Hostname: page-component-8448b6f56d-qsmjn Total loading time: 0 Render date: 2024-04-15T16:01:10.603Z Has data issue: false hasContentIssue false

NMR structures of membrane proteins in phospholipid bilayers

Published online by Cambridge University Press:  17 July 2014

Jasmina Radoicic
Affiliation:
Department of Chemistry and Biochemistry, University of California, San Diego, La Jolla, CA 92093, USA
George J. Lu
Affiliation:
Department of Chemistry and Biochemistry, University of California, San Diego, La Jolla, CA 92093, USA
Stanley J. Opella*
Affiliation:
Department of Chemistry and Biochemistry, University of California, San Diego, La Jolla, CA 92093, USA
*
*Author for correspondence: S. J. Opella, Department of Chemistry and Biochemistry, University of California, San Diego, La Jolla, CA 92093, USA. Tel.: 858 822-4820; Email: sopella@ucsd.edu

Abstract

Membrane proteins have always presented technical challenges for structural studies because of their requirement for a lipid environment. Multiple approaches exist including X-ray crystallography and electron microscopy that can give significant insights into their structure and function. However, nuclear magnetic resonance (NMR) is unique in that it offers the possibility of determining the structures of unmodified membrane proteins in their native environment of phospholipid bilayers under physiological conditions. Furthermore, NMR enables the characterization of the structure and dynamics of backbone and side chain sites of the proteins alone and in complexes with both small molecules and other biopolymers. The learning curve has been steep for the field as most initial studies were performed under non-native environments using modified proteins until ultimately progress in both techniques and instrumentation led to the possibility of examining unmodified membrane proteins in phospholipid bilayers under physiological conditions. This review aims to provide an overview of the development and application of NMR to membrane proteins. It highlights some of the most significant structural milestones that have been reached by NMR spectroscopy of membrane proteins, especially those accomplished with the proteins in phospholipid bilayer environments where they function.

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

5. References

Acharya, R., Carnevale, V., Fiorin, G., Levine, B. G., Polishchuk, A. L., Balannik, V., Samish, I., Lamb, R. A., Pinto, L. H., Degrado, W. F. & Klein, M. L. (2010). Structure and mechanism of proton transport through the transmembrane tetrameric M2 protein bundle of the influenza A virus. Proceedings of the National Academy of Sciences of the United State of America 107, 1507515080.Google Scholar
Ader, C., Schneider, R., Seidel, K., Etzkorn, M., Becker, S. & Baldus, M. (2008). Structural rearrangements of membrane proteins probed by water-edited solid-state NMR spectroscopy. Journal of the American Chemical Society 131, 170176.Google Scholar
Ahuja, S., Jahr, N., Im, S.-C., Vivekanandan, S., Popovych, N., Le Clair, S. V., Huang, R., Soong, R., Xu, J., Yamamoto, K., Nanga, R. P., Bridges, A., Waskell, L. & Ramamoorthy, A. (2013). A model of the membrane-bound cytochrome b5-cytochrome P450 complex from NMR and mutagenesis data. Journal of Biological Chemistry 288, 2208022095.Google Scholar
Almeida, F. C. & Opella, S. J. (1997). fd coat protein structure in membrane environments: structural dynamics of the loop between the hydrophobic trans-membrane helix and the amphipathic in-plane helix. Journal of Molecular Biology 270, 481495.Google Scholar
Andreas, L. B., Eddy, M. T., Chou, J. J. & Griffin, R. G. (2012). MAS NMR of the drug resistant S31N M2 proton transporter from influenza A. Journal of the American Chemical Society 134, 72157218.Google Scholar
Anfinsen, C. B. (1973). Principles that govern the folding of protein chains. Science 181, 223230.Google Scholar
Banerjee, R. & Datta, A. (1983). Proteoliposome as the model for the study of membrane-bound enzymes and transport proteins. Molecular Cell Biochemistry 50, 315.Google Scholar
Bauer, D. R., Opella, S. J., Nelson, D. J. & Pecora, R. (1975). Depolarized light scattering and carbon nuclear resonance measurements of the isotropic rotational correlation time of muscle calcium binding protein. Journal of the American Chemical Society 97, 25802582.Google Scholar
Bechinger, B., Kim, Y., Chirlian, L. E., Gesell, J., Neumann, J. M., Montal, M., Tomich, J., Zasloff, M. & Opella, S. J. (1991). Orientations of amphipathic helical peptides in membrane bilayers determined by solid-state NMR spectroscopy. Journal of Biomolecular NMR 1, 167173.Google Scholar
Blume, A., Wittebort, R. J., Das Gupta, S. K. & Griffin, R. G. (1982). Phase equilibria, molecular conformation, and dynamics in phosphatidylcholine/phosphatidylethanolamine bilayers. Biochemistry 21, 62436253.Google Scholar
Bogusky, M. J., Schiksnis, R. A., Leo, G. C. & Opella, S. J. (1987). Protein backbone dynamics by solid state and solution 15N NMR spectroscopy. Journal of Magnetic Resonance 72, 186190.Google Scholar
Bondarenko, V., Mowrey, D., Tillman, T., Cui, T., Liu, L. T., Xu, Y. & Tang, P. (2012). NMR structures of the transmembrane domains of the α4β2 nAChR. Biochimica et Biophysica Acta – Biomembranes 1818, 12611268.Google Scholar
Bondarenko, V., Tillman, T., Xu, Y. & Tang, P. (2010). NMR structure of the transmembrane domain of the n-acetylcholine receptor β2 subunit. Biochimica et Biophysica Acta – Biomembranes 1798, 16081614.Google Scholar
Bosch, C., Brown, L. R. & Wuthrich, K. (1980). Physicochemical characterization of glucagon-containing lipid micelles. Biochimica et Biophysica Acta – Biomembranes 603, 298312.Google Scholar
Buck-Koehntop, B. A., Mascioni, A., Buffy, J. J. & Veglia, G. (2005). Structure, dynamics, and membrane topology of stannin: a mediator of neuronal cell apoptosis induced by trimethyltin chloride. Journal of Molecular Biology 354, 652665.Google Scholar
