chem. 267, 2048920492.CrossRefGoogle Scholar
Carpenter, C. D., Ingraham, H. S., Cochet, C., Walton, G. M., Lazar, C. S., Sowadski, J. M., Rosenfeld, M. G. & Gill, G. N. (1991). Structural analysis of the transmembrane domain of the epidermal growth factor receptor. J. biol. Chem. 266, 57505755.CrossRefGoogle ScholarPubMed
Cheatham, B., Shoelson, S. E., Yamada, K., Goncalves, E. & Kahn, C. R. (1993). Substitution of the erb B-2 oncoprotein transmembrane domain activates the insulin receptor and modulates the action of insulin and insulin-receptor substrate 1. Proc. natn. Acad. Sci. U.S.A. 90, 73367340.CrossRefGoogle Scholar
Chothia, C. (1984). Principles that determine the structure of proteins. A. Rev. Biochem. 53, 537572.CrossRefGoogle ScholarPubMed
Chothia, C., Levitt, M. & Richardson, D. (1981). Helix-to-helix packing in proteins. J. molec. Biol. 145, 215250.CrossRefGoogle ScholarPubMed
Cohen, C. & Parry, D. A. D. (1990). α-helical coiled-coils and bundles: How to design an a-helical protein. Proteins: Struct. Fund. & Genetics 7, 115.CrossRefGoogle Scholar
Colley, K. J., Lee, E. U. & Paulson, J. C. (1992). The signal anchor and stem regions of the β-galactoside α2, 6-sialyltransferase may each act to localize the enzyme to the Golgi apparatus. J. biol. Chem. 267, 77847793.CrossRefGoogle ScholarPubMed
Cosson, P. & Bonifacino, J. S. (1992). Role of transmembrane domain interactions in the assembly of class II MHC molecules. Science, N.Y. 258, 659662.CrossRefGoogle ScholarPubMed
Cosson, P., Lankford, S. P., Bonifacino, J. S. & Klausner, R. D. (1991). Membrane protein association by potential intramembrane charge pairs. Nature, Lond. 351, 414416.CrossRefGoogle ScholarPubMed
Cowan, S. W., Schirmer, T., Rummel, G., Steirt, M., Ghosh, R., Pauptit, R. A., Joansonius, J. N. & Rosenbusch, J. P. (1992). Crystal structures explain the functional properties of two E. coli porins. Nature, Lond. 358, 727734.CrossRefGoogle ScholarPubMed
Crick, F. H. C. (1952). Is α-keratin a coiled coil? Nature, Lond. 170, 882883.CrossRefGoogle ScholarPubMed
Crick, F. H. C. (1953). The packing of α-helices: Simple coiled-coils. Acta crystallogr. 6, 689697.CrossRefGoogle Scholar
Deber, C. M., Khan, A. R., LI, Z., Joensson, C., Glibowicka, M. & Wang, J. (1993). Val to Ala mutations selectively alter helix-helix packing in the transmembrane segment of phage M13 coat protein. Proc. natn. Acad. Sci. U.S.A. 90, 1164811652.CrossRefGoogle ScholarPubMed
Deisenhofer, J., Epp, O., Miki, K., Huber, R. & Michel, H. (1985). Structure of the protein subunits in the photosynthetic reaction centre of Rhodopsendomonas viridis at 3 Å resolution. Nature, Lond. 318, 618624.CrossRefGoogle Scholar
Deisenhofer, J. & Michel, H. (1991). High-resolution structures of photosynthetic reaction centers. A. Rev. Biophys. Biophys. Chem. 20, 247266.CrossRefGoogle ScholarPubMed
Dohnal, J. C., Potempa, L. A. & Garvin, J. E. (1980). The molecular weights of three forms of glycophorin A in sodium dodecyl sulfate solution. Biochim. biophys. Acta. 621, 255264.CrossRefGoogle ScholarPubMed
Downward, J., Yarden, Y., Mayes, E., Scrace, G., Totty, N., Stockwell, P., Ullrich, A., Schlessinger, J. & Waterfield, M. D. (1984). Close similarity of epidermal growth factor receptor and v-erb-B oncogene protein sequences. Nature, Lond. 307, 521527.CrossRefGoogle ScholarPubMed
Dunten, R. L., Sahin-Toth, M. & Kaback, H. R. (1993). Role of the charge pair aspartic acid-237-lysine-358 in the lactose permease of Escherichia coli. Biochemistry 32, 31393145.CrossRefGoogle ScholarPubMed
Edelman, J. (1993). Quadratic minimization of predictors for protein secondary structure: Application to transmembrane α-helices. J. molec. Biol. 232, 165191.CrossRefGoogle ScholarPubMed
Eisenberg, D. (1984). Three dimensional structure of membrane and surface proteins. A. Rev. Biochem. 53, 595623.CrossRefGoogle ScholarPubMed
Eisenberg, D. & McLachlan, A. D. (1986). Solvation energy in protein folding and binding. Nature, Lond. 319, 199203.CrossRefGoogle ScholarPubMed
Engelman, D. M. & Zaccaï, G. (1980). Bacteriorhodopsin is an inside-out protein. Proc. natn. Acad. Sci. U.S.A. 77, 58945898.CrossRefGoogle ScholarPubMed
Engelman, D. M., Henderson, R., McLachlan, A. D. & Wallace, B. A. (1980). Path of the polypeptide in bacteriorhodopsin. Proc. natn. Acad. Sci. U.S.A. 77, 20232027.CrossRefGoogle ScholarPubMed
Engelman, D. M. & Steitz, T. A. (1981). The spontaneous insertion of proteins into and across membranes: The helical hairpin hypothesis. Cell 23, 411422.CrossRefGoogle ScholarPubMed
Engelman, D. M. (1982). An implication of the structure of bacteriorhodopsin. Globular membrane proteins are stabilized by polar interactions. Biophys. J. 37, 187188.CrossRefGoogle ScholarPubMed
Engelman, D. M., Steitz, T. A. & Goldman, A. (1986). Identifying nonpolar transbilayer helices in amino acid sequences of membrane proteins. A. Rev. Biophys. Biophys. Chem. 15, 321353.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 Å resolution. Nature, Lond. 300, 325330.CrossRefGoogle ScholarPubMed
Fraga, D., Hermolin, J. & Fillingame, R. H. (1994). Transmembrane helix-helix interactions in Fo suggested by suppressor mutations to Ala24 → Asp/Asp61 → Gly mutant of ATP synthase subunit c. J. biol. Chem. 269, 25622567.CrossRefGoogle Scholar
Frattali, A. L., Treadway, J. L. & Pessin, J. E. (1991). Evidence supporting a passive role for the insulin receptor transmembrane domain in insulin-dependent signal transduction. J. biol. Chem. 266, 98299834.CrossRefGoogle ScholarPubMed
