Hostname: page-component-cd9895bd7-7cvxr Total loading time: 0 Render date: 2024-12-29T01:07:12.047Z Has data issue: false hasContentIssue false

The safety dance: biophysics of membrane protein folding and misfolding in a cellular context

Published online by Cambridge University Press:  25 November 2014

Jonathan P. Schlebach
Affiliation:
Department of Biochemistry and Center for Structural Biology, Vanderbilt University School of Medicine, Nashville, Tennessee 37232, USA
Charles R. Sanders*
Affiliation:
Department of Biochemistry and Center for Structural Biology, Vanderbilt University School of Medicine, Nashville, Tennessee 37232, USA
*
*Author for correspondence: C. R. Sanders, Department of Biochemistry and Center for Structural Biology, Vanderbilt University School of Medicine, Nashville, Tennessee 37232, USA. Phone: 615-936-3756; Fax: 615-936-2211; Email: [email protected]

Abstract

Most biological processes require the production and degradation of proteins, a task that weighs heavily on the cell. Mutations that compromise the conformational stability of proteins place both specific and general burdens on cellular protein homeostasis (proteostasis) in ways that contribute to numerous diseases. Efforts to elucidate the chain of molecular events responsible for diseases of protein folding address one of the foremost challenges in biomedical science. However, relatively little is known about the processes by which mutations prompt the misfolding of α-helical membrane proteins, which rely on an intricate network of cellular machinery to acquire and maintain their functional structures within cellular membranes. In this review, we summarize the current understanding of the physical principles that guide membrane protein biogenesis and folding in the context of mammalian cells. Additionally, we explore how pathogenic mutations that influence biogenesis may differ from those that disrupt folding and assembly, as well as how this may relate to disease mechanisms and therapeutic intervention. These perspectives indicate an imperative for the use of information from structural, cellular, and biochemical studies of membrane proteins in the design of novel therapeutics and in personalized medicine.

Type
Review Article
Copyright
Copyright © Cambridge University Press 2014 

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

7. References

Adamian, L. & Liang, J. (2002). Interhelical hydrogen bonds and spatial motifs in membrane proteins: polar clamps and serine zippers. Proteins 47(2), 209218.Google Scholar
Allen, S., Curran, A., Templer, R., Meijberg, W. & Booth, P. (2004a). Controlling the folding efficiency of an integral membrane protein. Journal of Molecular Biology 342(4), 12931304.Google Scholar
Allen, S., Curran, A., Templer, R., Meijberg, W. & Booth, P. (2004b). Folding kinetics of an alpha helical membrane protein in phospholipid bilayer vesicles. Journal of Molecular Biology 342(4), 12791291.Google Scholar
Anfinsen, C. B. (1973). Principles that govern the folding of protein chains. Science 181(4096), 223230.Google Scholar
Baker, R. P. & Urban, S. (2012). Architectural and thermodynamic principles underlying intramembrane protease function. Nature Chemical Biology 8(9), 759768.Google Scholar
Balch, W. E., Morimoto, R. I., Dillin, A. & Kelly, J. W. (2008). Adapting proteostasis for disease intervention. Science 319(5865), 916919.Google Scholar
Barrera, F. N., Renart, M. L., Poveda, J. A., De Kruijff, B., Killian, J. A. & González-Ros, J. M. (2008). Protein self-assembly and lipid binding in the folding of the potassium channel KcsA. Biochemistry 47(7), 21232133.Google Scholar
Bayle, D., Weeks, D. & Sachs, G. (1995). The membrane topology of the rat sarcoplasmic and endoplasmic reticulum calcium ATPases by in vitro translation scanning. Journal of Biological Chemistry 270(43), 2567825684.Google Scholar
Bence, N. F., Sampat, R. M. & Kopito, R. R. (2001). Impairment of the ubiquitin-proteasome system by protein aggregation. Science 292(5521), 15521555.Google Scholar
Berk, J. L., Suhr, O. B., Obici, L., Sekijima, Y., Zeldenrust, S. R., Yamashita, T., Heneghan, M. A., Gorevic, P. D., Litchy, W. J., Wiesman, J. F., Nordh, E., Corato, M., Lozza, A., Cortese, A., Robinson-Papp, J., Colton, T., Rybin, D. V., Bisbee, A. B., Ando, Y., Ikeda, S., Seldin, D. C., Merlini, G., Skinner, M., Kelly, J. W., Dyck, P. J. & Diflunisal Trial, C. (2013). Repurposing diflunisal for familial amyloid polyneuropathy: a randomized clinical trial. Journal of the American Medical Association 310(24), 26582667.Google Scholar
Bolen, D. W. & Rose, G. D. (2008). Structure and energetics of the hydrogen-bonded backbone in protein folding. Annual Review of Biochemistry 77, 339362.CrossRefGoogle ScholarPubMed
Booth, P. & Curnow, P. (2009). Folding scene investigation: membrane proteins. Current Opinion in Structural Biology 19(1), 813.Google Scholar
Booth, P. J., Flitsch, S. L., Stern, L. J., Greenhalgh, D. A., Kim, P. S. & Khorana, H. G. (1995). Intermediates in the folding of the membrane protein bacteriorhodopsin. Nature Structural Biology 2(2), 139143.Google Scholar
Bowie, J. (2005). Solving the membrane protein folding problem. Nature 438(7068), 581589.Google Scholar
Bowie, J. U. (2011). Membrane protein folding: how important are hydrogen bonds? Current Opinion in Structural Biology 21(1), 4249.Google Scholar
Brockwell, D. J. & Radford, S. E. (2007). Intermediates: ubiquitous species on folding energy landscapes?. Current Opinion in Structural Biology 17(1), 3037.Google Scholar
Brodsky, J. L. (2012). Cleaning up: ER-associated degradation to the rescue. Cell 151(6), 11631167.Google Scholar
Brodsky, J. L. & Skach, W. R. (2011). Protein folding and quality control in the endoplasmic reticulum: recent lessons from yeast and mammalian cell systems. Current Opinion in Cell Biology 23(4), 464475.Google Scholar
Brown, C. R., Hong-Brown, L. Q. & Welch, W. J. (1997). Correcting temperature-sensitive protein folding defects. Journal of Clinical Investigation 99(6), 14321444.Google Scholar
Brown, M. F. (2012). Curvature forces in membrane lipid–protein interactions. Biochemistry 51(49), 97829795.Google Scholar
Buchberger, A., Bukau, B. & Sommer, T. (2010). Protein quality control in the cytosol and the endoplasmic reticulum: brothers in arms. Molecular Cell 40(2), 238252.Google Scholar
Buck, T. M. & Skach, W. R. (2005). Differential stability of biogenesis intermediates reveals a common pathway for aquaporin-1 topological maturation. Journal of Biological Chemistry 280(1), 261269.Google Scholar
Burgess, N., Dao, T., Stanley, A. & Fleming, K. (2008). Beta-barrel proteins that reside in the Escherichia coli outer membrane in vivo demonstrate varied folding behavior in vitro. Journal of Biological Chemistry 283(39), 2674826758.Google Scholar
Calamini, B., Silva, M. C., Madoux, F., Hutt, D. M., Khanna, S., Chalfant, M. A., Saldanha, S. A., Hodder, P., Tait, B. D., Garza, D., Balch, W. E. & Morimoto, R. I. (2012). Small-molecule proteostasis regulators for protein conformational diseases. Nature Chemical Biology 8(2), 185196.Google Scholar
Cannon, K. S. & Cresswell, P. (2001). Quality control of transmembrane domain assembly in the tetraspanin CD82. EMBO Journal 20(10), 24432453.Google Scholar
Cao, Z. & Bowie, J. U. (2012). Shifting hydrogen bonds may produce flexible transmembrane helices. Proceedings of the National Academy of Sciences of the United States of America 109(21), 81218126.Google Scholar