Cady, S. D., Schmidt-Rohr, K., Wang, J., Soto, C. S., Degrado, W. F. & Hong, M. (2010). Structure of the amantadine binding site of influenza M2 proton channels in lipid bilayers. Nature 463, 689692.Google Scholar
Can, T. V., Sharma, M., Hung, I., GOR'KOV, P. L., Brey, W. W. & Cross, T. A. (2012). Magic angle spinning and oriented sample solid-state NMR structural restraints combine for influenza A M2 protein functional insights. Journal of the American Chemical Society 134, 90229025.Google Scholar
Cao, J., Burke, J. E. & Dennis, E. A. (2013). Using hydrogen/deuterium exchange mass spectrometry to define the specific interactions of the phospholipase A2 superfamily with lipid substrates, inhibitors, and membranes. The Journal of Biological Chemistry 288, 18061813.Google Scholar
Cegelski, L. (2013). REDOR NMR for drug discovery. Bioorganic and Medicinal Chemistry Letters 23, 57675775.Google Scholar
Chan, J. C. C. & Tycko, R. (2003). Recoupling of chemical shift anisotropies in solid-state NMR under high-speed magic-angle spinning and in uniformly 13C-labeled systems. Journal of Chemical Physics 118, 83788389.Google Scholar
Chan, S. I., Feigenson, G. W. & Seiter, C. H. (1971). Nuclear relaxation studies of lecithin bilayers. Nature 231, 110112.Google Scholar
Chan, S. I., Seiter, C. H. & Feigenson, G. W. (1972). Anisotropic and restricted molecular motion in lecithin bilayers. Biochemical and Biophysical Research Communications 46, 14881492.Google Scholar
Chan, S. I., Sheetz, M. P., Seiter, C. H., Feigenson, G. W., Hsu, M. C., Lau, A. & Yau, A. (1973). Nuclear magnetic resonance studies of the structure of model membrane systems: the effect of surface curvature. Annals of the New York Academy of Sciences 222, 499522.Google Scholar
Chen, M. Y., Maldarelli, F., Karczewski, M. K., Willey, R. L. & Strebel, K. (1993). Human immunodeficiency virus type 1 Vpu protein induces degradation of CD4 in vitro: the cytoplasmic domain of CD4 contributes to Vpu sensitivity. Journal of Virological 67, 38773884.Google Scholar
Cherry, R. J. (1978). Measurement of protein rotational diffusion in membranes by flash photolysis. Methods Enzymology 54, 4761.Google Scholar
Comellas, G. & Rienstra, C. M. (2013). Protein structure determination by magic-angle spinning solid-state NMR, and insights into the formation, structure, and stability of amyloid fibrils. Annual Review of Biophysics 42, 515536.Google Scholar
Cone, R. A. (1972). Rotational diffusion of rhodopsin in the visual receptor membrane. Nature New Biology 236, 3943.Google Scholar
Cook, G. A., Dawson, L. A., Tian, Y. & Opella, S. J. (2013). Three-dimensional structure and interaction studies of Hepatitis C Virus p7 in 1,2-dihexanoyl-sn-glycero-3-phospholcholine by solution nuclear magnetic resonance. Biochemistry 52, 52955303.Google Scholar
Cross, T., Murray, D. & Watts, A. (2013). Helical membrane protein conformations and their environment. European Biophysics Journal 42, 731755.Google Scholar
Cross, T. A. & Opella, S. J. (1979). NMR of fd coat protein. Journal of Supramolecular Structure 11, 139145.Google Scholar
Cross, T. A. & Opella, S. J. (1980). Structural properties of fd coat protein in sodium dodecyl sulfate micelles. Biochemical and Biophysical Research Communications 92, 478484.Google Scholar
Cross, T. A. & Opella, S. J. (1985). Protein structure by solid state nuclear magnetic resonance. Residues 40 to 45 of bacteriophage fd coat protein. Journal of Molecular Biology 182, 367381.Google Scholar
Das, B. B., Nothnagel, H. J., Lu, G. J., Son, W. S., Tian, Y., Marassi, F. M. & Opella, S. J. (2012). Structure determination of a membrane protein in proteoliposomes. Journal of the American Chemical Society 134, 20472056.Google Scholar
De Angelis, A. A., Howell, S. C., Nevzorov, A. A. & Opella, S. J. (2006). Structure determination of a membrane protein with two trans-membrane helices in aligned phospholipid bicelles by solid-state NMR spectroscopy. Journal of the American Chemical Society 128, 1225612267.Google Scholar
De Angelis, A. A., Nevzorov, A. A., Park, S. H., Howell, S. C., Mrse, A. A. & Opella, S. J. (2004). High-resolution NMR spectroscopy of membrane proteins in aligned bicelles. Journal of the American Chemical Society 126, 1534015341.Google Scholar
De Angelis, A. A. & Opella, S. J. (2007). Bicelle samples for solid-state NMR of membrane proteins. Nature Protocols 2, 23322338.Google Scholar
Devaux, P. & McConnell, H. M. (1973). Equality of the rates of lateral diffusion of phosphatidylethanolamine and phosphatidyl choline spin labels in rabbit sarcoplasmic reticulum. Annals of the New York Academy Science 222, 489498.Google Scholar
Devaux, P. & Mcdonnell, H. M. (1972). Lateral diffusion in spin-labeled phosphatidylcholine multilayers. Journal of the American Chemical Society 94, 44754481.Google Scholar
Dickerson, R. E., Reddy, J., Pinkerton, M. & Steinrauf, L. K. (1962). A 6 angstrom model of triclinic lysozyme. Nature 196, 1178.Google Scholar
Do Hoa, Q., Wittlich, M., Glück Julian, M., Möckel, L., Willbold, D., Koenig Bernd, W. & Heise, H. (2013). Full-length Vpu and human CD4(372–433) in phospholipid bilayers as seen by magic angle spinning NMR. Biological Chemistry 394, 14531463.Google Scholar
Durr, H. N. U., Soong, R. & Ramamoorthy, A. (2013). When detergent meets bilayer: birth and coming of age of lipid bicelles. Progress in NMR Spectroscopy 69, 122.Google Scholar
Durr, U. H., Waskell, L. & Ramamoorthy, A. (2007). The cytochromes P450 and b5 and their reductases – promising targets for structural studies by advanced solid-state NMR spectroscopy. Biochimica et Biophysica Acta – Biomembranes 1768, 32353259.Google Scholar
Edidin, M. (1974). Rotational and translational diffusion in membranes. Annual Review of Biophysics and Bioengineering 3, 179201.Google Scholar