Fljii, J., Maruyama, K., Tada, M. & MacLennan, D. H. (1989). Expression and sitespecific mutagenesis of phospholamban: Studies of residues involved in phosphorylation and pentamer formation. J. biol. Chem. 264, 1295012955.Google Scholar
Flrthmayr, H. & Marchesi, V. T. (1976). Subunit structure of human erythrocyte glycophorin A. Biochemistry 15, 11371144.CrossRefGoogle Scholar
Gazit, E. & Shai, Y. (1993). Structural characterization, membrane insertion, and specific assembly within phospholipid membranes of hydrophobic segments from Bacillus thuringiensis var. israelensis cytolytic toxin. Biochemistry 32, 1236312371.CrossRefGoogle Scholar
Gerber, G. E., Anderegg, R. J., Herlihy, W. C., Gray, C. P., Beimann, K. & Khorana, H. G. (1979). Partial primary structure of bacteriorhodopsin: Sequencing methods for membrane proteins. Proc. natn. Acad. Sci. U.S.A. 76, 227231.CrossRefGoogle ScholarPubMed
Gerwert, K., Hess, B., Soppa, J. & Oesterhelt, D. (1989). Role of aspartate-96 in proton translocation by bacteriorhodopsin. Proc. natn. Acad. Sci. U.S.A. 86,49434947.CrossRefGoogle ScholarPubMed
Gierasch, L. M. (1989). Signal sequences. Biochemistry 28, 923930.CrossRefGoogle ScholarPubMed
Gilles-Gonzalez, M. A., Engelman, D. M. & Khorana, H. G. (1991). Structurefunction studies of bacteriorhodopsin XV: Effects of deletions in loops B–C and E–F on bacteriorhodopsin chromophore and structure. J. biol. Chem. 266, 85458550.CrossRefGoogle ScholarPubMed
Girvin, M. E. & Fillingame, R. H. (1993). Helical structure and folding of subunit c of F1 F0 ATP synthase: 1H NMR resonance assignments and nOe analysis. Biochemistry 32, 1216712177.CrossRefGoogle Scholar
Girvin, M. E. & Fillingame, R. H. (1994). Hairpin folding of subunit c of F1 F0 ATP synthase: 1H distance measurements to nitroxide-derivatized aspartyl-61. Biochemistry, 33, 665674.CrossRefGoogle Scholar
Goldstein, D. J., Kulke, R., Dimaio, D. & Schlegel, R. (1992 a). A glutamine residue in the membrane-associating domain of the bovine papillomavirus type 1 E5 oncoprotein mediates its binding to a transmembrane component of the vacuolar H+ ATPase. J. Virol. 66, 405413.CrossRefGoogle ScholarPubMed
Goldstein, D. J., Andersson, T., Sparkowski, J. J. & Schlegel, R. (1992 b). The BPV-1 E5 protein, the 16 kDa membrane pore-forming protein and the PDGF receptor exist in a complex that is dependent on hydrophobic transmembrane interactions. EMBO J. 11, 48514859.CrossRefGoogle Scholar
Havelka, W. A., Henderson, R., Heymann, J. A. W. & Oesterhelt, D. (1993). Projection structure of halorhodopsin from Halobacterium halobium at 6 Å resolution obtained by electron cryo-microscopy. J. molec. Biol. 234, 837846.CrossRefGoogle ScholarPubMed
Hempstead, B. L., Martin-Zanca, D., Kaplan, D. R., Parada, L. F. & Chao, M. V. (1991). High-affinity NGF binding requires coexpression of the trk proto-oncogene and the low-affinity NGF receptor. Nature, Lond. 350, 678683.CrossRefGoogle ScholarPubMed
Henderson, R. & Unwin, P. N. T. (1975). Three-dimensional model of purple membrane obtained by electron microscopy. Nature, Lond. 257, 2832.CrossRefGoogle ScholarPubMed
Henderson, R., Baldwin, J. M., Ceska, T. A., Zemlin, F., Beckmann, E. & Downing, K. H. (1990). Model for the structure of bacteriorhodopsin based on high-resolution electron cryo-microscopy. J. molec. Biol. 213, 899929.CrossRefGoogle ScholarPubMed
Hendrickson, W. A. (1992). Modes of transduction. Curr. Biol. 2, 5759.CrossRefGoogle ScholarPubMed
Hennecke, S. & Cosson, P. (1993). Role of transmembrane domains in assembly and intracellular transport of the CD8 molecule. J. biol. Chem. 268, 2660726612.CrossRefGoogle ScholarPubMed
Henry, G. D. & Sykes, B. D. (1990). Detergent-solubilized M13 coat protein exists as an asymmetric dimer. J. molec. Biol. 212, 1114.CrossRefGoogle ScholarPubMed
Henry, G. D. & Sykes, B. D. (1992). Assignment of amide 1H and 15N NMR resonances in detergent-solubilized M13 coat protein: A model for the coat protein dimer. Biochemistry 31, 52845291.CrossRefGoogle ScholarPubMed
Honig, B. H. & Hubbell, W. L. (1984). Stability of salt bridges in membrane proteins. Proc. natn. Acad. Sci. U.S.A. 81, 54125416.CrossRefGoogle ScholarPubMed
Huang, K.-S., Bayley, H., Liao, M.-J., London, E. & Khorana, H. G. (1981). Refolding of an integral membrane protein: Denaturation, renaturation, and reconstitution of intact bacteriorhodopsin and two proteolytic fragments. J. biol. Chem. 256, 38023809.CrossRefGoogle ScholarPubMed
Hunt, J. F., Bousche, O., Earnest, T. N., Kalghati, K., Reilly, K., Horvath, C., Rothschild, K. J. & Engelman, D. M. (1994). A biophysical dissection of a prototypical α-helical integral membrane protein. (In preparation.)Google Scholar
Hurwitz, D. R., Emanuel, S. L., Nathan, M. H., Sarver, N., Ullrich, A., Felder, S., Lax, I. & Schlessinger, J. (1991). EGF induces increased ligand binding affinity and dimerization of soluble epidermal growth factor receptor extracellular domain. J. biol. Chem. 266, 2203522043.CrossRefGoogle Scholar
Jähnig, F. (1983). Thermodynamics and kinetics of protein incorporation into membranes. Proc. natn. Acad. Sci. U.S.A. 80, 36913695.CrossRefGoogle ScholarPubMed
Jakab, G. & Kranias, E. G. (1988). Phosphorylation and dephosphorylation of purified phospholamban and associated phosphatidylinositides. Biochemistry 27, 37993806.CrossRefGoogle ScholarPubMed
