Abstract
Membrane proteins play many critical roles in cells, mediating flow of material and information across cell membranes. They have evolved to perform these functions in the environment of a cell membrane, whose physicochemical properties are often different from those of common cell membrane mimetics used for structure determination. As a result, membrane proteins are difficult to study by traditional methods of structural biology, and they are significantly underrepresented in the protein structure databank. Solid-state Nuclear Magnetic Resonance (SSNMR) has long been considered as an attractive alternative because it allows for studies of membrane proteins in both native-like membranes composed of synthetic lipids and in cell membranes. Over the past decade, SSNMR has been rapidly developing into a major structural method, and a growing number of membrane protein structures obtained by this technique highlights its potential. Here we discuss membrane protein sample requirements, review recent progress in SSNMR methodologies, and describe recent advances in characterizing membrane proteins in the environment of a cellular membrane.
Keywords: membrane protein, protein structure, solid-state NMR, lipid bilayer, cell membrane
Introduction
Membrane proteins play many important roles in biology of the cell, from regulation of the membrane transport to signal transduction, both between the extracellular and intracellular media and between different intracellular compartments. Although it is generally accepted that their structure and function are sensitive to the environment, the majority of high-resolution structural methods require membrane proteins to be prepared in the form either completely devoid of, or containing a small amount of lipids. Amongst high-resolution structure determination methods, X-ray crystallography has been the most prolific to date, providing the most detailed view of molecular structure at the same time. Since the first crystallographic structure of myoglobin,1 ∼90,000 protein structures have been determined by this method out of a total of more than 100,000 structures, and impressive progress has been achieved in crystallography of membrane proteins.2–5 Solution NMR approaches have also continued to mature,6–9 and provided a few structures of polytopic α-helical membrane proteins in detergent micelles.10–14 Despite this impressive progress in membrane protein structure determination, sample preparation remains the main challenge for both X-ray crystallography and solution NMR, and the number of membrane protein structures in the protein data bank is less than 2% of the total number of structures.
Membrane proteins have evolved to perform their functions in the environment of a cell membrane. Being able to study them in the native-like environment of the lipid bilayer ensures that the sample represents a biologically relevant protein conformation, that is, as close as possible to the one present in living cells. Solid-state NMR (SSNMR) has a long history of applications to the studies of structure and dynamics of membrane proteins in phospholipids. While earlier studies focused primarily on samples with selectively isotopically labeled sites, or small ligands and cofactors,15–18 the emergence of high-field magnets, new probe technologies,19,20 methods for sample preparation,21 and new SSNMR methodologies over the past decade have greatly expanded the range of membrane proteins amenable to structural investigations, and allowed detailed studies of the samples with uniform incorporation of isotopic 13C and 15N labels. Similar to solution NMR, multiple individual sites can now be resolved and uniquely assigned in the solid-state, paving the way for in-depth structural, dynamic and functional analyses. The potential of SSNMR has initially been demonstrated on soluble model proteins prepared in a microcrystalline state, such as Bovine Pancreatic Trypsin Inhibitor (BPTI),22 α-spectrin SH3 domain,23,24 ubiquitin,25,26 the β1 immunoglobulin binding domain of protein G (GB1),27 and thioredoxin,28 and recently has been expanded to amyloid fibril forming proteins,29–32 which are not readily accessible to X-ray crystallography and solution NMR. SSNMR has also been successfully used to determine structures of several membrane proteins including influenza A M2 proton channel33 and its complex with the drug amantadine,34 phospholamban,35 mycobacterial cell division regulatory protein CrgA,36 transmembrane domain of Yersinia enterocolitica adhesin A (YadA),37 light-sensitive receptor Anabaena Sensory Rhodopsin (ASR)38 and G-protein coupled chemokine CXCR1 receptor.39 In addition, X-ray structures of disulfide bond formation protein B (DsbB),40 and its complex with DsbA41 have been refined to higher resolution by combining electron densities obtained from X-ray diffraction and solid-state NMR restraints. SSNMR offers even more exciting possibility of studying membrane protein structure in situ, either in cell membranes or even in whole cells.42–48 The main focus of this review is on the recent progress in solid-state NMR membrane protein structure determination, primarily concentrating on the magic angle spinning solid-state NMR, as well as on the prospects of studying membrane proteins in cell membrane environment.