Cao, Z., Schlebach, J., Park, C. & Bowie, J. U. (2011). Thermodynamic stability of bacteriorhodopsin mutants measured relative to the bacterioopsin unfolded state. Biochimica et Biophysica Acta 1818(4), 10491054.Google Scholar
Carlile, G. W., Keyzers, R. A., Teske, K. A., Robert, R., Williams, D. E., Linington, R. G., Gray, C. A., Centko, R. M., Yan, L., Anjos, S. M., Sampson, H. M., Zhang, D., Liao, J., Hanrahan, J. W., Andersen, R. J. & Thomas, D. Y. (2012). Correction of F508del-CFTR trafficking by the sponge alkaloid latonduine is modulated by interaction with PARP. Chemical Biology 19(10), 12881299.Google Scholar
Cestèle, S., Schiavon, E., Rusconi, R., Franceschetti, S. & Mantegazza, M. (2013). Nonfunctional NaV1·1 familial hemiplegic migraine mutant transformed into gain of function by partial rescue of folding defects. Proceedings of the National Academy of Sciences of the United States of America 110(43), 1754617551.Google Scholar
Chang, Y. C. & Bowie, J. U. (2014). Measuring membrane protein stability under native conditions. Proceedings of the National Academy of Sciences of the United States of America 111(1), 219224.Google Scholar
Chen, F., Degnin, C., Laederich, M., Horton, W. A. & Hristova, K. (2011). The A391E mutation enhances FGFR3 activation in the absence of ligand. Biochimica et Biophysica Acta 1808(8), 20452050.Google Scholar
Chen, M. & Zhang, J. T. (1999). Topogenesis of cystic fibrosis transmembrane conductance regulator (CFTR): regulation by the amino terminal transmembrane sequences. Biochemistry 38(17), 54715477.Google Scholar
Cheng, S. H., Gregory, R. J., Marshall, J., Paul, S., Souza, D. W., White, G. A., O'Riordan, C. R. & Smith, A. E. (1990). Defective intracellular transport and processing of CFTR is the molecular basis of most cystic fibrosis. Cell 63(4), 827834.Google Scholar
Cohen, F. E. & Kelly, J. W. (2003). Therapeutic approaches to protein-misfolding diseases. Nature 426(6968), 905909.CrossRefGoogle ScholarPubMed
Coppinger, J. A., Hutt, D. M., Razvi, A., Koulov, A. V., Pankow, S., Yates, J. R. III & Balch, W. E. (2012). A chaperone trap contributes to the onset of cystic fibrosis. PLoS ONE 7(5), e37682.Google Scholar
Curnow, P. & Booth, P. (2007). Combined kinetic and thermodynamic analysis of alpha-helical membrane protein unfolding. Proceedings of the National Academy of Sciences of the United States of America 104(48), 1897018975.Google Scholar
Curnow, P. & Booth, P. (2009). The transition state for integral membrane protein folding. Proceedings of the National Academy of Sciences of the United States of America 106(3), 773778.Google Scholar
Curnow, P. & Booth, P. J. (2010). The contribution of a covalently bound cofactor to the folding and thermodynamic stability of an integral membrane protein. Journal of Molecular Biology 403(4), 630642.Google Scholar
Curnow, P., Di Bartolo, N. D., Moreton, K. M., Ajoje, O. O., Saggese, N. P. & Booth, P. J. (2011). Stable folding core in the folding transition state of an alpha-helical integral membrane protein. Proceedings of the National Academy of Sciences of the United States of America 108(34), 1413314138.Google Scholar
Cymer, F. & Von Heijne, G. (2013). Cotranslational folding of membrane proteins probed by arrest-peptide-mediated force measurements. Proceedings of the National Academy of Sciences of the United States of America 110(36), 1464014645.Google Scholar
Dedola, S., Izumi, M., Makimura, Y., Seko, A., Kanamori, A., Sakono, M., Ito, Y. & Kajihara, Y. (2014). Folding of synthetic homogeneous glycoproteins in the presence of a glycoprotein folding sensor enzyme. Angewandte Chemie International Edition English 53(11), 28832887.Google Scholar
Denks, K., Vogt, A., Sachelaru, I., Petriman, N. A., Kudva, R. & Koch, H. G. (2014). The Sec translocon mediated protein transport in prokaryotes and eukaryotes. Molecular Membrane Biology 31(2–3), 5884.Google Scholar
Denning, G. M., Anderson, M. P., Amara, J. F., Marshall, J., Smith, A. E. & Welsh, M. J. (1992). Processing of mutant cystic fibrosis transmembrane conductance regulator is temperature-sensitive. Nature 358(6389), 761764.Google Scholar
Devaraneni, P. K., Conti, B., Matsumura, Y., Yang, Z., Johnson, A. E. & Skach, W. R. (2011). Stepwise insertion and inversion of a type II signal anchor sequence in the ribosome-Sec61 translocon complex. Cell 146(1), 134147.Google Scholar
Dhaunchak, A. S. & Nave, K. A. (2007). A common mechanism of PLP/DM20 misfolding causes cysteine-mediated endoplasmic reticulum retention in oligodendrocytes and Pelizaeus-Merzbacher disease. Proceedings of the National Academy of Sciences of the United States of America 104(45), 1781317818.Google Scholar
Dill, K. A. & Maccallum, J. L. (2012). The protein-folding problem, 50 years on. Science 338(6110), 10421046.Google Scholar
Dill, K. A., Ghosh, K. & Schmit, J. D. (2011). Physical limits of cells and proteomes. Proceedings of the National Academy of Sciences of the United States of America 108(44), 1787617882.Google Scholar
Doura, A. K., Kobus, F. J., Dubrovsky, L., Hibbard, E. & Fleming, K. G. (2004). Sequence context modulates the stability of a GxxxG-mediated transmembrane helix–helix dimer. Journal of Molecular Biology 341(4), 991998.Google Scholar
Dutta, A., Kim, T. Y., Moeller, M., Wu, J., Alexiev, U. & Klein-Seetharaman, J. (2010). Characterization of membrane protein non-native states. 2. The SDS-unfolded states of rhodopsin. Biochemistry 49(30), 63296340.Google Scholar
Eckford, P. D., Ramjeesingh, M., Molinski, S., Pasyk, S., Dekkers, J. F., Li, C., Ahmadi, S., Ip, W., Chung, T. E., Du, K., Yeger, H., Beekman, J., Gonska, T. & Bear, C. E. (2014). VX-809 and related corrector compounds exhibit secondary activity stabilizing active F508del-CFTR after its partial rescue to the cell surface. Chemical Biology, 21(5), 666678.Google Scholar
Egea, P. F. & Stroud, R. M. (2010). Lateral opening of a translocon upon entry of protein suggests the mechanism of insertion into membranes. Proceedings of the National Academy of Sciences of the United States of America 107(40), 1718217187.Google Scholar
Eletto, D., Maganty, A., Eletto, D., Dersh, D., Makarewich, C., Biswas, C., Paton, J. C., Paton, A. W., Doroudgar, S., Glembotski, C. C. & Argon, Y. (2012). Limitation of individual folding resources in the ER leads to outcomes distinct from the unfolded protein response. Journal of Cell Science 125(Pt 20), 48654875.Google Scholar
Ellgaard, L. & Helenius, A. (2003). Quality control in the endoplasmic reticulum. Nature Reviews Molecular Cell Biology 4(3), 181191.Google Scholar
Engelman, D. M. & Steitz, T. A. (1981). The spontaneous insertion of proteins into and across membranes: the helical hairpin hypothesis. Cell 23(2), 411422.Google Scholar
Engelman, D. M., Chen, Y., Chin, C. N., Curran, A. R., Dixon, A. M., Dupuy, A. D., Lee, A. S., Lehnert, U., Matthews, E. E., Reshetnyak, Y. K., Senes, A. & Popot, J. L. (2003). Membrane protein folding: beyond the two stage model. FEBS Letters 555(1), 122125.Google Scholar
Escusa-Toret, S., Vonk, W. I. & Frydman, J. (2013). Spatial sequestration of misfolded proteins by a dynamic chaperone pathway enhances cellular fitness during stress. Nature Cellular Biology 15(10), 12311243.Google Scholar