Jeffery, C. J. & Koshland, D. E. Jr., (1994 a). A single hydrophobic to hydrophobic substitution in the transmembrane domain impairs aspartate receptor function. (Submitted.)CrossRefGoogle Scholar
Jeffery, C. J. & Koshland, D. E. Jr., (1994 a). A single hydrophobic to hydrophobic substitution in the transmembrane domain impairs aspartate receptor function. Biochemistry 33, 34583463.CrossRefGoogle ScholarPubMed
Jeffery, C. J. & Koshland, D. E. Jr., (1994 b). The E. coli aspartate receptor: Effects of random amino acid substitutions in the second transmembrane domain. (Submitted.)Google Scholar
Kaback, H. R. (1992). In and out and up and down with lac permease. International Review of Cytology 137A (ed. Jeon, K. W. and Friedlander, M.) pp. 97125. New York: Academic Press.Google Scholar
Kaback, H. R., Jung, K., Jung, H., Wu, J., Prive, G. G. & Zen, K. (1993). What's new with lactose permease ?. J. Bioenerget. Biomemb. 25. (In the Press.)Google Scholar
Kahn, T. W., Sturtevant, J. M. & Engelman, D. M. (1992). Thermodynamic measurements of the contributions of helix-connecting loops and of retinal to the stability of bacteriorhodopsin. Biochemistry 31, 88298839.CrossRefGoogle Scholar
Kashles, O., Szapary, D., Bellot, F., Ullrich, A., Schlessinger, J. & Schmidt, A. (1988). Ligand-induced stimulation of epidermal growth factor receptor mutants with altered transmembrane regions. Proc. natn. Acad. Sci. U.S.A. 85, 95679571.CrossRefGoogle ScholarPubMed
Kashles, O., Yarden, Y., Fischer, R., Ullrich, A. & Schlessinger, J. (1991). A dominant negative mutation suppresses the function of normal epidermal growth factor receptors by heterodimerization. Molec. Cell. Biol. 11, 14541463.Google ScholarPubMed
Kataoka, M., Kahn, T. W., Tsujiuchi, T., Engelman, D. M. & Tokunaga, F. (1992). Bacteriorhodopsin reconstituted from two individual helices and the complementary five-helix fragment is photoactive. Photochem. Photobiol. 56, 895901.CrossRefGoogle ScholarPubMed
Kim, S.-H. (1994). ‘Frozen’ dynamic dimer model for transmembrane signaling in bacterial chemotaxis receptors. Protein Science 3, 159165.CrossRefGoogle ScholarPubMed
Kinet, J. P. (1990). Antibody-cell interactions: Fc receptors. Cell 57, 351354.CrossRefGoogle Scholar
Klausner, R. D., Lippincott-Schwartz, J. & Bonifacino, J. S. (1990). The T-cell antigen receptor: Insights into organelle biology. A. Rev. Cell Biol. 6, 403431.CrossRefGoogle ScholarPubMed
Klausner, R. D. & Sitia, R. (1990). Protein degradation in the endoplasmic reticulum. Cell 62, 611614.CrossRefGoogle ScholarPubMed
Kobilka, B. K., Kobilka, T. S., Daniel, K., Regan, J. W., Caron, M. G. & Lefkowitz, R. J. (1988). Chimeric α1,-β2-adrenergic receptors: Delineation of domains involved in effector coupling and ligand binding specificity. Science, N. Y. 240, 13101316.CrossRefGoogle Scholar
Köster, W. & Braun, V. (1990). Iron (III) hydroxamate transport of Escherichia coli:Restoration of iron supply by coexpression of the N- and C-terminal halves of the cytoplasmic membrane protein Fhu B cloned on separate plasmids. Molec. gen. Genet. 223, 379384.CrossRefGoogle Scholar
Kovacs, R. J., Nelson, M. T., Simmerman, H. K. B. & Jones, L. R. (1988). Phospholamban forms Ca2+-selective channels in lipid bilayers. J. biol. Chem. 263, 1836418368.CrossRefGoogle ScholarPubMed
Krauss, N., Hinrichs, W., Witt, I., Fromme, P., Pritzkow, W., Dauter, Z., Betzel, C., Wilson, K., Witt, H. T. & Saenger, W. (1993). Three-dimensional structure of system I of photosynthesis at 6 A resolution. Nature, Lond. 361, 326331.CrossRefGoogle Scholar
Kühlbrandt, W. & Wang, D. N. (1991). Three-dimensional structure of plant lightharvesting complex determined by electron crystallography. Nature, Lond. 350, 130134.CrossRefGoogle ScholarPubMed
Kühlbrandt, W., Wang, D. N. & Fujiyoshi, Y. (1994). Atomic model of plant lightharvesting complex. Nature, Lond. 367, 614621.CrossRefGoogle ScholarPubMed
Kulke, R., Horwitz, B. H., Zibello, T. & Dimaio, D. (1992). The central hydrophobic domain of the bovine papillomavirus E5 transforming protein can be functionally replaced by many hydrophobic amino acid sequences containing a glutamine. J. Virol. 66, 505511.CrossRefGoogle ScholarPubMed
Kurosaki, T., Gander, I. & Ravetch, J. V. (1991). A subunit common to an IgG Fc receptor and the T-cell receptor mediates assembly through different interactions. Proc. natn. Acad. Sci. U.S.A. 88, 38373841.CrossRefGoogle Scholar
Kurosaki, T. & Ravetch, J. V. (1989). A single amino acid in the glycosyl phosphatidylinositol attachment domain determines the membrane topology of FcyR III. Nature, Lond. 342, 805807.CrossRefGoogle Scholar
Kyte, J. & Doolittle, R. F. (1982). A simple method for displaying the hydropathic character of a protein. J. molec. Biol. 157, 105132.CrossRefGoogle ScholarPubMed
Langosch, D., Hartung, K., Grell, E., Bamberg, E. & Betz, H. (1991). Ion channel formation by synthetic transmembrane segments of the inhibitory glycine receptor - a model study. Biochim. biophys. Acta 1063, 3644.CrossRefGoogle ScholarPubMed
Lanier, L. L., Yu, G. & Phillips, J. H. (1989). Co-association of CD3ζwith a receptor (CD 16) for IgG Fc on human natural killer cells. Nature, Lond. 342, 803805.CrossRefGoogle Scholar
Lanier, L. L., Yu, G. & Phillips, J. H. (1991). Analysis of FcyR II (CD16) membrane expression and association with CD3ζ and FcεRI-γ by site-directed mutation. J. Immun. 146, 15711576.CrossRefGoogle Scholar