Protein Expression and Isotope Labeling
Preparation of samples of stable functional membrane proteins in sufficient quantities is the main challenge for all structural methods. For solid-state NMR studies, one must recombinantly produce a few milligrams of isotopically 13C,15N-enriched (and, sometimes, deuterated) protein. The most frequently used are E. coli based heterologous expression systems, as they offer flexibility in the choice of constructs and labeling strategies, and are relatively inexpensive. Much effort has been directed at the development of the bacterial expression, refolding and functional purification protocols for the production of eukaryotic membrane proteins for NMR, most notably G-protein coupled receptors, including neuropeptide Y2 receptor,49 human cannabinoid Type 2 receptor,50 and human chemokine CXCR1 receptor.51
The use of lower eukaryotes such as Pichia pastoris is another attractive route for the production of isotopically labeled eukaryotic proteins. The utility of this approach for membrane proteins has been demonstrated using eukaryotic (fungal) rhodopsin from Leptosphaeria maculans,52 and human water channel aquaporin 1.53 Both proteins gave solid-state NMR spectra of excellent quality, suitable for high-resolution structural and dynamic characterization. On the other hand, Pichia uses methanol as the main carbon source, and offers a limited flexibility in the choice of 13C-labeling schemes, e.g., not allowing sparse 13C labeling (e.g., by using 2-13C or 1,3-13C labeled glycerol, 1-13C, 2-13C labeled glucose, etc., as carbon sources),54–57 which is easily achieved in bacterial E.coli based systems. Production of sparsely labeled samples is essential part of SSNMR methodology, as it is widely used in SSNMR to both improve spectral resolution for spectroscopic assignments,58,59 and to facilitate distance measurements for structure determination.24 Other species of yeast, such as Kluyveromyces lactis, which can use sugars as carbon source, are being explored to correct this deficiency.60,61
More advanced eukaryotic expression systems, for example, transfected mammalian cells and baculovirus-infected insect cells, have their own challenges such as relatively low expression yields, high cost of production, incomplete labeling, as well as lack of reliable deuteration and uniform labeling protocols, even though these parameters keep improving.62,63 Nevertheless, there are a number of successful SSNMR studies of selectively labeled GPCRs (most notably, visual rhodopsin) produced in these systems.64,65 New eukaryotic expression systems are being developed. For example, a recent study demonstrated successful isotope labeling of bovine visual rhodopsin in a roundworm Caenorhabditis elegans.66 One should also mention significant progress in the cell-free expression methods, which have been used to produce isotopically labeled membrane proteins14,67–69 with several applications in solid-state NMR.70,71
Solid-State NMR Approaches and Sample Formulation
The term “solid-state NMR” implies application of NMR to molecular systems that lack significant motion. In contrast to solution NMR, which requires rapid tumbling of molecules, and is therefore limited by size of the molecule, SSNMR can be applied to protein and/or protein complexes of arbitrary molecular weight. Because of lack of fast molecular tumbling SSNMR spectra are broadened by anisotropic dipolar interactions and chemical shift anisotropies, which need to be removed for site-specific resolution. There are two principal approaches to accomplish this, either by uniformly orienting molecules relative to a magnetic field (oriented sample NMR, OS-NMR), or by applying magic angle spinning (MAS) technique (Fig. 1). The OS-NMR approach relies on the uniaxial alignment of the sample, and the strong angular dependence of the 15N chemical shift anisotropy and 15N-1H dipolar couplings on the molecular orientation with respect to the magnetic field is used to determine the orientation of individual sites relative to the common axis. Most commonly, amino acid selectively 15N-labeled proteins, either incorporated into the bilayers mechanically aligned on glass plates or reconstituted in bicelles, are used to acquire NMR spectra correlating 15N-1H dipolar couplings with 15N chemical shift. In addition, magnetically aligned bicelles72,73 provide high quality of alignment, and they have been used to study membrane proteins, including CXCR1,74 OmpX,75 and cytochrome b5.76
Figure 1.