Faham, S., Yang, D., Bare, E., Yohannan, S., Whitelegge, J. & Bowie, J. (2004). Side-chain contributions to membrane protein structure and stability. Journal of Molecular Biology 335(1), 297305.Google Scholar
Fan, J. Q., Ishii, S., Asano, N. & Suzuki, Y. (1999). Accelerated transport and maturation of lysosomal alpha-galactosidase A in Fabry lymphoblasts by an enzyme inhibitor. Nature Medicine 5(1), 112115.Google Scholar
Farinha, C. M., Matos, P. & Amaral, M. D. (2013). Control of cystic fibrosis transmembrane conductance regulator membrane trafficking: not just from the endoplasmic reticulum to the Golgi. FEBS Journal 280(18), 43964406.Google Scholar
Feige, M. J. & Hendershot, L. M. (2013). Quality control of integral membrane proteins by assembly-dependent membrane integration. Molecular Cell 51(3), 297309.Google Scholar
Ficker, E., Obejero-Paz, C. A., Zhao, S. & Brown, A. M. (2002). The binding site for channel blockers that rescue misprocessed human long QT syndrome type 2 ether-a-gogo-related gene (HERG) mutations. Journal of Biological Chemistry 277(7), 49894998.Google Scholar
Findlay, H. E., Rutherford, N. G., Henderson, P. J. & Booth, P. J. (2010). Unfolding free energy of a two-domain transmembrane sugar transport protein. Proceedings of the National Academy of Sciences of the United States of America 107(43), 1845118456.Google Scholar
Fleming, K. G. (2002). Standardizing the free energy change of transmembrane helix–helix interactions. Journal of Molecular Biology 323(3), 563571.Google Scholar
Fleming, K. G. (2014). Energetics of membrane protein folding. Annual Review of Biophysics 43, 233255.Google Scholar
Fleming, K. G. & Engelman, D. M. (2001). Specificity in transmembrane helix–helix interactions can define a hierarchy of stability for sequence variants. Proceedings of the National Academy of Sciences of the United States of America 98(25), 1434014344.Google Scholar
Fleming, K. G., Ackerman, A. L. & Engelman, D. M. (1997). The effect of point mutations on the free energy of transmembrane alpha-helix dimerization. Journal of Molecular Biolology 272(2), 266275.Google Scholar
Fontanini, A., Chies, R., Snapp, E. L., Ferrarini, M., Fabrizi, G. M. & Brancolini, C. (2005). Glycan-independent role of calnexin in the intracellular retention of Charcot–Marie-tooth 1A Gas3/PMP22 mutants. Journal of Biological Chemistry 280(3), 23782387.Google Scholar
Fortun, J., Dunn, W. A., Joy, S., Li, J. & Notterpek, L. (2003). Emerging role for autophagy in the removal of aggresomes in Schwann cells. Journal of Neuroscience 23(33), 1067210680.Google Scholar
Fortun, J., Li, J., Go, J., Fenstermaker, A., Fletcher, B. S. & Notterpek, L. (2005). Impaired proteasome activity and accumulation of ubiquitinated substrates in a hereditary neuropathy model. Journal of Neurochemistry 92(6), 15311541.Google Scholar
Gafvelin, G. & Von Heijne, G. (1994). Topological “frustration” in multispanning E. coli inner membrane proteins. Cell 77(3), 401412.CrossRefGoogle ScholarPubMed
Gamerdinger, M., Kaya, A. M., Wolfrum, U., Clement, A. M. & Behl, C. (2011). BAG3 mediates chaperone-based aggresome-targeting and selective autophagy of misfolded proteins. EMBO Reports 12(2), 149156.Google Scholar
Garcia-Mata, R., Bebok, Z., Sorscher, E. J. & Sztul, E. S. (1999). Characterization and dynamics of aggresome formation by a cytosolic GFP-chimera. Journal of Cell Biology 146(6), 12391254.Google Scholar
Garcia-Mata, R., Gao, Y. S. & Sztul, E. (2002). Hassles with taking out the garbage: aggravating aggresomes. Traffic 3(6), 388396.CrossRefGoogle ScholarPubMed
Gessmann, D., Chung, Y. H., Danoff, E. J., Plummer, A. M., Sandlin, C. W., Zaccai, N. R. & Fleming, K. G. (2014). Outer membrane beta-barrel protein folding is physically controlled by periplasmic lipid head groups and BamA. Proceedings of the National Academy of Sciences of the United States of America 111(16), 58785883.Google Scholar
Geva, Y. & Schuldiner, M. (2014). The back and forth of cargo exit from the endoplasmic reticulum. Current Biology 24(3), R130R136.Google Scholar
Ghosh, K. & Dill, K. A. (2009). Computing protein stabilities from their chain lengths. Proceedings of the National Academy of Sciences of the United States of America 106(26), 1064910654.Google Scholar
Ghosh, K. & Dill, K. (2010). Cellular proteomes have broad distributions of protein stability. Biophysical Journal 99(12), 39964002.Google Scholar
Gidalevitz, T., Ben-Zvi, A., Ho, K. H., Brignull, H. R. & Morimoto, R. I. (2006). Progressive disruption of cellular protein folding in models of polyglutamine diseases. Science 311(5766), 14711474.Google Scholar
Gidalevitz, T., Kikis, E. A. & Morimoto, R. I. (2010). A cellular perspective on conformational disease: the role of genetic background and proteostasis networks. Current Opinion in Structural Biology 20(1), 2332.Google Scholar
Goldberg, A. F., Loewen, C. J. & Molday, R. S. (1998). Cysteine residues of photoreceptor peripherin/rds: role in subunit assembly and autosomal dominant retinitis pigmentosa. Biochemistry 37(2), 680685.Google Scholar
Gong, Q., Jones, M. A. & Zhou, Z. (2006). Mechanisms of pharmacological rescue of trafficking-defective hERG mutant channels in human long QT syndrome. Journal of Biological Chemistry 281(7), 40694074.Google Scholar
Gratkowski, H., Lear, J. D. & Degrado, W. F. (2001). Polar side chains drive the association of model transmembrane peptides. Proceedings of the National Academy of Sciences of the United States of America 98(3), 880885.Google Scholar
Harris, N. J., Findlay, H. E., Simms, J., Liu, X. & Booth, P. J. (2014). Relative domain folding and stability of a membrane transport protein. Journal of Molecular Biology 426(8), 18121825.Google Scholar
He, L. & Hristova, K. (2008). Pathogenic activation of receptor tyrosine kinases in mammalian membranes. Journal of Molecular Biology 384(5), 11301142.Google Scholar
Hebert, D. N. & Molinari, M. (2012). Flagging and docking: dual roles for N-glycans in protein quality control and cellular proteostasis. Trends in Biochemical Sciences 37(10), 404410.Google Scholar
Heinrich, S. U. & Rapoport, T. A. (2003). Cooperation of transmembrane segments during the integration of a double-spanning protein into the ER membrane. EMBO Journal 22(14), 36543663.Google Scholar
Heinrich, S. U., Mothes, W., Brunner, J. & Rapoport, T. A. (2000). The Sec61p complex mediates the integration of a membrane protein by allowing lipid partitioning of the transmembrane domain. Cell 102(2), 233244.Google Scholar
Hermansson, M. & Von Heijne, G. (2003). Inter-helical hydrogen bond formation during membrane protein integration into the ER membrane. Journal of Molecular Biology 334(4), 803809.Google Scholar
Hessa, T., Kim, H., Bihlmaier, K., Lundin, C., Boekel, J., Andersson, H., Nilsson, I., White, S. H. & Von Heijne, G. (2005). Recognition of transmembrane helices by the endoplasmic reticulum translocon. Nature 433(7024), 377381.Google Scholar
Hessa, T., Meindl-Beinker, N. M., Bernsel, A., Kim, H., Sato, Y., Lerch-Bader, M., Nilsson, I., White, S. H. & Von Heijne, G. (2007). Molecular code for transmembrane-helix recognition by the Sec61 translocon. Nature 450(7172), 10261030.Google Scholar
Hong, H. & Bowie, J. U. (2011). Dramatic destabilization of transmembrane helix interactions by features of natural membrane environments. Journal of the American Chemical Society 133(29), 1138911398.Google Scholar