Lankford, S. P., Cosson, P., Bonifacino, J. S. & Klausner, R. D. (1993). Transmembrane domain length affects charge-mediated retention and degradation of proteins within the endoplasmic reticulum. J. biol. Chem. 268, 48144820.CrossRefGoogle ScholarPubMed
Large, T. H., Weskamp, G., Helder, J. C., Radeke, M. J., Misko, T. P., Shooter, E. M. & Reichardt, L. F. (1989). Structure and developmental expression of the nerve growth factor receptor in the chicken central nervous system. Neuron 2, 11231134.CrossRefGoogle ScholarPubMed
Lax, I., Mitra, A. K., Ravera, C., Hurwitz, D. R., Rubinstein, M., Ullrich, A., Stroud, R. M. & Schlessinger, J. (1991). Epidermal growth factor (EGF) induces oligomerization of soluble, extracellular, ligand-binding domain of EGF receptor. J. biol. Chem. 266, 1382813833.CrossRefGoogle ScholarPubMed
Lee, A. W. & Nienhuis, A. W. (1992). Functional dissection of structural domains in the receptor for colony-stimulating factor-1. J. biol. Chem. 267, 1647216483.CrossRefGoogle ScholarPubMed
Lee, B. & Richards, F. M. (1971). The interpretation of protein structures: Estimation of static accessibility. J. molec. Biol. 55, 379400.CrossRefGoogle ScholarPubMed
Lee, J., Dull, T. J., Lax, I., Schlessinger, J. & Ullrich, A. (1989). HER2 cytoplasmic domain generates normal mitogenic and transforming signals in a chimeric receptor. EMBO J. 8, 167173.CrossRefGoogle Scholar
Lehväslaiho, H., Lehtola, L., Sistonen, L. & Alitalo, K. (1989). A chimeric EGFR/neu proto-oncogene allows EGF to regulate neu tyrosine kinase and cell transformation. EMBO J. 8, 159166.CrossRefGoogle Scholar
Lemmon, M. A., Flanagan, J. M., Hunt, J. F., Adair, B. D., Bormann, B.-J., Dempsey, C. E. & Engelman, D. M. (1992 a). Glycophorin A dimerization is driven by specific interactions between transmembrane a-helices. J. biol. Chem. 267, 76837689.CrossRefGoogle Scholar
Lemmon, M. A., Flanagan, J. M., Treutlein, H. R., Zhang, J. & Engelman, D. M. (1992 b). Sequence-specific dimerization of transmembrane α-helices. Biochemistry, 31, 1271912725.CrossRefGoogle Scholar
Lemmon, M. A. (1993). Sequence-specific oligomerization of transmembrane α-helices: Their role in the oligomerization of integral membrane proteins. Ph.D. Thesis, Yale University, New Haven, CT, USA.Google Scholar
Lemmon, M. A., Treutlein, H. R., Adams, P., Brunger, A. T. & Engelman, D. M. (1994). A dimerization motif for transmembrane α-helices. Nature Struct. Biol. 1, 157163.CrossRefGoogle ScholarPubMed
Lenard, J. & Singer, S. J. (1966). Protein conformation in cell membrane preparations as studied by optical rotatory dispersion and circular dichroism. Proc. natn. Acad. Sci. U.S.A. 56, 18281835.CrossRefGoogle ScholarPubMed
Li, Z., Glibowicka, M., Joensson, C. & Deber, C. M. (1992). Conformational states of mutant M13 coat proteins are regulated by transmembrane residues.J. biol. Chem. 268, 45844587.CrossRefGoogle Scholar
Liao, M. J., London, E. & Khorana, H. G. (1983). Regeneration of native bacteriorhodopsin from two chymotryptic fragments. J. biol. Chem. 258, 99499955.CrossRefGoogle ScholarPubMed
Liao, M.J., Huang, K. S. & Khorana, H. G. (1984). Regeneration of native bacteriorhodopsin structure from fragments. J. biol. Chem. 259, 42004204.CrossRefGoogle ScholarPubMed
Lofts, F. J., Hurst, H. C., Sternberg, M. J. E. & Gullick, W. J. (1993). Specific short transmembrane sequences can inhibit transformation by the mutant neu growth factor receptor in vitro and in vivo. Oncogene 8, 28132820.Google ScholarPubMed
Longo, N., Shuster, R. C., Griffin, L. D., Langley, S. D. & Elsas, L. J. (1992). Activation of insulin receptor signaling by a single amino-acid substitution in the transmembrane domain. J. biol. Chem. 267, 1241612419.CrossRefGoogle ScholarPubMed
Lynch, B. A. & Koshland, D. E. (1991). Disulfide cross-linking studies of the transmembrane regions of the aspartate sensory receptor of Escherichia coli. Proc. natn. Acad. Sci. U.S.A. 88, 1040210406.CrossRefGoogle ScholarPubMed
Machamer, C. E. & ROSE, J. K. (1987). A specific transmembrane domain of a coronavirus Ei glycoprotein is required for its retention in the Golgi region. J. Cell Biol. 105, 12051214.CrossRefGoogle Scholar
Machamer, C. E. (1991). Golgi retention signals: Do membranes hold the key? Trends. Cell Biol. 1, 141144.CrossRefGoogle ScholarPubMed
Machamer, C. E., Grim, M. G., Esquela, A., Chung, S. W., Rolls, M., Ryan, K. & Swift, A. M. (1993). Retention of a cis Golgi protein requires polar residues on one face of a predicted a-helix in the transmembrane domain. Molec. Biol. Cell 4, 695704.CrossRefGoogle Scholar
Madden, D. R., Garboczi, D. N. & Wiley, D. C. (1993). The antigenic identity of peptide-MHC complexes: A comparison of the conformations of five viral peptides presented by HLA-A2. Cell 75, 693708.CrossRefGoogle ScholarPubMed
Maggio, R., Vogel, Z. & Wess, J. (1993). Reconstitution of functional muscarinic receptors by co-expression of amino- and carboxy-terminal receptor fragments. FEBS Lett. 319, 195200.CrossRefGoogle Scholar
Manolios, N., Bonifacino, J. S. & Klausner, R. D.Transmembrane helical interactions and the assembly of the T cell receptor complex. (1990). Science, N.Y. 249, 274277.CrossRefGoogle ScholarPubMed
Manolios, N., Letournier, F., Bonifacino, J. S. & Klausner, R. D. (1991). Pairwise, cooperative and inhibitory interactions describe the assembly and probable structure of the T-cell antigen receptor. EMBO J. 10, 16431651.CrossRefGoogle ScholarPubMed