Etzkorn, M., Martell, S., Andronesi, O. C., Seidel, K., Engelhard, M. & Baldus, M. (2007). Secondary structure, dynamics, and topology of a seven-helix receptor in native membranes, studied by solid-state NMR spectroscopy. Angewandte Chemie (International ed. English) 46, 459462.Google Scholar
Finer, E. G., Flook, A. G. & Hauser, H. (1972). The nature and origin of the NMR spectrum of unsonicated and sonicated aqueous egg yolk lecithin dispersions. Biochimica et Biophysica Acta – Biomembranes 260, 5969.Google Scholar
Foster, T. L., Thompson, G. S., Kalverda, A. P., Kankanala, J., Bentham, M., Wetherill, L. F., Thompson, J., Barker, A. M., Clarke, D., Noerenberg, M., Pearson, A. R., Rowlands, D. J., Homans, S. W., Harris, M., Foster, R. & Griffin, S. (2013). Structure-guided design affirms inhibitors of hepatitis C virus p7 as a viable class of antivirals targeting virion release. Hepatology 59, 408422.Google Scholar
Franks, W. T., Linden, A. H., Kunert, B., Van Rossum, B. J. & Oschkinat, H. (2012). Solid-state magic-angle spinning NMR of membrane proteins and protein-ligand interactions. European Journal of Cell Biology 91, 340348.Google Scholar
Früh, V., Zhou, Y., Chen, D., Loch, C., Ab, E., Grinkova, Y. N., Verheij, H., Sligar, S. G., Bushweller, J. H. & Siegal, G. (2010). Application of fragment-based drug discovery to membrane proteins: identification of ligands of the integral membrane enzyme DsbB. Chemistry and Biology 17, 881891.Google Scholar
Gall, C. M., Cross, T. A., Diverdi, J. A. & Opella, S. J. (1982). Protein dynamics by solid-state NMR: aromatic rings of the coat protein in fd bacteriophage. Proceedings of the National Academy Science of the United State of America 79, 101105.Google Scholar
Gautier, A., Kirkpatrick, J. P. & Nietlispach, D. (2008). Solution-state NMR spectroscopy of a seven-helix transmembrane protein receptor: backbone assignment, secondary structure, and dynamics. Angewandte Chemie (International Edition in English) 47, 72977300.Google Scholar
Gautier, A., Mott, H. R., Bostock, M. J., Kirkpatrick, J. P. & Nietlispach, D. (2010). Structure determination of the seven-helix transmembrane receptor sensory rhodopsin II by solution NMR spectroscopy. Nature Structural and Molecular Biology 17, 768774.Google Scholar
Goldbourt, A., Gross, B. J., Day, L. A. & Mcdermott, A. (2007). Filamentous phage studied by magic-angle spinning NMR: resonance assignment and secondary structure of the coat protein in Pf1. Journal of the American Chemical Society 129, 23382344.Google Scholar
Goncalves, J. A., Eilers, M., South, K., Opefi, C. A., Laissue, P., Reeves, P. J. & Smith, S. O. (2013). Magic angle spinning nuclear magnetic resonance spectroscopy of G protein-coupled receptors. Methods in Enzymology 522, 365389.Google Scholar
Gonzalez, M. E. & Carrasco, L. (2003). Viroporins. FEBS Letters 552, 2834.Google Scholar
Gustavsson, M., Traaseth, N. J. & Veglia, G. (2012). Probing ground and excited states of phospholamban in model and native lipid membranes by magic angle spinning NMR spectroscopy. Biochimica et Biophysica Acta – Biomembranes 1818, 146153.Google Scholar
Gustavsson, M., Verardi, R., Mullen, D. G., Mote, K. R., Traaseth, N. J., Gopinath, T. & Veglia, G. (2013). Allosteric regulation of SERCA by phosphorylation-mediated conformational shift of phospholamban. Proceedings of the National Academy of Sciences of the United State of America 110, 1733817343.Google Scholar
Gutowsky, H. S. & Pake, G. E. (1950). Structural investigations by means of nuclear magnetism. II. Hindered rotation in solids. Journal of Chemical Physics 18, 162170.Google Scholar
Henderson, R. & Unwin, P. N. T. (1975). Three-dimensional model of purple membrane obtained by electron microscopy. Nature 257, 2832.Google Scholar
Henry, G. D. & Sykes, B. D. (1990). Structure and dynamics of detergent-solubilized M13 coat protein (an integral membrane protein) determined by 13C and 15N nuclear magnetic resonance spectroscopy. Biochemistry and Cell Biology 68, 318329.Google Scholar
Herzfeld, J. & Berger, A. E. (1980). Sideband intensities in NMR spectra of samples spinning at the magic angle. Journal of Chemical Physics 73, 60216030.Google Scholar
Herzfeld, J., Mulliken, C. M., Siminovitch, D. J. & Griffin, R. G. (1987). Contrasting molecular dynamics in red and purple membrane fractions of the Halobacterium halobium. Biophysical Journal 52, 855858.Google Scholar
Hong, M., Zhang, Y. & Hu, F. (2012). Membrane protein structure and dynamics from NMR spectroscopy. Annual Review of Physical Chemistry 63, 124.Google Scholar
Howell, S. C., Mesleh, M. F. & Opella, S. J. (2005). NMR structure determination of a membrane protein with two transmembrane helices in micelles: MerF of the bacterial mercury detoxification system. Biochemistry 44, 51965206.Google Scholar
Hu, F., Luo, W. & Hong, M. (2010). Mechanisms of proton conduction and gating in influenza M2 proton channels from solid-state NMR. Science 330, 505508.Google Scholar
Hu, J., Asbury, T., Achuthan, S., Li, C., Bertram, R., Quine, J. R., Fu, R. & Cross, T. A. (2007). Backbone structure of the amantadine-blocked trans-membrane domain M2 proton channel from influenza A virus. Biophysical Journal 92, 43354343.Google Scholar
Hu, J., Fu, R., Nishimura, K., Zhang, L., Zhou, H.-X., Busath, D. D., Vijayvergiya, V. & Cross, T. A. (2006). Histidines, heart of the hydrogen ion channel from influenza A virus: toward an understanding of conductance and proton selectivity. Proceedings of the National Academy of Sciences of the United State of America 103, 68656870.Google Scholar
Hubbell, W. L. & McConnell, H. M. (1971). Molecular motion in spin-labeled phospholipids and membranes. Journal of the American Chemical Society 93, 314326.Google Scholar
Inaba, K., Murakami, S., Nakagawa, A., Iida, H., Kinjo, M., Ito, K. & Suzuki, M. (2009). Dynamic nature of disulphide bond formation catalysts revealed by crystal structures of DsbB. EMBO Journal 28, 779791.Google Scholar