Hong, H. & Tamm, L. (2004). Elastic coupling of integral membrane protein stability to lipid bilayer forces. Proceedings of the National Academy of Sciences of the United States of America 101(12), 40654070.Google Scholar
Hong, H., Joh, N., Bowie, J. & Tamm, L. (2009). Methods for measuring the thermodynamic stability of membrane proteins. Methods in Enzymology 455, 213236.Google Scholar
Hong, H., Blois, T. M., Cao, Z. & Bowie, J. U. (2010). Method to measure strong protein-protein interactions in lipid bilayers using a steric trap. Proceedings of the National Academy of Sciences of the United States of America 107(46), 1980219807.Google Scholar
Houck, S. A. & Cyr, D. M. (2012). Mechanisms for quality control of misfolded transmembrane proteins. Biochimica et Biophysica Acta 1818(4), 11081114.Google Scholar
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. Journal of Biological Chemistry 256(8), 38023809.Google Scholar
Huysmans, G. H., Baldwin, S. A., Brockwell, D. J. & Radford, S. E. (2010). The transition state for folding of an outer membrane protein. Proceedings of the National Academy of Sciences of the United States of America 107(9), 40994104.Google Scholar
Huysmans, G. H., Radford, S. E., Baldwin, S. A. & Brockwell, D. J. (2012). Malleability of the folding mechanism of the outer membrane protein PagP: parallel pathways and the effect of membrane elasticity. Journal of Molecular Biology 416(3), 453464.Google Scholar
Hwa, J., Reeves, P. J., Klein-Seetharaman, J., Davidson, F. & Khorana, H. G. (1999). Structure and function in rhodopsin: further elucidation of the role of the intradiscal cysteines, Cys-110, -185, and -187, in rhodopsin folding and function. Proceedings of the National Academy of Sciences of the United States of America 96(5), 19321935.Google Scholar
Illergard, K., Kauko, A. & Elofsson, A. (2011). Why are polar residues within the membrane core evolutionary conserved? Proteins 79(1), 7991.Google Scholar
Janovick, J. A., Stewart, M. D., Jacob, D., Martin, L. D., Deng, J. M., Stewart, C. A., Wang, Y., Cornea, A., Chavali, L., Lopez, S., Mitalipov, S., Kang, E., Lee, H. S., Manna, P. R., Stocco, D. M., Behringer, R. R. & Conn, P. M. (2013). Restoration of testis function in hypogonadotropic hypogonadal mice harboring a misfolded GnRHR mutant by pharmacoperone drug therapy. Proceedings of the National Academy of Sciences of the United States of America 110(52), 2103021035.Google Scholar
Jaswal, S. S., Sohl, J. L., Davis, J. H. & Agard, D. A. (2002). Energetic landscape of alpha-lytic protease optimizes longevity through kinetic stability. Nature 415(6869), 343346.Google Scholar
Jefferson, R. E., Blois, T. M. & Bowie, J. U. (2013). Membrane proteins can have high kinetic stability. Journal of the American Chemical Society 135(40), 1518315190.Google Scholar
Jetten, A. M. & Suter, U. (2000). The peripheral myelin protein 22 and epithelial membrane protein family. Progress in Nucleic Acid Research and Molecular Biology 64, 97129.Google Scholar
Joh, N., Min, A., Faham, S., Whitelegge, J., Yang, D., Woods, V. & Bowie, J. (2008). Modest stabilization by most hydrogen-bonded side-chain interactions in membrane proteins. Nature 453(7199), 12661270.Google Scholar
Joh, N. H., Oberai, A., Yang, D., Whitelegge, J. P. & Bowie, J. U. (2009). Similar energetic contributions of packing in the core of membrane and water-soluble proteins. Journal of the American Chemical Society 131(31), 1084610847.Google Scholar
Johnson, A. E. & Van Waes, M. A. (1999). The translocon: a dynamic gateway at the ER membrane. Annual Review of Cell and Developmental Biology 15, 799842.Google Scholar
Johnston, J. A., Ward, C. L. & Kopito, R. R. (1998). Aggresomes: a cellular response to misfolded proteins. Journal of Cell Biology 143(7), 18831898.Google Scholar
Johnston, J. A., Illing, M. E. & Kopito, R. R. (2002). Cytoplasmic dynein/dynactin mediates the assembly of aggresomes. Cell Motility and the Cytoskeleton 53(1), 2638.Google Scholar
Jung, J., Coe, H. & Michalak, M. (2011). Specialization of endoplasmic reticulum chaperones for the folding and function of myelin glycoproteins P0 and PMP22. FASEB Journal 25(11), 39293937.Google Scholar
Kaganovich, D., Kopito, R. & Frydman, J. (2008). Misfolded proteins partition between two distinct quality control compartments. Nature 454(7208), 10881095.Google Scholar
Kanehara, K., Xie, W. & Ng, D. T. (2010). Modularity of the Hrd1 ERAD complex underlies its diverse client range. Journal of Cell Biology 188(5), 707716.Google Scholar
Kanki, T., Sakaguchi, M., Kitamura, A., Sato, T., Mihara, K. & Hamasaki, N. (2002). The tenth membrane region of band 3 is initially exposed to the luminal side of the endoplasmic reticulum and then integrated into a partially folded band 3 intermediate. Biochemistry 41(47), 1397313981.Google Scholar
Kanner, E. M., Klein, I. K., Friedlander, M. & Simon, S. M. (2002). The amino terminus of opsin translocates “posttranslationally” as efficiently as cotranslationally. Biochemistry 41(24), 77077715.Google Scholar
Kauko, A., Hedin, L. E., Thebaud, E., Cristobal, S., Elofsson, A. & Von Heijne, G. (2010). Repositioning of transmembrane alpha-helices during membrane protein folding. Journal of Molecular Biology 397(1), 190201.Google Scholar
Kaushal, S. & Khorana, H. G. (1994). Structure and function in rhodopsin. 7. Point mutations associated with autosomal dominant retinitis pigmentosa. Biochemistry 33(20), 61216128.Google Scholar
Kawaguchi, Y., Kovacs, J. J., Mclaurin, A., Vance, J. M., Ito, A. & Yao, T. P. (2003). The deacetylase HDAC6 regulates aggresome formation and cell viability in response to misfolded protein stress. Cell 115(6), 727738.Google Scholar
Kelly, J. W. & Balch, W. E. (2006). The integration of cell and chemical biology in protein folding. Nature Chemical Biology 2(5), 224227.Google Scholar
Khushoo, A., Yang, Z., Johnson, A. E. & Skach, W. R. (2011). Ligand-driven vectorial folding of ribosome-bound human CFTR NBD1. Molecular Cell 41(6), 682692.Google Scholar
Kim, B. L., Schafer, N. P. & Wolynes, P. G. (2014). Predictive energy landscapes for folding alpha-helical transmembrane proteins. Proceedings of the National Academy of Sciences of the United States of America 111(30), 1103111036.Google Scholar
Kim, S. J. & Skach, W. R. (2012). Mechanisms of CFTR folding at the endoplasmic reticulum. Frontiers in Pharmacology 3, 201.Google Scholar
Kopito, R. R. (2000). Aggresomes, inclusion bodies and protein aggregation. Trends in Cell Biology 10(12), 524530.Google Scholar
Kowalski, J. M., Parekh, R. N., Mao, J. & Wittrup, K. D. (1998a). Protein folding stability can determine the efficiency of escape from endoplasmic reticulum quality control. Journal of Biological Chemistry 273(31), 1945319458.Google Scholar
Kowalski, J. M., Parekh, R. N. & Wittrup, K. D. (1998b). Secretion efficiency in Saccharomyces cerevisiae of bovine pancreatic trypsin inhibitor mutants lacking disulfide bonds is correlated with thermodynamic stability. Biochemistry 37(5), 12641273.Google Scholar
Krishnamani, V. & Lanyi, J. K. (2011). Structural changes in bacteriorhodopsin during in vitro refolding from a partially denatured state. Biophysical Journal 100(6), 15591567.Google Scholar
Krishnamani, V., Hegde, B. G., Langen, R. & Lanyi, J. K. (2012). Secondary and tertiary structure of bacteriorhodopsin in the SDS denatured state. Biochemistry 51(6), 10511060.Google Scholar