Matsumoto, A. K., Kopicky-Burd, J., Carter, R. H., Tuveson, D. A., Tedder, T. T. & Fearon, D. T. (1991). Interaction of the complement and immune systems: A signal transduction complex of the B-lymphocyte containing complement receptor type 2 and CD19. J. exp. Med. 173, 5564.CrossRefGoogle ScholarPubMed
Matsumoto, A. K., Martin, D. R., Carter, R. H., Klickstein, L. B., Ahearn, J. M. & Fearon, D. T. (1993). Functional dissection o the CD21/CD19/TAPA-1/Leu-13 complex of B lymphocytes. J. exp. Med. 178, 14071417.CrossRefGoogle Scholar
McGinnes, L., Sergel, T. & Morrison, T. (1993). Mutations in the transmembrane domain of the HN protein of newcastle disease virus affect the structure and activity of the protein. Virology 196, 101110.CrossRefGoogle Scholar
Mendelsohn, R., Dluhy, R. A., Crawford, T. & Mantsch, H. H. (1984). Interaction of glycophorin with phosphatidylserine: A Fourier Transform infrared investigation. Biochemistry 23, 14981504.CrossRefGoogle ScholarPubMed
Milburn, M. V., Privé, G. G., Milligan, D. L., Scott, W. G., Yeh, J., Jancarik, J., Koshland, D. E. Jr. & Kim, S.-H. (1991). Three dimensional structures of the ligand binding domain of the bacterial aspartate receptor with and without a ligand. Science, N.Y. 254, 13421347.CrossRefGoogle ScholarPubMed
Millar, D. G. & Shore, G. C. (1993). The signal anchor sequence of mitochondrial Mas70p contains an oligomerization domain. J. biol. Chem. 268, 1840318406.CrossRefGoogle ScholarPubMed
Milligan, D. L. & Koshland, D. E. (1988). Site-directed cross-linking: Establishing the dimeric structure of the aspartate receptor of bacterial chemotaxis. J. biol. Chem. 263, 62686275.CrossRefGoogle ScholarPubMed
Moe, G. R., Bollag, G. E. & Koshland, D. E. Jr. (1989). Transmembrane signalling by a chimera of the Escherichia coli aspartate receptor and the human insulin receptor. Proc. natn. Acad. Sci. U.S.A. 86, 56835687.CrossRefGoogle ScholarPubMed
Mohammadi, M., Honegger, A., Sorokin, A., Ullrich, A., Schlessinger, J. & Hurwitz, D. R. (1993). Aggregation-induced activation of the epidermal growth factor receptor protein tyrosine kinase. Biochemistry 32, 87428748.CrossRefGoogle ScholarPubMed
Montal, M. (1990). Molecular anatomy and molecular design of channel proteins. FASEB J. 4, 26232635.CrossRefGoogle ScholarPubMed
Munro, S. (1991). Sequences within and adjacent to the transmembrane segment of α- 2, 6-sialyltransferase specify Golgi retention. EMBO J. 10, 35773588.CrossRefGoogle Scholar
Nilges, M. & Brünger, A. T. (1991). Automated modeling of coiled coils: Application to the GCN4 dimerization region. Protein Engineering 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. Fund. & Genetics 15, 133146.CrossRefGoogle Scholar
Nilsson, T., Lucocq, J. M., Mackay, D. & Warren, G. (1991). The membrane spanning domain of β1, 4-galactosyltransferase specifies trans Golgi localization. EMBO J. 10, 35673575.CrossRefGoogle ScholarPubMed
Nilsson, T., Slusarewicz, P., Hoe, M. H. & Warren, G. (1993). Kin recognition: A model for the retention of Golgi enzymes. FEBS Letts. 330, 14.CrossRefGoogle Scholar
Nilsson, T., Hoe, M. H., Slusarewicz, P., Rabouille, C., Watson, R., Hunte, F., Watzele, G., Berger, E. G. & Warren, G. (1994). Kin recognition between medial Golgi enzymes in HeLa cells. EMBO J. 13, 562574.CrossRefGoogle ScholarPubMed
Oiki, S., Danho, W. & Montal, M. (1988). Channel protein engineering: Synthetic 22- mer peptide from the primary structure of the voltage-sensitive sodium channel forms ionic channels in lipid bilayers. Proc. natn. Acad. Sci. U.S.A. 85, 23932397.CrossRefGoogle ScholarPubMed
Osborne, R. S. & Silhavy, T. J. (1993). PrlA suppressor mutations cluster in regions Corresponding to three distinct topological domains. EMBO J. 12, 33913398.CrossRefGoogle ScholarPubMed
Pakula, A. A. & Simon, M. I. (1992). Determination of transmembrane protein structure by disulfide cross-linking: The Escherichia coli Tar receptor. Proc. natn. Acad. Sci. U.S.A. 89, 41444148.CrossRefGoogle ScholarPubMed
Pelham, H. R. B. (1989). Control of protein exit from the endoplasmic reticulum. A. Rev. Cell Biol. 5, 123.CrossRefGoogle ScholarPubMed
Pelham, H. R. B. & Munro, S. (1993). Sorting of membrane proteins in the secretory pathway. Cell 75, 603605.CrossRefGoogle ScholarPubMed
Pervushin, K. V. & Arseniev, A. S. (1992). Three-dimensional structure of (1–36) bacterioopsin in methanol-chloroform mixture and SDS micelles determined by 2D 1H-NMR spectroscopy. FEBS Lett. 308, 190196.CrossRefGoogle Scholar
Petti, L., Nilson, L., Kulke, R., Leptak, C., Riese, D. J., Zibello, T. & Dimaio, D. (1991 a). The E5 mini-oncogene of bovine papillomavirus: Biological activities, genetic analysis, and proposed mechanisms of action. Origins of Human Cancer: A Comprehensive Review. Cold Spring Harbor: Cold Spring Harbor Laboratory Press, pp. 707713.Google Scholar
Petti, L., Nilson, L. A. & Dimaio, D. (1991 b). Activation of the platelet-derived growth factor receptor by the bovine papillomavirus E5 transforming protein. EMBO J. 10, 845855.CrossRefGoogle ScholarPubMed
Petti, L. & Dimaio, D. (1992). Stable association between the bovine papillomavirus E5 transforming protein and activated platelet-derived growth factor receptor in transformed mouse cells. Proc. natn. Acad. Sci. U.S.A. 89, 87368740.CrossRefGoogle ScholarPubMed
Popot, J.-L., Trewhella, J. & Engelman, D. M. (1986). Reformation of crystalline purple membrane from purified bacteriorhodopsin fragments. EMBO J. 5, 30393044CrossRefGoogle ScholarPubMed
Popot, J.-L., Gerchman, S. E. & Engelman, D. M. (1987). Refolding of bacteriorhodopsin in lipid bilayers: a thermodynamically controlled two-stage process. J. molec. Biol. 198, 655676.CrossRefGoogle ScholarPubMed