Inaba, K., Murakami, S., Suzuki, M., Nakagawa, A., Yamashita, E., Okada, K. & Ito, K. (2006). Crystal structure of the DsbB-DsbA complex reveals a mechanism of disulfide bond generation. Cell 127, 789801.Google Scholar
Judge, P. J. & Watts, A. (2011). Recent contributions from solid-state NMR to the understanding of membrane protein structure and function. Current Opinion in Chemical Biology 15, 690695.Google Scholar
Kendrew, J. C., Bodo, G., Dintzis, H. M., Parrish, R. G., Wyckoff, H. W. & Philips, D. C. (1958). A three-dimensional model of the myoglobin molecule obtained by X-ray analysis. Nature 181, 666.Google Scholar
Keniry, M. A., Gutowsky, H. S. & Oldfield, E. (1984). Surface dynamics of the integral membrane protein bacteriorhodopsin. Nature 307, 383386.Google Scholar
Keough, K. M., Oldfield, E. & Chapman, D. (1973). Carbon-13 and proton nuclear magnetic resonance of unsonicated model and mitochondrial membranes. Chemistry and Physics of Lipids 10, 3750.Google Scholar
Kinsey, R. A., Kintanar, A. & Oldfield, E. (1981a). Dynamics of amino acid side chains in membrane proteins by high field solid state deuterium nuclear magnetic resonance spectroscopy. Phenylalanine, tyrosine, and tryptophan. Journal of Biological Chemistry 256, 90289036.Google Scholar
Kinsey, R. A., Kintanar, A., Tsai, M. D., Smith, R. L., Janes, N. & Oldfield, E. (1981b). First observation of amino acid side chain dynamics in membrane proteins using high field deuterium nuclear magnetic resonance spectroscopy. The Journal of Biological Chemistry 256, 41464149.Google Scholar
Knight, M. J., Felli, I. C., Pierattelli, R., Emsley, L. & Pintacuda, G. (2013). Magic angle spinning NMR of paramagnetic proteins. Accounts of Chemical Research 46, 21082116.Google Scholar
Knox, R. W., Lu, G. J., Opella, S. J. & Nevzorov, A. A. (2010). A resonance assignment method for oriented-sample solid-state NMR of proteins. Journal of the American Chemical Society 132, 82558257.Google Scholar
Kruger-Koplin, R. D., Sorgen, P. L., Druieger-Koplin, S. T., Rivera-Torres, I. O., Cahill, S. M., Krulwich, T. A. & Girvin, M. E. (2004). An evaluation of detergents for NMR structural studies of membrane proteins. Journal of Biomolecular NMR 28, 19801987.Google Scholar
Lamberth, S., Schmid, H., Muenchbach, M., Vorherr, T., Krebs, J., Carafoli, E. & Griesinger, C. (2000). NMR solution structure of phospholamban. Helvetica Chimica Acta 83, 21412152.Google Scholar
Lauterwein, J., Bosch, C., Brown, L. R. & Wuthrich, K. (1979). Physicochemical studies of the protein-lipid interactions in melittin-containing micelles. Biochimica et Biophysica Acta – Biomembranes 556, 244264.Google Scholar
Leo, G. C., Colnago, L. A., Valentine, K. G. & Opella, S. J. (1987). Dynamics of fd coat protein in lipid bilayers. Biochemistry 26, 854862.Google Scholar
Lewis, B. A., Harbison, G. S., Herzfeld, J. & Griffin, R. G. (1985). NMR structural analysis of a membrane protein: bacteriorhodopsin peptide backbone orientation and motion. Biochemistry 24, 46714679.Google Scholar
Li, D., Lyons, J. A., Pye, V. E., Vogeley, L., Aragao, D., Kenyon, C. P., Shah, S. T. A., Doherty, C., Aherne, M. & Caffrey, M. (2013). Crystal structure of the integral membrane diacylglycerol kinase. Nature 497, 521524.Google Scholar
Li, Y., Berthold, D. A., Gennis, R. B. & Rienstra, C. M. (2008). Chemical shift assignment of the transmembrane helices of DsbB, a 20-kDa integral membrane enzyme, by 3D magic-angle spinning NMR spectroscopy. Protein Science 17, 199204.Google Scholar
Liao, S. Y., Fritzsching, K. J. & Hong, M. (2013). Conformational analysis of the full-length M2 protein of the influenza a virus using solid-state NMR. Protein Science 22, 16231638.Google Scholar
Lichtenberg, D., Petersen, N. O., Girardet, J. L., Kainosho, M., Kroon, P. A., Seiter, C. H., Feigenson, G. W. & Chan, S. I. (1975). The interpretation of proton magnetic resonance linewidths for lecithin dispersions. Effect of particle size and chain packing. Biochimica et Biophysica Acta – Biomembranes 382, 1021.Google Scholar
Lombard, J., Lopez-Garcia, P. & Moreira, D. (2012). The early evolution of lipid membranes and the three domains of life. Nature Reviews Microbiology 10, 507515.Google Scholar
Lu, G. J. & Opella, S. J. (2014). Resonance assignments of a membrane protein in phospholipid bilayers by combining multiple strategies of oriented sample solid-state NMR. Journal of Biomolecular NMR 58, 6981.Google Scholar
Lu, G. J., Park, S. H. & Opella, S. J. (2012). Improved 1H amide resonance line narrowing in oriented sample solid-state NMR of membrane proteins in phospholipid bilayers. Journal of Magnetic Resonance 220, 5461.Google Scholar
Lu, G. J., Son, W. S. & Opella, S. J. (2011). A general assignment method for oriented sample (OS) solid-state NMR of proteins based on the correlation of resonances through heteronuclear dipolar couplings in samples aligned parallel and perpendicular to the magnetic field. Journal of Magnetic Resonance 209, 195206.Google Scholar
Lu, G. J., Tian, Y., Vora, N., Marassi, F. M. & Opella, S. J. (2013). The structure of the mercury transporter MerF in phospholipid bilayers: a large conformational rearrangement results from N-terminal truncation. Journal of the American Chemical Society 135, 92999302.Google Scholar
Lu, J.-X., Sharpe, S., Ghirlando, R., Yau, W.-M. & Tycko, R. (2010). Oligomerization state and supramolecular structure of the HIV-1 Vpu protein transmembrane segment in phospholipid bilayers. Protein Science 19, 18771896.Google Scholar
Ma, C., Marassi, F., Jones, D., Straus, S., Bour, S., Strebel, K., Schubert, U., Oblatt-Montal, M., Montal, M. & Opella, S. (2002). Expression, purification, and activities of full-length and truncated versions of the integral membrane protein Vpu from HIV-1. Protein Science 11, 546557.Google Scholar
Malojčić, G., Owen, R. L., Grimshaw, J. P. A. & Glockshuber, R. (2008). Preparation and structure of the charge-transfer intermediate of the transmembrane redox catalyst DsbB. FEBS Letters 582, 33013307.Google Scholar