Lau, F. & Bowie, J. (1997). A method for assessing the stability of a membrane protein. Biochemistry 36(19), 58845892.Google Scholar
Lee, M. C., Miller, E. A., Goldberg, J., Orci, L. & Schekman, R. (2004). Bi-directional protein transport between the ER and Golgi. Annual Review of Cell and Developmental Biology 20, 87123.Google Scholar
Li, E., You, M. & Hristova, K. (2006). FGFR3 dimer stabilization due to a single amino acid pathogenic mutation. Journal of Molecular Biology 356(3), 600612.Google Scholar
Li, E., Wimley, W. C. & Hristova, K. (2012a). Transmembrane helix dimerization: beyond the search for sequence motifs. Biochimica et Biophysica Acta 1818(2), 183193.Google Scholar
Li, J., Parker, B., Martyn, C., Natarajan, C. & Guo, J. (2012b). The PMP22 gene and its related diseases. Molecular Neurobiology 47(2), 673698.Google Scholar
Liu, X., Garriga, P. & Khorana, H. G. (1996). Structure and function in rhodopsin: correct folding and misfolding in two point mutants in the intradiscal domain of rhodopsin identified in retinitis pigmentosa. Proceedings of the National Academy of Sciences of the United States of America 93(10), 45544559.Google Scholar
London, E. & Khorana, H. (1982). Denaturation and renaturation of bacteriorhodopsin in detergents and lipid–detergent mixtures. Journal of Biological Chemistry 257(12), 70037011.Google Scholar
Lu, H. & Booth, P. J. (2000). The final stages of folding of the membrane protein bacteriorhodopsin occur by kinetically indistinguishable parallel folding paths that are mediated by pH. Journal of Molecular Biology 299(1), 233243.Google Scholar
Lu, W., Chai, Q., Zhong, M., Yu, L., Fang, J., Wang, T., Li, H., Zhu, H. & Wei, Y. (2012). Assembling of AcrB trimer in cell membrane. Journal of Molecular Biology 423(1), 123134.Google Scholar
Lu, Y., Xiong, X., Helm, A., Kimani, K., Bragin, A. & Skach, W. R. (1998). Co- and posttranslational translocation mechanisms direct cystic fibrosis transmembrane conductance regulator N terminus transmembrane assembly. Journal of Biological Chemistry 273(1), 568576.Google Scholar
Lu, Y., Turnbull, I. R., Bragin, A., Carveth, K., Verkman, A. S. & Skach, W. R. (2000). Reorientation of aquaporin-1 topology during maturation in the endoplasmic reticulum. Molecular Biology of the Cell 11(9), 29732985.CrossRefGoogle ScholarPubMed
Lundbaek, J. A., Andersen, O. S., Werge, T. & Nielsen, C. (2003). Cholesterol-induced protein sorting: an analysis of energetic feasibility. Biophysical Journal 84(3), 20802089.Google Scholar
Maattanen, P., Kozlov, G., Gehring, K. & Thomas, D. Y. (2006). ERp57 and PDI: multifunctional protein disulfide isomerases with similar domain architectures but differing substrate–partner associations. Biochemistry and Cell Biology 84(6), 881889.Google Scholar
Mackenzie, K. R. & Fleming, K. G. (2008). Association energetics of membrane spanning alpha-helices. Current Opinion in Structural Biology 18(4), 412419.Google Scholar
Martin, G. M., Chen, P. C., Devaraneni, P. & Shyng, S. L. (2013). Pharmacological rescue of trafficking-impaired ATP-sensitive potassium channels. Frontiers in Physiology 4, 386.Google Scholar
Matthews, E. E., Zoonens, M. & Engelman, D. M. (2006). Dynamic helix interactions in transmembrane signaling. Cell 127(3), 447450.Google Scholar
Meacham, G. C., Lu, Z., King, S., Sorscher, E., Tousson, A. & Cyr, D. M. (1999). The Hdj-2/Hsc70 chaperone pair facilitates early steps in CFTR biogenesis. EMBO Journal 18(6), 14921505.Google Scholar
Meindl-Beinker, N. M., Lundin, C., Nilsson, I., White, S. H. & Von Heijne, G. (2006). Asn- and Asp-mediated interactions between transmembrane helices during translocon-mediated membrane protein assembly. EMBO Reports 7(11), 11111116.Google Scholar
Merulla, J., Fasana, E., Solda, T. & Molinari, M. (2013). Specificity and regulation of the endoplasmic reticulum-associated degradation machinery. Traffic 14(7), 767777.Google Scholar
Michelsen, K., Yuan, H. & Schwappach, B. (2005). Hide and run. Arginine-based endoplasmic-reticulum-sorting motifs in the assembly of heteromultimeric membrane proteins. EMBO Reports 6(8), 717722.Google Scholar
Miyano, M., Ago, H., Saino, H., Hori, T. & Ida, K. (2010). Internally bridging water molecule in transmembrane alpha-helical kink. Current Opinion in Structural Biology 20(4), 456463.Google Scholar
Molinari, M., Eriksson, K. K., Calanca, V., Galli, C., Cresswell, P., Michalak, M. & Helenius, A. (2004). Contrasting functions of calreticulin and calnexin in glycoprotein folding and ER quality control. Molecular Cell 13(1), 125135.Google Scholar
Moon, C. P. & Fleming, K. G. (2011). Side-chain hydrophobicity scale derived from transmembrane protein folding into lipid bilayers. Proceedings of the National Academy of Sciences of the United States of America 108(25), 1017410177.Google Scholar
Moon, C. P., Kwon, S. & Fleming, K. G. (2011). Overcoming hysteresis to attain reversible equilibrium folding for outer membrane phospholipase A in phospholipid bilayers. Journal of Molecular Biology 413(2), 484494.Google Scholar
Morello, J. P., Salahpour, A., Laperriere, A., Bernier, V., Arthus, M. F., Lonergan, M., Petaja-Repo, U., Angers, S., Morin, D., Bichet, D. G. & Bouvier, M. (2000). Pharmacological chaperones rescue cell-surface expression and function of misfolded V2 vasopressin receptor mutants. Journal of Clinical Investigation 105(7), 887895.Google Scholar
Moss, K., Helm, A., Lu, Y., Bragin, A. & Skach, W. R. (1998). Coupled translocation events generate topological heterogeneity at the endoplasmic reticulum membrane. Molecular Biology of the Cell 9(9), 26812697.Google Scholar
Mu, T. W., Ong, D. S., Wang, Y. J., Balch, W. E., Yates, J. R. III, Segatori, L. & Kelly, J. W. (2008). Chemical and biological approaches synergize to ameliorate protein-folding diseases. Cell 134(5), 769781.Google Scholar
Myers, J. K., Mobley, C. K. & Sanders, C. R. (2008). The peripheral neuropathy-linked Trembler and Trembler-J mutant forms of peripheral myelin protein 22 are folding-destabilized. Biochemistry 47(40), 1062010629.Google Scholar
Naef, R. & Suter, U. (1999). Impaired intracellular trafficking is a common disease mechanism of PMP22 point mutations in peripheral neuropathies. Neurobiology of Disease 6(1), 114.Google Scholar
Nagy, J. K. & Sanders, C. R. (2002). A critical residue in the folding pathway of an integral membrane protein. Biochemistry 41(29), 90219025.Google Scholar
Nagy, J. K. & Sanders, C. R. (2004). Destabilizing mutations promote membrane protein misfolding. Biochemistry 43(1), 1925.Google Scholar
Ng, D. P. & Deber, C. M. (2010). Modulation of the oligomerization of myelin proteolipid protein by transmembrane helix interaction motifs. Biochemistry 49(32), 68966902.Google Scholar
Ng, D. P., Poulsen, B. E. & Deber, C. M. (2012). Membrane protein misassembly in disease. Biochimica et Biophysica Acta 1818(4), 11151122.Google Scholar
Notterpek, L., Ryan, M. C., Tobler, A. R. & Shooter, E. M. (1999). PMP22 accumulation in aggresomes: implications for CMT1A pathology. Neurobiology of Disease 6(5), 450460.Google Scholar
Odolczyk, N., Fritsch, J., Norez, C., Servel, N., Da Cunha, M. F., Bitam, S., Kupniewska, A., Wiszniewski, L., Colas, J., Tarnowski, K., Tondelier, D., Roldan, A., Saussereau, E. L., Melin-Heschel, P., Wieczorek, G., Lukacs, G. L., Dadlez, M., Faure, G., Herrmann, H., Ollero, M., Becq, F., Zielenkiewicz, P. & Edelman, A. (2013). Discovery of novel potent DeltaF508-CFTR correctors that target the nucleotide binding domain. EMBO Molecular Medicine 5(10), 14841501.Google Scholar