Popot, J.-L., Engelman, D. M., Gurel, O. & Zaccaï, G. (1989). Tertiary structure of bacteriorhodopsin: Positions and bientations of helices A and B in the structural map determined by neutron diffraction. J. molec. Biol. 210, 829847.CrossRefGoogle Scholar
Popot, J.-L. & Engelman, D. M. (1990). Membrane protein folding and oligomerization: The two-stage model. Biochemistry 29, 40314037.CrossRefGoogle ScholarPubMed
Popot, J.-L. & De Vitry, C. (1990). On the microassembly of integral membrane proteins. A. Rev. Biophys. Biophys. Chem. 19, 369403.CrossRefGoogle ScholarPubMed
Popot, J.-L. (1993). Integral membrane protein structure: Transmembrane α-helices as autonomous folding domains. Curr. Op. Struct. Biol. 3, 532540.CrossRefGoogle Scholar
Rapaport, D. & Shai, Y. (1992). Aggregation and organization of pardaxin in lipid bilayers: A fluorescence energy transfer study. J. biol. Chem. 267, 65026509.CrossRefGoogle ScholarPubMed
Rashin, A. A., Iofin, M. & Honig, B. H. (1986). Internal cavities and buried waters in globular proteins. Biochemistry 25, 36193625.CrossRefGoogle ScholarPubMed
Ravetch, J. V. & Kinet, J. P. (1991). Fc receptors. A. Rev. Immunol. 9, 457492.CrossRefGoogle ScholarPubMed
Redemann, N., Holzmann, B., Von Rüden, T., Wagner, E. F., Schlessinger, J. & Ullrich, A. (1992). Anti-oncogenic activity of signaling-defective epidermal growth factor receptor mutants. Molec. Cell. Biol. 12, 491498.Google Scholar
Rees, D. C., Komiya, H., Yeates, T. O., Allen, J. P. & Feher, G. (1989 a). The bacterial photosynthetic reaction center as a model for membrane proteins. A. Rev. Biochem. 58, 607633.CrossRefGoogle Scholar
Rees, D. C., DeAntonio, L. & Eisenberg, D. (1989 b). Hydrophobic organization of membrane proteins. Science, N.Y. 245, 510513.CrossRefGoogle ScholarPubMed
Rees, D. C., Chirino, A. J., Kim, K.-H. & Komita, H. (1994). Membrane protein structure and stability: Implications of the first crystallographic analyses. In Membrane Protein Structure: An Experimental Approach. (ed. White, S. H.). Oxford: Oxford University Press. (In the Press.)Google Scholar
Riedel, H., Dull, T. J., Schlessinger, J. & Ullrich, A. (1986). A chimeric receptor allows insulin to stimulate tyrosine kinase activity of epidermal growth factor receptors. Nature, Lond. 324, 6870.CrossRefGoogle Scholar
Richards, F. M. (1977). Areas, volumes, packing, and protein structure. A. Rev. Biophys. Bioeng. 6, 151176.CrossRefGoogle ScholarPubMed
Rodrigues, G. & Park, M. (1993). Dimerization mediated through a leucine-zipper activates the oncogenic potential of the met receptor tyrosine kinase. Molec. Cell. Biol. 13, 67116722.Google ScholarPubMed
Romeo, C. & Seed, B. (1991). Cellular immunity to HIV activated by CD4 fused to T-cell or Fc receptor polypeptides. Cell 64, 10371046.CrossRefGoogle ScholarPubMed
Rothman, J. E. (1989). GTP and methionine bristles. Nature, Lond. 340, 433434.CrossRefGoogle ScholarPubMed
Russo, R. N., Shaper, N. L., Taatjes, D. J. & Shaper, J. H. (1992). β1, 4-galactosyltransferase: A short NH2-terminal fragment that includes the cytoplasmic and transmembrane domain is sufficient for Golgi retention. J. biol. Chem. 267, 92419247.CrossRefGoogle Scholar
Rutledge, T., Cosson, P., Manolios, N., Bonifacino, J. S. & Klausner, R. D. (1992). Transmembrane helical interactions: zeta chain dimerization and functional association with the T-cell receptor. EMBO J. 11, 32453254.CrossRefGoogle Scholar
Saito, T., Weiss, A., Gunter, E. M. & Shevach, R. N. (1987). Cell surface T3 expression requires the presence of both γ, α and β chains of the T-cell receptor. J. Immunol. 139, 625631.CrossRefGoogle Scholar
Schechter, A. L., Stern, D. F., Vaidyanathan, L., Decker, S. J., Drebin, J. A., Greene, M. I. & Weinberg, R. A. (1984). The neu oncogene: an erb-B-related gene encoding a 185000-M r. tumour antigen. Nature, Lond. 312, 513516.CrossRefGoogle ScholarPubMed
Schertler, G. F. X., Villa, C. & Henderson, R. (1993). Projection structure of rhodopsin. Nature, Lond. 362, 770772.CrossRefGoogle ScholarPubMed
Schlessinger, J. & Ullrich, A. (1992). Growth factor signaling by receptor tyrosine kinases. Neuron 9, 383391.CrossRefGoogle ScholarPubMed
Seelig, J. & Seelig, A. (1980). Lipid conformation in model membranes and biological membranes. Q. Rev. Biophys. 13, 1961.CrossRefGoogle ScholarPubMed
Segatto, O., King, C. R., Pierce, J. H., Di Fiore, P. P. & Aaronson, S. A. (1988). Different structural alterations up-regulate in vitro tyrosine kinase activity and transforming potency of the erbB-2 gene. Molec. Cell. Biol. 8, 55705574.Google Scholar
Segrest, J. P., Kahane, I., Jackson, R. L. & Marchesi, V. T. (1973). Major glycoprotein of the human erythrocyte membrane: evidence for an amphipathic molecular structure. Archs Biochem. Biophys. 155, 167183.CrossRefGoogle ScholarPubMed
Sharp, K. A., Nicholls, A., Fine, R. F. & Honig, B. (1991 a). Reconciling the magnitude of the microscopic and macroscopic hydrophobic effects. Science, N. Y. 252, 106109.CrossRefGoogle ScholarPubMed
Sharp, K. A., Nicholls, A., Friedman, R. & Honig, B. (1991 b). Extracting hydrophobic free energies from experimental data: Relationship to protein folding and theoretical models. Biochemistry 30, 96869697.CrossRefGoogle ScholarPubMed
Shaw, A. C., Mitchell, R. N., Weaver, Y. K., Campos-Torres, J., Abbas, A. K. & Leder, P. (1990). Mutations of immunoglobulin transmembrane and cytoplasmic domains: Effects on intracellular signalling and antigen presentation. Cell 63, 381392.CrossRefGoogle ScholarPubMed