Solid-state NMR approaches to membrane protein structure determination. (A) Protein molecules can be uniformly oriented with respect to the magnetic field Bo NMR spectra correlating the 1H-15N dipolar coupling to the 15N chemical shift show a circular pattern, which reports on the topology of the protein andits backbone structure. An example on the left is a spectrum of the transmembrane domain of virus protein “u” comprising residues 1–32.80 Reprinted from Journal of Molecular Biology, 333, S.H. Park, A.A. Mrse, A.A. Nevzorov M. F. Mesleh, M. Oblatt-Montal, M. Montal and S.J. Opella, “Three-dimensional Structure of the Channel-forming Trans-membrane Domain of Virus Protein “u” (Vpu) from HIV-1” (2003) 409–424. Copyright 2003, with permission from Elsevier. (B) Magic angle spinning (MAS) NMR is applied to unoriented samples. MAS averages out anisotropic interactions to their isotropic values, and results in spectra that resemble those from samples in solution phase. An example on the right is a 2D NCA spectrum of Anabaena Sensory Rhodopsin correlating isotropic chemical shifts of 15N and 13Cα atoms.81 Reproduced with permission from Angewandte Chemie Int Ed., ‘‘Conformation of a seven-helical transmembrane photosensor n the lipid environment’’ 50 (2011) 1302–1305. Copyright 2011 John Wiley & Sons, Inc. Additionally, in the case when a protein undergoes fast (<10−6 s) rotational diffusion, correlating the 1H-15N dipolar couplings to the 15N chemical shift would result in circular patterns similar to that shown on the left (Rotational Alignment), and can be used as an independent way to obtain backbone structure, or combined with distance measurements under MAS conditions.
In oriented samples, the topology and structure of a membrane protein can be determined from the circular spectral patterns, called PISA wheels, or polarity index slant angle (Fig. 1),77,78 which are directly related to the periodicity of an α-helix and to its orientation with respect to the external field (and, consequently, to the membrane plane). The majority of solid-state NMR structures of membrane proteins available so far have been determined using this approach (http://www.drorlist.com/nmr/MPNMR.html). While the initial applications of OS-NMR have been mostly limited to small proteins and peptides because of the spectral complexity and low sensitivity of heteronuclear 15N detection, very promising SSNMR data for larger proteins have been published in recent years, including well-resolved spectra of the multidrug resistance transporter EmrE79 and a structure of cell division regulatory protein CrgA.36
Magic Angle Spinning (MAS) refers to a technique in which unoriented samples are spun around an axis oriented at 54.7° with respect to the external magnetic field (Fig. 1). MAS coherently averages anisotropic dipolar couplings and chemical shift anisotropies to their isotropic values. A typical carbon or nitrogen MAS spectrum of extensively isotopically labeled membrane protein contains hundreds to thousands of lines, and the spectral widths of individual lines become the most important parameter that essentially determines whether a protein is amenable to high-resolution structural analysis. The spectral resolution and the line widths depend on sample homogeneity, which in turn is strongly affected by the sample preparation. Even slight structural heterogeneity may cause large inhomogeneous line broadening, and render high-resolution structure determination impossible.
One of the major advantages of MAS solid-state NMR is that it allows examining many sample forms ranging from membrane proteins in a precipitated and/or microcrystalline state, to proteins reconstituted in model membranes, or even in more complex and chemically heterogeneous cellular membranes. In Figure 2 we show a few selected examples of typical two-dimensional carbon-carbon correlation spectra of membrane proteins prepared under different conditions. They all have comparably high resolution with tens of cross peaks resolved, indicating high level of structural homogeneity, while some difference in the appearance can be attributed to different methods employed to obtain spectral correlations.
Figure 2.