Marassi, F. M., Das, B. B., Lu, G. J., Nothnagel, H. J., Park, S. H., Son, W. S., Tian, Y. & Opella, S. J. (2011). Structure determination of membrane proteins in five easy pieces. Methods 55, 363369.Google Scholar
Marassi, F. M., Ma, C., Gratkowski, H., Straus, S. K., Strebel, K., Oblatt-Montal, M., Montal, M. & Opella, S. J. (1999). Correlation of the structural and functional domains in the membrane protein Vpu from HIV-1. Proceedings of the National Academy Science of the United State of America 96, 1433614341.Google Scholar
Marassi, F. M. & Opella, S. J. (2003). Simultaneous assignment and structure determination of a membrane protein from NMR orientational restraints. Protein Science 12, 403411.Google Scholar
McConnell, H. M. & Hubbell, W. L. (1971). Molecular motion in spin-labeled phospholipids and membranes. Journal of the American Chemical Society 93, 314326.Google Scholar
McConnell, H. M., Wright, K. L. & Mcfarland, B. G. (1972). The fraction of the lipid in a biological membrane that is in a fluid state: a spin label assay. Biochemical and Biophysical Research Communications 47, 273281.Google Scholar
Mcdermott, A. (2009). Structure and dynamics of membrane proteins by magic angle spinning solid-state NMR. Annual Review of Biophyics 38, 385403.Google Scholar
Mcdonnell, P. A., Shon, K., Kim, Y. & Opella, S. J. (1993). fd coat protein structure in membrane environments. Journal of Molecular Biology 233, 447463.Google Scholar
Mclaughlin, A. C., Cullis, P. R., Hemminga, M. A., Hoult, D. I., Radda, G. K., Ritchie, G. A., Seeley, P. J. & Richards, R. E. (1975). Application of 31P NMR to model and biological membrane systems. FEBS Letters 57, 213218.Google Scholar
Mehring, M., Griffin, R. G. & Waugh, J. S. (1971). 19F Shielding tensors from coherently narrowed NMR powder spectra. Journal of Chemical Physics 55, 746755.Google Scholar
Miao, Y., Qin, H., Fu, R., Sharma, M., Can, T. V., Hung, I., Luca, S., GOR'KOV, P. L., Brey, W. W. & Cross, T. A. (2012). M2 proton channel structural validation from full-length protein samples in synthetic bilayers and E. coli membranes. Angewandte Chemie International Edition in English 51, 83838386.Google Scholar
Miller, S. L. (1953). A production of amino acids under possible primitive earth conditions. Science 117, 528529.Google Scholar
Montal, M. & Opella, S. J. (2002). The structure of the M2 channel-lining segment from the nicotinic acetylcholine receptor. Biochimica et Biophysica Acta – Biomembranes 1565, 287293.Google Scholar
Montserret, R., Saint, N., Vanbelle, C., Salvay, A. S. G., Simorre, J.-P., Ebel, C., Sapay, N., Renisio, J.-G., Böckmann, A., Steinmann, E., Pietschmann, T., Dubuisson, J., Chipot, C. & Penin, F. S. (2010). NMR structure and ion channel activity of the p7 protein from hepatitis C virus. Journal of Biological Chemistry 285, 3144631461.Google Scholar
Mueller, L. J. & Dunn, M. F. (2013). NMR crystallography of enzyme active sites: probing chemically-detailed, three-dimensional structure in tryptophan synthase. Accounts of Chemical Research 46, 20082017.Google Scholar
Murray, D. T., Das, N. & Cross, T. A. (2013). Solid state NMR strategy for characterizing native membrane protein structures. Accounts of Chemical Research 46, 21722181.Google Scholar
Nambudripad, R., Stark, W., Opella, S. J. & Makowski, L. (1991). Membrane-mediated assembly of filamentous bacteriophage Pf1 coat protein. Science 252, 13051308.Google Scholar
Nevzorov, A. A. (2008). Mismatched Hartmann–Hahn conditions cause proton-mediated intermolecular magnetization transfer between dilute low-spin nuclei in NMR of static solids. Journal of the American Chemical Society 130, 1128211283.Google Scholar
Nevzorov, A. A. & Opella, S. J. (2003). Structural fitting of PISEMA spectra of aligned proteins. Journal of Magnetic Resonance 160, 3339.Google Scholar
Nevzorov, A. A. & Opella, S. J. (2007). Selective averaging for high-resolution solid-state NMR spectroscopy of aligned samples. Journal of Magnetic Resonance 185, 5970.Google Scholar
Oldfield, E., Chapman, D. & Derbyshire, W. (1971). Deuteron resonance: a novel approach to the study of hydrocarbon chain mobility in membrane systems. FEBS Letters 16, 102104.Google Scholar
Opella, S. (2013a). Structure determination of membrane proteins by nuclear magnetic resonance spectrocopy. Annual Review of Analysis Chemistry 6, 305328.Google Scholar
Opella, S. J. (2013b). Structure determination of membrane proteins in their native phospholipid bilayer environment by rotationally aligned solid-state NMR spectroscopy. Accounts of Chemical Research 46, 21452153.Google Scholar
Opella, S. J., Cross, T. A., Diverdi, J. A. & Sturm, C. F. (1980). Nuclear magnetic resonance of the filamentous bacteriophage fd. Biophysical Journal 32, 531548.Google Scholar
Opella, S. J., Marassi, F. M., Gesell, J. J., Valente, A. P., Kim, Y., Oblatt-Montal, M. & Montal, M. (1999). Structures of the M2 channel-lining segments from nicotinic acetylcholine and NMDA receptors by NMR spectroscopy. Nature Structural Biology 6, 374379.Google Scholar
Opella, S. J. & Waugh, J. S. (1977). Two-dimensional 13C NMR of highly oriented polyethylene. Journal of Chemical Physics 66, 49194924.Google Scholar
Opella, S. J., Zeri, A. C. & Park, S. H. (2008). Structure, dynamics, and assembly of filamentous bacteriophages by nuclear magnetic resonance spectroscopy. Annual Review of Physical Chemistry 59, 635657.Google Scholar
Ouyang, B., Xie, S., Berardi, M. J., Zhao, X., Dev, J., Yu, W., Sun, B. & Chou, J. J. (2013). Unusual architecture of the p7 channel from hepatitis C virus. Nature 498, 521525.Google Scholar
Oxenoid, K. & Chou, J. J. (2005). The structure of phospholamban pentamer reveals a channel-like architecture in membranes. Proceedings of the National Academy Science of the United State of America 102, 1087010875.Google Scholar