Shin, J., Lee, S. & Strominger, J. L. (1993). Translocation of TCRa chains into the lumen of the endoplasmic reticulum and their degradation. Science, N. Y. 259, 19011904.CrossRefGoogle Scholar
Showers, M. O., Demartino, J. C., Saito, Y. & D'Andrea, A. D. (1993). Fusion of the erythropoietin receptor and the Friend spleen focus-forming virus gp55 glycoprotein transforms a factor-dependent hematopoietic cell line. Molec. Cell. Biol. 13, 739748.Google ScholarPubMed
Sigrist, H., Wenger, R. H., Kislig, E. & Wüthrich, M. (1988). Refolding of bacteriorhodopsin. Eur.J. Biochem. 177, 125133.Google ScholarPubMed
Simmerman, H. K. B., Collins, J. H., Theibert, J. L., Wegener, A. D. & Jones, L. R. (1986). Sequence analysis of phospholamban: Identification of phosphorylation sites and two major structural domains. J. biol. Chem. 261, 1333313341CrossRefGoogle ScholarPubMed
Simmerman, H. K. B., Lovelace, D. E. & Jones, L. R. (1989). Secondary structure of detergent-solubilized phospholamban, a phosphorylatable, oligomeric protein of cardiac sarcoplasmic reticulum. Biochim. biophys. Acta. 997, 322329.CrossRefGoogle ScholarPubMed
Singer, S. J. (1962). The properties of proteins in nonaqueous solvents. Adv. Protein Chem. 17, 169.Google Scholar
Singer, S. J. (1990). The structure and insertion of integral proteins in membranes. A. Rev. Cell Biol. 6, 247296.CrossRefGoogle ScholarPubMed
Smith, S. & Blobel, G. (1993). The first membrane spanning region of the lamin B receptor is sufficient for sorting to the inner nuclear membrane. J. Cell Biol. 120, 631637.CrossRefGoogle Scholar
Sorokin, A., Lemmon, M. A., Ullrich, A. & Schlessinger, J. (1994). Stabilization of an active dimeric form of the Epidermal Growth Factor receptor by introduction of an inter-receptor disulfide bond. J. biol. Chem. 269, 97529759.CrossRefGoogle ScholarPubMed
Spaargaren, M., Defize, L. H. K., Boonstra, J. & De Laat, S. W. (1991). Antibodyinduced dimerization activates the epidermal growth factor receptor tyrosine kinase. J. biol. Chem. 266, 17331739.CrossRefGoogle ScholarPubMed
Stern, D. F., Kamps, M. P. & Cao, H. (1988). Oncogenic activation of p185neu stimulates tyrosine phosphorylation in vivo. Molec. Cell. Biol. 8, 39693973.Google ScholarPubMed
Sternberg, M. J. E. & Gullick, W. J. (1989). Neu receptor dimerization. Nature, Lond. 339, 587.CrossRefGoogle ScholarPubMed
Sternberg, M. J. E. & Gullick, W. J. (1990). A sequence motif in the transmembrane region of growth factor receptors with tyrosine kinase activity mediates dimerization. Protein Engineering, 3, 245248.CrossRefGoogle ScholarPubMed
Stoddard, B. L., Bui, J. D. & Koshland, D. E. Jr. (1992). Structure and dynamics of transmembrane signaling by the Escherichia coli aspartate receptor. Biochemistry 31, 1197811983.CrossRefGoogle ScholarPubMed
Suryanarayana, S., von Zatrow, M. & Kobilka, B. K. (1992). Identification of intramolecular interactions in adrenergic receptors. J. biol. Chem. 267; 2199121994.CrossRefGoogle ScholarPubMed
Sussman, J. J., Bonifacino, J. S., Lippincott-Schwartz, J., Weissman, A. M., Saito, T., Klausner, R. D. & Ashwell, J. D. (1988). Failure to synthesize the T-cell CD3-ζ chain: Structure and function of a partial T-cell receptor complex. Cell 52, 8595.CrossRefGoogle ScholarPubMed
Swift, A. M. & Machamer, C. E. (1991). A Golgi retention signal in a membrane-spanning domain of coronavirus E1 protein. J. Cell Biol. 115, 1930.CrossRefGoogle Scholar
Tan, L., Turner, J. & Weiss, A. (1991). Regions of the T-cell receptor α and β chains that are responsible for interactions with CD3. J. exp. Med. 173, 12471256.CrossRefGoogle ScholarPubMed
Tang, B. L., Wong, S. H., Low, S. H. & Hon, H. (1992). The transmembrane domain of N-glucosaminyltransferase I contains a Golgi retention signal. J. biol. Chem. 267, 1012210126.CrossRefGoogle ScholarPubMed
Teasdale, R. D., D'Agostaro, G. & Gleeson, P. A. (1992). The signal for Golgi retention of bovine β1, 4-galactosyltransferase is in the transmembrane domain. J. biol. Chem. 267, 40844096.CrossRefGoogle Scholar
Teufel, M., Pompejus, M., Humbel, B., Friedrich, K. & Fritz, H.-J. (1993). Properties of bacteriorhodopsin derivatives constructed by insertion of an exogenous epitope into extra-membrane loops. EMBO J. 12, 33993408.CrossRefGoogle ScholarPubMed
Tomita, M., Furthmayr, H. & Marchesi, V. T. (1978). Primary structure of human erythrocyte glycophorin A. Isolation and characterization of peptides and complete amino acid sequence. Biochemistry 17, 47564770.CrossRefGoogle ScholarPubMed
Traxler, B. & Beckwith, J. (1992). Assembly of a hetero-oligomeric membrane protein complex. Proc. natn. Acad. Sci. U.S.A. 89, 1085210856.CrossRefGoogle ScholarPubMed
Treutlein, H. R., Lemmon, M. A., Engelman, D. M. & Brünger, A. T. (1992). The glycophorin A transmembrane domain dimer: Sequence-specific propensity for a right-handed supercoil of helices. Biochemistry 31, 1272612732.CrossRefGoogle ScholarPubMed
Treutlein, H. R., Lemmon, M. A., Engelman, D. M. & Brünger, A. T. (1993). Simulation of helix association in membranes: Modeling the glycophorin A transmembrane domain. In: Proceedings of the 26th Hawaii International Conference on System Sciences, Los Alamitos, CA. IEEE Computer Society Press. 1, 708714.Google Scholar
Ullrich, A. & Schlessinger, J. (1990). Signal transduction by receptors with tyrosine kinase activity. Cell 61, 203212.CrossRefGoogle ScholarPubMed