Öjemalm, K., Higuchi, T., Jiang, Y., Langel, Ü., Nilsson, I., White, S. H., Suga, H. & Von Heijne, G. (2011). Apolar surface area determines the efficiency of translocon-mediated membrane-protein integration into the endoplasmic reticulum. Proceedings of the National Academy of Sciences of the United States of America 108(31), E359E364.Google Scholar
Olzmann, J. A., Kopito, R. R. & Christianson, J. C. (2013). The mammalian endoplasmic reticulum-associated degradation system. Cold Spring Harbor Perspectives in Biology 5(9), 116.Google Scholar
Ong, D. S. & Kelly, J. W. (2011). Chemical and/or biological therapeutic strategies to ameliorate protein misfolding diseases. Current Opinion in Cell Biology 23(2), 231238.Google Scholar
Otero, J. H., Lizak, B. & Hendershot, L. M. (2010). Life and death of a BiP substrate. Seminars in Cell and Developmental Biology 21(5), 472478.Google Scholar
Otzen, D. (2003). Folding of DsbB in mixed micelles: a kinetic analysis of the stability of a bacterial membrane protein. Journal of Molecular Biology 330(4), 641649.Google Scholar
Otzen, D. E. (2011). Mapping the folding pathway of the transmembrane protein DsbB by protein engineering. Protein Engineering Design and Selection 24(1–2), 139149.Google Scholar
Ozcan, U., Yilmaz, E., Ozcan, L., Furuhashi, M., Vaillancourt, E., Smith, R. O., Gorgun, C. Z. & Hotamisligil, G. S. (2006). Chemical chaperones reduce ER stress and restore glucose homeostasis in a mouse model of type 2 diabetes. Science 313(5790), 11371140.Google Scholar
Pareek, S., Notterpek, L., Snipes, G. J., Naef, R., Sossin, W., Laliberté, J., Iacampo, S., Suter, U., Shooter, E. M. & Murphy, R. A. (1997). Neurons promote the translocation of peripheral myelin protein 22 into myelin. Journal of Neuroscience 17(20), 77547762.Google Scholar
Park, C., Zhou, S., Gilmore, J. & Marqusee, S. (2007). Energetics-based protein profiling on a proteomic scale: identification of proteins resistant to proteolysis. Journal of Molecular Biology 368(5), 14261437.Google Scholar
Partridge, A. W., Therien, A. G. & Deber, C. M. (2004). Missense mutations in transmembrane domains of proteins: phenotypic propensity of polar residues for human disease. Proteins 54(4), 648656.Google Scholar
Perlmutter, D. H. (2002). Chemical chaperones: a pharmacological strategy for disorders of protein folding and trafficking. Pediatric Research 52(6), 832836.Google Scholar
Petaja-Repo, U. E., Hogue, M., Bhalla, S., Laperriere, A., Morello, J. P. & Bouvier, M. (2002). Ligands act as pharmacological chaperones and increase the efficiency of delta opioid receptor maturation. EMBO Journal 21(7), 16281637.Google Scholar
Placone, J. & Hristova, K. (2012). Direct assessment of the effect of the Gly380Arg achondroplasia mutation on FGFR3 dimerization using quantitative imaging FRET. PLoS ONE 7(10), e46678.Google Scholar
Plaxco, K. W., Simons, K. T., Ruczinski, I. & Baker, D. (2000). Topology, stability, sequence, and length: defining the determinants of two-state protein folding kinetics. Biochemistry 39(37), 1117711183.Google Scholar
Popot, J. L. & Engelman, D. M. (1990). Membrane protein folding and oligomerization: the two-stage model. Biochemistry 29(17), 40314037.Google Scholar
Popot, J. L. & Engelman, D. M. (2000). Helical membrane protein folding, stability, and evolution. Annual Review of Biochemistry 69, 881922.Google Scholar
Popot, J. L., Trewhella, J. & Engelman, D. M. (1986). Reformation of crystalline purple membrane from purified bacteriorhodopsin fragments. EMBO Journal 5(11), 30393044.Google Scholar
Powers, E. T., Morimoto, R. I., Dillin, A., Kelly, J. W. & Balch, W. E. (2009). Biological and chemical approaches to diseases of proteostasis deficiency. Annual Review of Biochemistry 78, 959991.Google Scholar
Qu, B. H., Strickland, E. H. & Thomas, P. J. (1997). Localization and suppression of a kinetic defect in cystic fibrosis transmembrane conductance regulator folding. Journal of Biological Chemistry 272(25), 1573915744.Google Scholar
Rader, A. J., Anderson, G., Isin, B., Khorana, H. G., Bahar, I. & Klein-Seetharaman, J. (2004). Identification of core amino acids stabilizing rhodopsin. Proceedings of the National Academy of Sciences of the United States of America 101(19), 72467251.Google Scholar
Rajan, R. S., Illing, M. E., Bence, N. F. & Kopito, R. R. (2001). Specificity in intracellular protein aggregation and inclusion body formation. Proceedings of the National Academy of Sciences of the United States of America 98(23), 1306013065.Google Scholar
Ramsey, B. W., Davies, J., Mcelvaney, N. G., Tullis, E., Bell, S. C., Drevinek, P., Griese, M., Mckone, E. F., Wainwright, C. E., Konstan, M. W., Moss, R., Ratjen, F., Sermet-Gaudelus, I., Rowe, S. M., Dong, Q., Rodriguez, S., Yen, K., Ordonez, C., Elborn, J. S. & Group, V. X. S. (2011). A CFTR potentiator in patients with cystic fibrosis and the G551D mutation. New England Journal of Medicine 365(18), 16631672.Google Scholar
Riley, M., Wallace, B., Flitsch, S. & Booth, P. (1997). Slow alpha helix formation during folding of a membrane protein. Biochemistry 36(1), 192196.Google Scholar
Robben, J. H., Sze, M., Knoers, N. V. & Deen, P. M. (2006). Rescue of vasopressin V2 receptor mutants by chemical chaperones: specificity and mechanism. Molecular Biology of the Cell 17(1), 379386.Google Scholar
Roth, D. M. & Balch, W. E. (2011). Modeling general proteostasis: proteome balance in health and disease. Current Opinion in Cell Biology 23(2), 126134.Google Scholar
Rowe, S. M. & Verkman, A. S. (2013). Cystic fibrosis transmembrane regulator correctors and potentiators. Cold Spring Harbor Perspectives in Medicine 3(7), 115.Google Scholar
Ryan, M. C., Shooter, E. M. & Notterpek, L. (2002). Aggresome formation in neuropathy models based on peripheral myelin protein 22 mutations. Neurobiology of Disease 10(2), 109118.Google Scholar
Ryno, L. M., Wiseman, R. L. & Kelly, J. W. (2013). Targeting unfolded protein response signaling pathways to ameliorate protein misfolding diseases. Current Opinion in Chemical Biology 17(3), 346352.Google Scholar
Sadlish, H. & Skach, W. R. (2004). Biogenesis of CFTR and other polytopic membrane proteins: new roles for the ribosome–translocon complex. Journal of Membrane Biology 202(3), 115126.Google Scholar
Sadlish, H., Pitonzo, D., Johnson, A. E. & Skach, W. R. (2005). Sequential triage of transmembrane segments by Sec61alpha during biogenesis of a native multispanning membrane protein. Nature Structural and Molecular Biology 12(10), 870878.Google Scholar
Sakakura, M., Hadziselimovic, A., Wang, Z., Schey, K. L. & Sanders, C. R. (2011). Structural basis for the Trembler-J phenotype of Charcot-Marie-Tooth disease. Structure 19(8), 11601169.Google Scholar
Saliba, R. S., Munro, P. M., Luthert, P. J. & Cheetham, M. E. (2002). The cellular fate of mutant rhodopsin: quality control, degradation and aggresome formation. Journal of Cell Science 115(Pt 14), 29072918.Google Scholar
Sampson, H. M., Robert, R., Liao, J., Matthes, E., Carlile, G. W., Hanrahan, J. W. & Thomas, D. Y. (2011). Identification of a NBD1-binding pharmacological chaperone that corrects the trafficking defect of F508del-CFTR. Chemical Biology 18(2), 231242.Google Scholar