Oxenoid, K., Rice, A. J. & Chou, J. J. (2007). Comparing the structure and dynamics of phospholamban pentamer in its unphosphorylated and pseudo-phosphorylated states. Protein Science 16, 19771983.Google Scholar
Page, R. C., Moore, J. D., Nguyen, H. B., Sharma, M., Chase, R., Gao, F. P., Mobley, C. K., Sanders, C. R., Ma, L., Sonnichsen, F. D., Lee, S., Howell, S. C., Opella, S. J. & Cross, T. A. (2006). Comprehensive evaluation of solution nuclear magnetic resonance spectroscopy sample preparation for helical integral membrane proteins. Journal of Structural and Functional Genomics 7, 5164.Google Scholar
Pandey, M. K. & Ramamoorthy, A. (2013). Quantum chemical calculations of amide-15N chemical shift anisotropy tensors for a membrane-bound cytochrome-b5. Journal of Physical Chemistry B 117, 859867.Google Scholar
Park, S. H., Casagrande, F., Cho, L., Albrecht, L. & Opella, S. J. (2011a). Interactions of interleukin-8 with the human chemokine receptor CXCR1 in phospholipid bilayers by NMR spectroscopy. Journal of Molecular Biology 414, 194203.Google Scholar
Park, S. H., Casagrande, F., Das, B. B., Albrecht, L., Chu, M. & Opella, S. J. (2011b). Local and global dynamics of the G protein-coupled receptor CXCR1. Biochemistry 50, 23712380.Google Scholar
Park, S. H., Das, B. B., Casagrande, F., Tian, Y., Nothnagel, H. J., Chu, M., Kiefer, H., Maier, K., De Angelis, A. A., Marassi, F. M. & Opella, S. J. (2012). Structure of the chemokine receptor CXCR1 in phospholipid bilayers. Nature 491, 779783.Google Scholar
Park, S. H., Das, B. B., Deangelis, A. A., Scrima, M. & Opella, S. J. (2010a). Mechanically, magnetically, and ‘rotationally aligned’ membrane proteins in phospholipid bilayers give equivalent angular constraints for NMR structure determination. Journal of Physical Chemistry B 114, 13995–13003.Google Scholar
Park, S. H., De Angelis, A. A., Nevzorov, A. A., Wu, C. H. & Opella, S. J. (2006). Three-dimensional structure of the transmembrane domain of Vpu from HIV-1 in aligned phospholipid bicelles. Biophysical Journal 91, 30323042.Google Scholar
Park, S. H., Marassi, F. M., Black, D. & Opella, S. J. (2010b). Structure and dynamics of the membrane-bound form of Pf1 coat protein: implications of structural rearrangement for virus assembly. Biophysical Journal 99, 14651474.Google Scholar
Park, S. H., Mrse, A. A., Nevzorov, A. A., De Angelis, A. A. & Opella, S. J. (2005). Rotational diffusion of membrane proteins in aligned phospholipid bilayers by solid-state NMR spectroscopy. Journal of Magnetic Resonance 178, 162165.Google Scholar
Park, S. H., Mrse, A. A., Nevzorov, A. A., Mesleh, M. F., Oblatt-Montal, M., Montal, M. & Opella, S. J. (2003). Three-dimensional structure of the channel-forming trans-membrane domain of virus protein “u” (Vpu) from HIV-1. Journal of Molecular Biology 333, 409424.Google Scholar
Park, S. H. & Opella, S. J. (2007). Conformational changes induced by a single amino acid substitution in the trans-membrane domain of Vpu: implications for HIV-1 susceptibility to channel blocking drugs. Protein Science 16, 22052215.Google Scholar
Parthasarathy, S., Nishiyama, Y. & Ishii, Y. (2013). Sensitivity and resolution enhanced solid-state NMR for paramagnetic systems and biomolecules under very fast magic angle spinning. Accounts of Chemical Research 46, 21272135.Google Scholar
Perutz, M. F., Rossman, M. G., Cullis, A. F., Muirhead, H. & Georg, W. (1960). Structure of hemoglobin: a three-dimensional Fourier synthesis at 5·5 A resolution, obtained by X-ray analysis. Nature 185, 416422.Google Scholar
Petsko, G. A. & Ringe, D. (2008). Protein Structure and Function. Oxford University Press: Oxford.Google Scholar
Poget, S. F. & Girvin, M. E. (2007). Solution NMR of membrane proteins in bilayer mimics: small is beautiful, but sometimes bigger is better. Biochimica et Biophysica Acta – Biomembranes 1768, 30983106.Google Scholar
Pollesello, P. & Annila, A. (2002). Structure of the 1–36 N-terminal fragment of human phospholamban phosphorylated at Ser-16 and Thr-17. Biophysical Journal 83, 484490.Google Scholar
Poo, M.-M. & Cone, R. A. (1973). Lateral diffusion of rhodopsin in Necturus rods. Experimental Eye Research 17, 503510.Google Scholar
Poo, M.-M. & Cone, R. A. (1974). Lateral diffusion of rhodopsin in the photoreceptor membrane. Nature 247, 438441.Google Scholar
Popot, J. L. (2010). Amphipols, nanodiscs, and fluorinated surfactants: three nonconventional approaches to studying membrane proteins in aqueous solutions. Annual Review of Biochemistry 79, 737777.Google Scholar
Prosser, R. S., Hunt, S. A., Dinatale, J. A. & Vold, R. R. (1996). Magnetically aligned membrane model systems with positive order parameter: switching the sign of Szz with paramagnetic ions. Journal of the American Chemical Society 118, 269270.Google Scholar
Rajarathnam, K., Clark-Lewis, I. & Sykes, B. D. (1995). 1H NMR solution structure of an active monomeric interleukin-8. Biochemistry 34, 1298312990.Google Scholar
Ramachandran, G. N., Ramakrishnan, C. & Sasisekharan, V. (1963). Stereochemistry of polypeptide chain configurations. Journal of Molecular Biology 7, 9599.Google Scholar
Richardson, J. S. (1981). Anatomy and taxonomy of protein structures. Advances in Protein Chemistry 34, 167339.Google Scholar
Saito, H., Ando, I. & Ramamoorthy, A. (2010). Chemical shift tensor – the heart of NMR: insights into biological aspects of proteins. Progress in NMR Spectroscopy 57, 181228.Google Scholar
Sanders, C. R. & Oxenoid, K. (2000). Customizing model membranes and samples for NMR spectroscopic studies of complex membrane proteins. Biochimica et Biophysica Acta – Biomembranes 1508, 129145.Google Scholar
Schnell, J. & Chou, J. (2008). Structure and mechanism of the M2 proton channel of influenza A virus. Nature 451, 591595.Google Scholar
Schubert, U., Bour, S., Ferrer-Montiel, A. V., Montal, M., Maldarell, F. & Strebel, K. (1996). The two biological activities of human immunodeficiency virus type 1 Vpu protein involve two separable structural domains. Journal of Virology 70, 809819.Google Scholar