Unwin, N. (1993). Nicotinic acetylcholine receptor at 9 Å resolution. J. molec. Biol. 229, 11011124.CrossRefGoogle ScholarPubMed
Wang, J. & Pullman, A. (1991). Do helices in membranes prefer to form bundles or stay dispersed in the lipid phase ? Biochim. biophys. Acta 1070, 493496.CrossRefGoogle ScholarPubMed
Watowich, S. S., Yoshimura, A., Longmore, G. D., Hilton, D. J., Yoshimura, Y. & Lodish, H. F. (1992). Homodimerization and constitutive activation of the erythropoietin receptor. Proc. natn. Acad. Sci. U.S.A. 89, 21402144.CrossRefGoogle ScholarPubMed
Weiner, D. B., Liu, J., Cohen, J. A., Williams, W. V. & Green, M. I. (1989). A point mutation in the neu oncogene mimics ligand induction of receptor aggregation. Nature, Lond. 339, 230231.CrossRefGoogle ScholarPubMed
Weiss, M. S., Wacker, T., Weckesser, J., Welte, W. & Schulz, G. E. (1990). The three dimensional structure of porin from Rhodobacter capsulatus at 3 Å resolution. FEBS Lett. 267, 268272.CrossRefGoogle ScholarPubMed
Weiss, M. S., Abele, U., Weckesser, J., Welte, W., Schiltz, E. & Schultz, G. E. (1991). Molecular architecture and electrostatic properties of a bacterial porin. Science, N.Y. 254, 16271630.CrossRefGoogle ScholarPubMed
Weisz, O. A., Swift, A. M. & Machamer, C. E. (1993). Oligomerization of a membrane protein correlates with its retention in the Golgi complex. J. Cell Biol. 122, 11861196.CrossRefGoogle ScholarPubMed
Welsh, E. J., Thom, D., Morris, E. R. & Rees, D. A. (1985). Molecular organization of glycophorin A: implications for membrane interactions. Biopolymers, 24, 23012332.CrossRefGoogle ScholarPubMed
Whitley, P., Nilsson, L. & von Heijne, G. (1993). Three-dimensional model for the membrane domain of Escherichia coli leader peptidase based on disulfide mapping. Biochemistry 32, 85348539.CrossRefGoogle ScholarPubMed
Wides, R. J., Zak, N. B. & Shilo, B.-Z. (1990). Enhancement of tyrosine kinase activity of the Drosophila epidermal growth factor homolog by alterations of the transmembrane domain. Eur.J. Biochem. 189, 637645.CrossRefGoogle ScholarPubMed
Williams, G. T., Venkitaraman, A. R., Gilmore, D. J. & Neuberger, M. S. (1990). The sequence of the μ transmembrane segment determines the tissue specificity of the transport of immunoglobulin M to the cell surface. J. exp. Med. 171, 947952.CrossRefGoogle ScholarPubMed
Wong, S. H., Low, S. H. & Hong, W. (1992). The 17-residue transmembrane domain of beta-galactoside alpha 2, 6-sialyltransferase is sufficient for Golgi retention. J. Cell Biol. 117, 245258.CrossRefGoogle ScholarPubMed
Wozniak, R. W. & Blobel, G. (1992). The single transmembrane segment of gp210 is sufficient for sorting to the pore membrane domain of the nuclear envelope. J. Cell Biol. 119, 14411449.CrossRefGoogle Scholar
Wrubel, W., Stochaj, U., Sonnewald, U., Theres, C. & Ehring, R. (1990). Reconstitution of an active lactose carrier by simultaneous synthesis of two complementary protein fragments. J. Bacteriol. 172, 53745381.CrossRefGoogle ScholarPubMed
Yamada, K., Goncalves, E., Kahn, C. R. & Shoelson, S. E. (1992). Substitution of the insulin receptor transmembrane domain with the c-neu/erb B2 transmembrane domain constitutively activates the insulin receptor in vitro. J. biol. Chem. 267, 1245212461.CrossRefGoogle Scholar
Yamamoto, T., Ikawa, S., Akiyama, T., Semba, K., Nomura, N., Miyajima, N., Saito, T. & Toyoshima, K. (1986). Similarity of protein encoded by the human c-erb-B-2 gene to epidermal growth factor receptor. Nature, Lond. 319, 230234.CrossRefGoogle ScholarPubMed
Yan, H., Schlessinger, J. & Chao, M. V. (1991). Chimeric NGF-EGF receptors define domains responsible for neuronal differentiation. Science, N.Y. 252, 561563.CrossRefGoogle ScholarPubMed
Yarden, Y. (1990). Agonistic antibodies stimulate the kinase encoded by the neu protooncogene in living cells but the oncogenic mutant is constitutively active. Proc. natn. Acad. Sci. U.S.A. 87, 25692573.CrossRefGoogle ScholarPubMed
Yeates, T. O., Komiya, H., Rees, D. C., Allen, J. P. & Feher, G. (1987). Structure of The reaction center from Rhodobacter sphaeroides R-26: Membrane-protein interactions. Proc. natn. Acad. Sci. U.S.A. 84, 64386442.CrossRefGoogle ScholarPubMed
Yeh, J. I. T., Biemann, H.-P., Pandit, J., Koshland, D. E. Jr., & Kim, S.-H. (1993). Three-dimensional structure of the ligand binding domains of wild-type bacterial chemotaxis receptor: Structural comparison to the cross-linked mutant forms and conformational changes upon binding. J. biol. Chem. 268, 97879792.CrossRefGoogle Scholar
Zhang, Y.-P., Lewis, R. N. A. H., Hodges, R. S. & McElhaney, R. N. (1992). Interaction of a peptide model of a hydrophobic transmembrane α-helical segment of a membrane protein with phosphatidylcholine bilayers: Differential scanning calorimetric and FTIR spectroscopic studies. Biochemistry 31, 1157911588.CrossRefGoogle ScholarPubMed
Zhou, M., Felder, S., Rubinstein, M., Hurwitz, D. R., Ullrich, A., Lax, I. & Schlessinger, J. (1993). Real time measurements of kinetics of EGF binding to soluble EGF-receptor monomers and dimers support the dimerization model for receptor activation. Biochemistry 32, 81938198.CrossRefGoogle ScholarPubMed
Zon, L. I., Moreau, J.-F., Koo, J.-W., Mathey-Prevot, B. & D'Andrea, A. D. (1992). The erythropoietin receptor transmembrane region is necessary for activation by the friend spleen focus-forming virus gp55 glycoprotein. Molec. Cell. Biol. 12, 29492957.Google ScholarPubMed