Sanders, C. & Myers, J. (2004). Disease-related misassembly of membrane proteins. Annual Review of Biophysics and Biomolecular Structure 33, 2551.Google Scholar
Sanders, C. R. & Mittendorf, K. F. (2011). Tolerance to changes in membrane lipid composition as a selected trait of membrane proteins. Biochemistry 50(37), 78587867.Google Scholar
Sanders, C. R. & Nagy, J. K. (2000). Misfolding of membrane proteins in health and disease: the lady or the tiger? Current Opinion in Structural Biology 10(4), 438442.Google Scholar
Sanders, C. R., Ismail-Beigi, F. & Mcenery, M. W. (2001). Mutations of peripheral myelin protein 22 result in defective trafficking through mechanisms which may be common to diseases involving tetraspan membrane proteins. Biochemistry 40(32), 94539459.Google Scholar
Sato, B. K., Schulz, D., Do, P. H. & Hampton, R. Y. (2009). Misfolded membrane proteins are specifically recognized by the transmembrane domain of the Hrd1p ubiquitin ligase. Molecular Cell 34(2), 212222.Google Scholar
Sato, S., Ward, C. L., Krouse, M. E., Wine, J. J. & Kopito, R. R. (1996). Glycerol reverses the misfolding phenotype of the most common cystic fibrosis mutation. Journal of Biological Chemistry 271(2), 635638.Google Scholar
Sawkar, A. R., Cheng, W. C., Beutler, E., Wong, C. H., Balch, W. E. & Kelly, J. W. (2002). Chemical chaperones increase the cellular activity of N370S beta-glucosidase: a therapeutic strategy for Gaucher disease. Proceedings of the National Academy of Sciences of the United States of America 99(24), 1542815433.Google Scholar
Schlebach, J. P., Kim, M. S., Joh, N. H., Bowie, J. U. & Park, C. (2011). Probing membrane protein unfolding with pulse proteolysis. Journal of Molecular Biology 406(4), 545551.Google Scholar
Schlebach, J. P., Cao, Z., Bowie, J. U. & Park, C. (2012). Revisiting the folding kinetics of bacteriorhodopsin. Protein Science 21(1), 97106.Google Scholar
Schlebach, J. P., Peng, D., Kroncke, B. M., Mittendorf, K. F., Narayan, M., Carter, B. D. & Sanders, C. R. (2013). Reversible folding of human peripheral myelin protein 22, a tetraspan membrane protein. Biochemistry 52(19), 32293241.Google Scholar
Schnell, D. J. & Hebert, D. N. (2003). Protein translocons: multifunctional mediators of protein translocation across membranes. Cell 112(4), 491505.Google Scholar
Sehgal, P., Mogensen, J. E. & Otzen, D. E. (2005). Using micellar mole fractions to assess membrane protein stability in mixed micelles. Biochimica et Biophysica Acta 1716(1), 5968.Google Scholar
Sekijima, Y., Wiseman, R. L., Matteson, J., Hammarstrom, P., Miller, S. R., Sawkar, A. R., Balch, W. E. & Kelly, J. W. (2005). The biological and chemical basis for tissue-selective amyloid disease. Cell 121(1), 7385.Google Scholar
Selkoe, D. J. (2003). Folding proteins in fatal ways. Nature 426(6968), 900904.Google Scholar
Sheppard, D. N., Rich, D. P., Ostedgaard, L. S., Gregory, R. J., Smith, A. E. & Welsh, M. J. (1993). Mutations in CFTR associated with mild-disease-form Cl-channels with altered pore properties. Nature 362(6416), 160164.Google Scholar
Skach, W. R., Calayag, M. C. & Lingappa, V. R. (1993). Evidence for an alternate model of human P-glycoprotein structure and biogenesis. Journal of Biological Chemistry 268(10), 69036908.Google Scholar
Skach, W. R., Shi, L. B., Calayag, M. C., Frigeri, A., Lingappa, V. R. & Verkman, A. S. (1994). Biogenesis and transmembrane topology of the CHIP28 water channel at the endoplasmic reticulum. Journal of Cell Biology 125(4), 803815.Google Scholar
Snapp, E. L., Reinhart, G. A., Bogert, B. A., Lippincott-Schwartz, J. & Hegde, R. S. (2004). The organization of engaged and quiescent translocons in the endoplasmic reticulum of mammalian cells. Journal of Cell Biology 164(7), 9971007.Google Scholar
Sontag, E. M., Vonk, W. I. & Frydman, J. (2014). Sorting out the trash: the spatial nature of eukaryotic protein quality control. Current Opinion in Cell Biology 26C, 139146.Google Scholar
Srinivasan, R., Henderson, B. J., Lester, H. A. & Richards, C. I. (2014). Pharmacological chaperoning of nAChRs: a therapeutic target for Parkinson's disease. Pharmacological Research 83, 2029.Google Scholar
Stanley, A. & Fleming, K. (2008). The process of folding proteins into membranes: challenges and progress. Archives of Biochemistry and Biophysics 469(1), 4666.Google Scholar
Sun, F., Zhang, R., Gong, X., Geng, X., Drain, P. F. & Frizzell, R. A. (2006). Derlin-1 promotes the efficient degradation of the cystic fibrosis transmembrane conductance regulator (CFTR) and CFTR folding mutants. Journal of Biological Chemistry 281(48), 3685636863.Google Scholar
Sung, C. H., Schneider, B. G., Agarwal, N., Papermaster, D. S. & Nathans, J. (1991). Functional heterogeneity of mutant rhodopsins responsible for autosomal dominant retinitis pigmentosa. Proceedings of the National Academy of Sciences of the United States of America 88(19), 88408844.Google Scholar
Suter, U. & Snipes, G. J. (1995). Biology and genetics of hereditary motor and sensory neuropathies. Annual Review of Neuroscience 18, 4575.Google Scholar
Suter, U., Moskow, J. J., Welcher, A. A., Snipes, G. J., Kosaras, B., Sidman, R. L., Buchberg, A. M. & Shooter, E. M. (1992). A leucine-to-proline mutation in the putative first transmembrane domain of the 22-kDa peripheral myelin protein in the trembler-J mouse. Proceedings of the National Academy of Sciences of the United States of America 89(10), 43824386.Google Scholar
Swanton, E., High, S. & Woodman, P. (2003). Role of calnexin in the glycan-independent quality control of proteolipid protein. EMBO Journal 22(12), 29482958.Google Scholar
Szeto, J., Kaniuk, N. A., Canadien, V., Nisman, R., Mizushima, N., Yoshimori, T., Bazett-Jones, D. P. & Brumell, J. H. (2006). ALIS are stress-induced protein storage compartments for substrates of the proteasome and autophagy. Autophagy 2(3), 189199.Google Scholar
Tamarappoo, B. K. & Verkman, A. S. (1998). Defective aquaporin-2 trafficking in nephrogenic diabetes insipidus and correction by chemical chaperones. Journal of Clinical Investigation 101(10), 22572267.Google Scholar
Tamborero, S., Vilar, M., Martinez-Gil, L., Johnson, A. E. & Mingarro, I. (2011). Membrane insertion and topology of the translocating chain-associating membrane protein (TRAM). Journal of Molecular Biology 406(4), 571582.Google Scholar
Tao, Y. X. & Conn, P. M. (2014). Chaperoning G Protein-coupled receptors: from cell biology to therapeutics. Endocrione Reviews er20131121.Google Scholar
Tardiff, D. F., Jui, N. T., Khurana, V., Tambe, M. A., Thompson, M. L., Chung, C. Y., Kamadurai, H. B., Kim, H. T., Lancaster, A. K., Caldwell, K. A., Caldwell, G. A., Rochet, J. C., Buchwald, S. L. & Lindquist, S. (2013). Yeast reveal a “druggable” Rsp5/Nedd4 network that ameliorates alpha-synuclein toxicity in neurons. Science 342(6161), 979983.Google Scholar
Taylor, R. C. & Dillin, A. (2011). Aging as an event of proteostasis collapse. Cold Spring Harbor Perspectives in Biology 3(5), 117.Google Scholar
Taylor, S. C., Ferguson, A. D., Bergeron, J. J. & Thomas, D. Y. (2004). The ER protein folding sensor UDP-glucose glycoprotein-glucosyltransferase modifies substrates distant to local changes in glycoprotein conformation. Nature Structural and Molecular Biology 11(2), 128134.Google Scholar