Seelig, J., Axel, F. & Limacher, H. (1973). Molecular architecture of bilayer membranes. Annals of the New York Academy Science 222, 588596.Google Scholar
Seiter, C. H. A. & Chan, S. I. (1973). Molecular motion in lipid bilayers: a nuclear magnetic resonance line width study. Journal of the American Chemical Society 95, 75417553.Google Scholar
Sengupta, I., Nadaud, P. S. & Jaroniec, C. P. (2013). Protein structure determination with paramagnetic solid-state NMR spectroscopy. Accounts of Chemical Research 46, 21172126.Google Scholar
Sharma, M., Yi, M., Dong, H., Qin, H., Peterson, E., Busath, D. D., Zhou, H.-X. & Cross, T. A. (2010). Insight into the mechanism of the influenza A proton channel from a structure in a lipid bilayer. Science 330, 509512.Google Scholar
Sharpe, S., Yau, W.-M. & Tycko, R. (2005). Structure and dynamics of the HIV-1 Vpu transmembrane domain revealed by solid-state NMR with magic-angle spinning. Biochemistry 45, 918933.Google Scholar
Sheetz, M. P. & Chan, S. I. (1972). Effect of sonication on the structure of lecithin bilayers. Biochemistry 11, 45734581.Google Scholar
Shi, L., Kawamura, I., Jung, K. H., Brown, L. S. & Ladizhansky, V. (2011). Conformation of a seven-helical transmembrane photosensor in the lipid environment. Angewandte Chemie (International Edition in English) 50, 13021305.Google Scholar
Shi, L. & Ladizhansky, V. (2012). Magic angle spinning solid-state NMR experiments for structural characterization of proteins. Methods in Molecular Biology 895, 153165.Google Scholar
Shon, K. J., Kim, Y., Colnago, L. A. & Opella, S. J. (1991). NMR studies of the structure and dynamics of membrane-bound bacteriophage Pf1 coat protein. Science 252, 13031305.Google Scholar
Singer, S. J. & Nicholson, G. L. (1972). The fluid mosaic model of the structure of cell membranes. Science 175, 720731.Google Scholar
Skasko, M., Wang, Y., Tian, Y., Tokarev, A., Munguia, J., Ruiz, A., Stephens, E. B., Opella, S. J. & Guatelli, J. (2012). HIV-1 Vpu protein antagonizes innate restriction factor BST-2 via lipid-embedded helix-helix interactions. Journal of Biological Chemistry 287, 5867.Google Scholar
Skelton, N. J., Quan, C., Reilly, D. & Lowman, H. (1999). Structure of a CXC chemokine-receptor fragment in complex with interleukin-8. Structure 7, 157168.Google Scholar
Smith, R. L. & Oldfield, E. (1984). Dynamic structure of membranes by deuterium NMR. Science 225, 280288.Google Scholar
Soong, R., Smith, P. E., Xu, J., Yamamoto, K., Im, S. C., Waskell, L. & Ramamoorthy, A. (2010). Proton-evolved local-field solid-state NMR studies of cytochrome b5 embedded in bicelles, revealing both structural and dynamical information. Journal of the American Chemical Society 132, 57795788.Google Scholar
Spudich, J. L. & Luecke, H. (2002). Sensory rhodopsin II: functional insights from structure. Current Opinion in Structural Biology 12, 540546.Google Scholar
Stouffer, A. L., Acharya, R., Salom, D., Levine, A. S., Di Costanzo, L., Soto, C. S., Tereshko, V., Nanda, V., Stayrook, S. & Degrado, W. F. (2008). Structural basis for the function and inhibition of an influenza virus proton channel. Nature 451, 596599.Google Scholar
Tanford, C. (1978). The hydrophobic effect and the organization of living matter. Science 200, 10121018.Google Scholar
Tang, M., Comellas, G. & Rienstra, C. M. (2013a). Advanced solid-state NMR approaches for structure determination of membrane proteins and amyloid fibrils. Accounts of Chemical Research 46, 20802088.Google Scholar
Tang, M., Nesbitt, A. E., Sperling, L. J., Berthold, D. A., Schwieters, C. D., Gennis, R. B. & Rienstra, C. M. (2013b). Structure of the disulfide bond generating membrane protein DsbB in the lipid bilayer. Journal of Molecular Biology 425, 16701682.Google Scholar
Tang, M., Sperling, L. J., Berthold, D. A., Nesbitt, A. E., Gennis, R. B. & Rienstra, C. M. (2011b). Solid-state NMR study of the charge-transfer complex between ubiquinone-8 and disulfide bond generating membrane protein DsbB. Journal of the American Chemical Society 133, 43594366.Google Scholar
Tang, M., Sperling, L. , Berthold, D., Schwieters, C., Nesbitt, A., Nieuwkoop, A., Gennis, R. & Rienstra, C. (2011a). High-resolution membrane protein structure by joint calculations with solid-state NMR and X-ray experimental data. Journal of Biomolecular NMR 51, 227233.Google Scholar
Thiriot, D. S., Nevzorov, A. A. & Opella, S. J. (2005). Structural basis of the temperature transition of Pf1 bacteriophage. Protein Science 14, 10641070.Google Scholar
Thiriot, D. S., Nevzorov, A. A., Zagyanskiy, L., Wu, C. H. & Opella, S. J. (2004). Structure of the coat protein in Pf1 bacteriophage determined by solid-state NMR spectroscopy. Journal of Molecular Biology 341, 869879.Google Scholar
Tian, Y., Schwieters, C. D., Opella, S. J. & Marassi, F. M. (2012). AssignFit: a program for simultaneous assignment and structure refinement from solid-state NMR spectra. Journal of Magnetic Resonance 214, 4250.Google Scholar
Traaseth, N. J., Shi, L., Verardi, R., Mullen, D. G., Barany, G. & Veglia, G. (2009). Structure and topology of monomeric phospholamban in lipid membranes determined by a hybrid solution and solid-state NMR approach. Proceedings of the National Academy of Sciences of the United State of America 106, 1016510170.Google Scholar
Traaseth, N. J. & Veglia, G. (2010). Probing excited states and activation energy for the integral membrane protein phospholamban by NMR CPMG relaxation dispersion experiments. Biochimica et Biophysica Acta – Biomembranes 1798, 7781.Google Scholar
Ullrich, S. J. & Glaubitz, C. (2013). Perspectives in enzymology of membrane proteins by solid-state NMR. Accounts of Chemical Research 46, 21642171.Google Scholar
Ulrich, A. S. (2005). Solid state 19G NMR methods for studying biomembranes. Progress in Nuclear Magnetic Resonance Spectroscopy 46, 121.Google Scholar
Unwin, N. (1995). Acetylcholine receptor channel imaged in the open state. Nature 373, 3743.Google Scholar
Van Horn, W. D., Kim, H.-J., Ellis, C. D., Hadziselimovic, A., Sulistijo, E. S., Karra, M. D., Tian, C., Sönnichsen, F. D. & Sanders, C. R. (2009). Solution nuclear magnetic resonance structure of membrane-integral diacylglycerol kinase. Science 324, 17261729.Google Scholar
Veksli, Z., Salsbury, N. J. & Chapman, D. (1969). Physical studies of phospholipids. XII. Nuclear magnetic resonance studies of molecular motion in some pure lecithin-water systems. Biochimica et Biophysica Acta – Biomembranes  183, 434446.Google Scholar
Verardi, R., Shi, L., Traaseth, N. J., Walsh, N. & Veglia, G. (2011). Structural topology of phospholamban pentamer in lipid bilayers by a hybrid solution and solid-state NMR method. Proceedings of the National Academy of Sciences of the United State of America 108, 91019106.Google Scholar
Vinogradova, O., Sonnichsen, F. & Sanders, C. R. II (1998). On choosing a detergent for solution NMR studies of membrane proteins. Journal of Biomolecuiar NMR 11, 381386.Google Scholar
Vostrikov, V. V., Mote, K. R., Verardi, R. & Veglia, G. (2013). Structural dynamics and topology of phosphorylated phospholamban homopentamer reveal its role in the regulation of calcium transport. Structure 21, 21192130.Google Scholar
Wacey, D., Kilbrun, M. R., Saunders, M., Cliff, J. & Brasier, M. D. (2011). Microfossils of sulphur-metabolizing cells in 3·4-billion-year-old rocks of Western Australia. Nature Geoscience 4, 698702.Google Scholar
Wang, J., Kim, S., Kovacs, F. & Cross, T. A. (2001). Structure of the transmembrane region of the M2 protein H+ channel. Protein Science 10, 22412250.Google Scholar
Wang, S., Munro, R. A., Shi, L., Kawamura, I., Okitsu, T., Wada, A., Kim, S. Y., Jung, K. H., Brown, L. S. & Ladizhansky, V. (2013a). Solid-state NMR spectroscopy structure determination of a lipid-embedded heptahelical membrane protein. Natural Methods 10, 10071012.Google Scholar
Wang, Y., Park, S. H., Tian, Y. & Opella, S. J. (2013b). Impact of histidine residues on the transmembrane helices of viroporins. Molecular Membrane Biology 30, 360369.Google Scholar
Watson, J. D. & Crick, F. H. C. (1953). Molecular structure of nucleic acids. Nature 171, 737738.Google Scholar
Weingarth, M. & Baldus, M. (2013). Solid-state NMR-based approaches for supramolecular structure elucidation. Accounts of Chemical Research 46, 20372046.Google Scholar
Wien, R. W., Morrisett, J. D. & Mcconnelll, H. M. (1972). Spin-label-induced nuclear relaxation. Distances between bound saccharides, histidines-15, and tryptophan-123 on lysozyme in solution. Biochemistry 11, 37073716.Google Scholar
Willbold, D., Hoffmann, S. & Rösch, P. (1997). Secondary structure and tertiary fold of the human immunodeficiency virus protein U (Vpu) cytoplasmic domain in solution. European Journal of Biochemistry 245, 581588.Google Scholar
Wittlich, M., Koenig, B. W., Stoldt, M., Schmidt, H. & Willbold, D. (2009). NMR structural characterization of HIV-1 virus protein U cytoplasmic domain in the presence of dodecylphosphatidylcholine micelles. FEBS Journal 276, 65606575.Google Scholar
Wuthrich, K. & Wagner, G. (1975). NMR investigations of the dynamics of the aromatic amino acid residues in the basic pancreatic trypsin inhibitor. FEBS Letter 50, 265268.Google Scholar
Wylie, B. J., Franks, W. T. & Rienstra, C. M. (2006). Determinations of 15N chemical shift anisotropy magnitudes in a uniformly 15N,13C-labeled microcrystalline protein by three-dimensional magic-angle spinning nuclear magnetic resonance spectroscopy. Journal of Physical Chemistry B 110, 1092610936.Google Scholar
Wylie, B. J., Sperling, L., Frericks, H., Shah, G., Franks, W. & Rienstra, C. (2007). Chemical-shift anisotropy measurements of amide and carbonyl resonances in a microcrystalline protein with slow magic-angle spinning NMR spectroscopy. Journal of the American Chemical Society 129, 53185319.Google Scholar
Xu, J., Soong, R., Im, S.-C., Waskell, L. & Ramamoorthy, A. (2010). INEPT-based separated local field NMR spectrsocopy: a unique approach to elucidate side-chain dynamics of membrane-associated proteins. Journal of the American Chemical Society 132, 99449947.Google Scholar
Yamamoto, K., Durr, H. N. U., Xu, J., Im, S.-C., Waskell, L. & Ramamoorthy, A. (2013a). Dynamic interaction between membrane bound full-length cytochrome P450 and cytochrome b5 observed by solid-state NMR spectroscopy. Science Report 3, 15.Google Scholar
Yamamoto, K., Gildenberg, M., Ahuja, S., Im, S. C., Pearcy, P., Waskell, L. & Ramamoorthy, A. (2013b). Probing the transmembrane structure and topology of microsomal cytochrome-p450 by solid-state NMR on temperature-resistant bicelles. Science Report 3, 2556.Google Scholar
Yan, S., Suiter, C. L., Hou, G., Zhang, H. & Polenova, T. (2013). Probing structure and dynamics of protein assemblies by magic angle spinning NMR spectroscopy. Accounts of Chemical Research 46, 20472058.Google Scholar
Zamoon, J., Mascioni, A., Thomas, D. D. & Veglia, G. (2003). NMR solution structure and topological orientation of monomeric phospholamban in dodecylphosphocholine micelles. Biophysical Journal 85, 25892598.Google Scholar
Zeri, A. C., Mesleh, M. F., Nevzorov, A. A. & Opella, S. J. (2003). Structure of the coat protein in fd filamentous bacteriophage particles determined by solid-state NMR spectroscopy. Proceedings of the National Academy Science of the United State of America 100, 64586463.Google Scholar
Zheng, J. & Jia, Z. (2013). Structural biology: tiny enzyme uses context to succeed. Nature 497, 445446.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
Zhou, Y., Cierpicki, T., Jimenez, R. H. F., Lukasik, S. M., Ellena, J. F., Cafiso, D. S., Kadokura, H., Beckwith, J. & Bushweller, J. H. (2008). NMR solution structure of the integral membrane enzyme DsbB: functional insights into DsbB-catalyzed disulfide bond formation. Molecular Cell 31, 896908.Google Scholar