Teasdale, R. D. & Jackson, M. R. (1996). Signal-mediated sorting of membrane proteins between the endoplasmic reticulum and the Golgi apparatus. Annual Review of Cell and Developmental Biology 12, 2754.Google Scholar
Therien, A. G., Grant, F. E. & Deber, C. M. (2001). Interhelical hydrogen bonds in the CFTR membrane domain. Nature Structural Biology 8(7), 597601.Google Scholar
Tobler, A. R., Notterpek, L., Naef, R., Taylor, V., Suter, U. & Shooter, E. M. (1999). Transport of Trembler-J mutant peripheral myelin protein 22 is blocked in the intermediate compartment and affects the transport of the wild-type protein by direct interaction. Journal of Neuroscience 19(6), 20272036.Google Scholar
Tobler, A. R., Liu, N., Mueller, L. & Shooter, E. M. (2002). Differential aggregation of the Trembler and Trembler J mutants of peripheral myelin protein 22. Proceedings of the National Academy of Sciences of the United States of America 99(1), 483488.Google Scholar
Van Den Berg, B., Clemons, W. M., Collinson, I., Modis, Y., Hartmann, E., Harrison, S. C. & Rapoport, T. A. (2004). X-ray structure of a protein-conducting channel. Nature 427(6969), 3644.Google Scholar
Van Goor, F., Hadida, S., Grootenhuis, P. D., Burton, B., Cao, D., Neuberger, T., Turnbull, A., Singh, A., Joubran, J., Hazlewood, A., Zhou, J., Mccartney, J., Arumugam, V., Decker, C., Yang, J., Young, C., Olson, E. R., Wine, J. J., Frizzell, R. A., Ashlock, M. & Negulescu, P. (2009). Rescue of CF airway epithelial cell function in vitro by a CFTR potentiator, VX-770. Proceedings of the National Academy of Sciences of the United States of America 106(44), 1882518830.Google Scholar
Van Goor, F., Hadida, S., Grootenhuis, P. D., Burton, B., Stack, J. H., Straley, K. S., Decker, C. J., Miller, M., Mccartney, J., Olson, E. R., Wine, J. J., Frizzell, R. A., Ashlock, M. & Negulescu, P. A. (2011). Correction of the F508del-CFTR protein processing defect in vitro by the investigational drug VX-809. Proceedings of the National Academy of Sciences of the United States of America 108(46), 1884318848.Google Scholar
Van Meer, G., Voelker, D. R. & Feigenson, G. W. (2008). Membrane lipids: where they are and how they behave. Nature Reviews Molecular Cell Biology 9(2), 112124.Google Scholar
Veerappan, A., Cymer, F., Klein, N. & Schneider, D. (2011). The tetrameric α-helical membrane protein GlpF unfolds via a dimeric folding intermediate. Biochemistry 50(47), 1022310230.Google Scholar
Vembar, S. S. & Brodsky, J. L. (2008). One step at a time: endoplasmic reticulum-associated degradation. Nature Reviews Molecular Cell Biology 9(12), 944957.Google Scholar
Virkki, M. T., Agrawal, N., Edsbacker, E., Cristobal, S., Elofsson, A. & Kauko, A. (2014). Folding of Aquaporin 1: multiple evidence that helix 3 can shift out of the membrane core. Protein Science 23(7), 981992.Google Scholar
Von Heijne, G. (1986). The distribution of positively charged residues in bacterial inner membrane proteins correlates with the trans-membrane topology. EMBO Journal 5(11), 30213027.Google Scholar
Von Heijne, G. (1992). Membrane protein structure prediction. Hydrophobicity analysis and the positive-inside rule. Journal of Molecular Biology 225(2), 487494.Google Scholar
Von Heijne, G. (2006). Membrane-protein topology. Nature Reviews Molecular Cell Biology 7(12), 909918.Google Scholar
Wang, X., Venable, J., Lapointe, P., Hutt, D. M., Koulov, A. V., Coppinger, J., Gurkan, C., Kellner, W., Matteson, J., Plutner, H., Riordan, J. R., Kelly, J. W., Yates, J. R. & Balch, W. E. (2006). Hsp90 cochaperone Aha1 downregulation rescues misfolding of CFTR in cystic fibrosis. Cell 127(4), 803815.Google Scholar
Ward, C. L. & Kopito, R. R. (1994). Intracellular turnover of cystic fibrosis transmembrane conductance regulator. Inefficient processing and rapid degradation of wild-type and mutant proteins. Journal of Biological Chemistry 269(41), 2571025718.Google Scholar
Warschawski, D. E., Arnold, A. A., Beaugrand, M., Gravel, A., Chartrand, E. & Marcotte, I. (2011). Choosing membrane mimetics for NMR structural studies of transmembrane proteins. Biochimica et Biophysica Acta 1808(8), 19571974.Google Scholar
Welch, W. J. & Brown, C. R. (1996). Influence of molecular and chemical chaperones on protein folding. Cell Stress and Chaperones 1(2), 109115.Google Scholar
Welsh, M. J. & Smith, A. E. (1993). Molecular mechanisms of CFTR chloride channel dysfunction in cystic fibrosis. Cell 73(7), 12511254.Google Scholar
White, S. (2009). Biophysical dissection of membrane proteins. Nature 459(7245), 344346.Google Scholar
White, S. H. (2003). Translocons, thermodynamics, and the folding of membrane proteins. FEBS Letters 555(1), 116121.Google Scholar
White, S. H. & Von Heijne, G. (2008). How translocons select transmembrane helices. Annual Review of Biophysics 37, 2342.Google Scholar
Wigley, W. C., Fabunmi, R. P., Lee, M. G., Marino, C. R., Muallem, S., Demartino, G. N. & Thomas, P. J. (1999). Dynamic association of proteasomal machinery with the centrosome. Journal of Cell Biology 145(3), 481490.Google Scholar
Wimley, W. C. & White, S. H. (1996). Experimentally determined hydrophobicity scale for proteins at membrane interfaces. Nature Structural Biology 3(10), 842848.Google Scholar
Wiseman, R. L. & Balch, W. E. (2005). A new pharmacology–drugging stressed folding pathways. Trends in Molecular Medicine 11(8), 347350.Google Scholar
Wiseman, R. L., Koulov, A., Powers, E., Kelly, J. W. & Balch, W. E. (2007a). Protein energetics in maturation of the early secretory pathway. Current Opinion in Cell Biology 19(4), 359367.Google Scholar
Wiseman, R. L., Powers, E. T., Buxbaum, J. N., Kelly, J. W. & Balch, W. E. (2007b). An adaptable standard for protein export from the endoplasmic reticulum. Cell 131(4), 809821.Google Scholar
Xia, K., Manning, M., Hesham, H., Lin, Q., Bystroff, C. & Colón, W. (2007). Identifying the subproteome of kinetically stable proteins via diagonal 2D SDS/PAGE. Proceedings of the National Academy of Sciences of the United States of America 104(44), 1732917334.Google Scholar
Xiong, X., Bragin, A., Widdicombe, J. H., Cohn, J. & Skach, W. R. (1997). Structural cues involved in endoplasmic reticulum degradation of G85E and G91R mutant cystic fibrosis transmembrane conductance regulator. Journal of Clinical Investigation 100(5), 10791088.Google Scholar
Yamamoto, K., Fujii, R., Toyofuku, Y., Saito, T., Koseki, H., Hsu, V. W. & Aoe, T. (2001). The KDEL receptor mediates a retrieval mechanism that contributes to quality control at the endoplasmic reticulum. EMBO Journal 20(12), 30823091.Google Scholar
You, M., Spangler, J., Li, E., Han, X., Ghosh, P. & Hristova, K. (2007). Effect of pathogenic cysteine mutations on FGFR3 transmembrane domain dimerization in detergents and lipid bilayers. Biochemistry 46(39), 1103911046.Google Scholar
Yu, Z., Sawkar, A. R. & Kelly, J. W. (2007). Pharmacologic chaperoning as a strategy to treat Gaucher disease. FEBS Journal 274(19), 49444950.Google Scholar
Zhou, H. X. & Cross, T. A. (2013). Influences of membrane mimetic environments on membrane protein structures. Annual Review of Biophysics 42, 361392.Google Scholar
Zocher, M., Zhang, C., Rasmussen, S. G., Kobilka, B. K. & Muller, D. J. (2012). Cholesterol increases kinetic, energetic, and mechanical stability of the human beta2-adrenergic receptor. Proceedings of the National Academy of Sciences of the United States of America 109(50), E3463E3472.Google Scholar