Magic angle spinning carbon-carbon correlation spectra of various membrane proteins. (A) 900 MHz spectrum of trimeric YadA.82 Reprinted by permission from Macmillan Publishers Ltd: Scientific Reports, 2, S.A. Shahid, S. Markovic, D. Linke, B.J. van Rossum, “Assignment and secondary structure of the YadA membrane protein by solid-state MAS NMR”, (2012) 803. Copyright 2012. (B) 800 MHz spectrum of chimeric potassium channel KcsA-Kv1.3.85 Adapted with permission from J. Am. Chem. Soc., 130, R. Schneider, C. Ader, A. Lange, K. Giller, S. Hornig, O. Pongs, S. Becker, M. Baldus, “Solid-state NMR spectroscopy applied to a chimeric potassium channel in lipid bilayers” (2008) 7427–7435. Copyright 2008 American Chemical Society. (C) 750 MHz spectrum of 40DsbB. Reprinted from Journal of Molecular Biology, 425, M. Tang, A.E. Nesbitt, L.J. Sperling, D.A. Berthold, C.D. Schwieters, R.B. Gennis, C.M. Rienstra, “Structure of the disulfide bond generating membrane protein DsbB in the lipid bilayer” (2013) 1670–1682. Copyright 2013, with permission from Elsevier. (D) 800 MHz spectrum of proteorhodopsin (unpublished data). (E) 900 MHz spectrum of VDAC1 sample prepared as 2D crystals.86 Adapted with permission from J. Am. Chem. Soc., 134, M.T. Eddy, T.C. Ong, L. Clark, O. Tejido, P.C.van der Wel, R. Garces, G. Wagner, T.K. Rostovtseva, R.G. Griffin, “Lipid dynamics and protein-lipid interactions in 2D crystals formed with the ß-barrel integral membrane protein VDAC1” Copyright 2012 American Chemical Society. (F) 800 MHz spectrum of water channel human aquaporin 1.53 Reprinted from Journal of Biomolecular NMR, S. Emami, Y. Fan, R. Munro, V. Ladizhansky, L.S. Brown, “Yeast-expressed human membrane protein aquaporin-1 yields excellent resolution of solid-state MAS NMR spectra” 55, (2013) 147155. Reproduced with kind permission from Springer Science and Business Media.
An important question in determining the sample preparation strategy is to what degree the resolution of SSNMR spectra is determined by the local order. For example, samples of YadA82 [Fig. 2(A)] and DsbB83 [Fig. 2(B)] proteins were prepared in precipitated/microcrystalline states. While small, poorly diffracting microcrystals are considerably easier to prepare than the high-quality crystals suitable for X-ray crystallographic studies, the resolution of NMR spectra is comparable between the two, as was demonstrated on polycrystalline and nanocrystalline ubiquitin.84 Dense packing achieved in crystals maximizes the amount of protein that can be packed in the active volume of the MAS NMR rotor, and offers a significant sensitivity advantage. Although proteins in small crystals are likely to be prone to the artifacts resulting from crystal contacts, restricted dynamics, etc., same as in large crystals used for protein crystallography, a variety of biochemical parameters can be varied with relative ease, allowing for the studies of protein function. Both structural and functional information can thus be deduced from such samples even in cases when the production of large crystals for X-ray analysis is problematic.
The environment of the lipid bilayer allows for a variety of native-like biochemical and biophysical conditions to be applied, including protein-to-lipid ratio, pH, temperature, bilayer fluidity and hydrophobic thickness, and is often the preparation of choice for SSNMR studies. Early publications argued that 2D crystalline preparation promotes local conformational homogeneity,87 and indeed a few proteins prepared in the form of 2D crystals yielded spectra of excellent quality, such as those of BR,88 OmpG,87 ASR,81 VDAC186 [Fig. 2(E)], and human aquaporin 153 [Fig. 2(F)]. It is important however for the 2D crystals to be macroscopically homogeneous. For example, 7TM light-driven proton pump proteorhodopsin (PR) forms 2D crystals of different symmetries under various lipid reconstitution conditions.89–91 In particular, PR embedded in the DOPC lipid matrix at high protein-to-lipid ratio (w/w 4:1) forms 2D assemblies of crystalline domains of hexagonal (and occasionally pentagonal) symmetry, which coexist with non-crystalline densely packed domains.91 While SSNMR samples prepared under these conditions yielded spectra of acceptable quality,92 even better resolution was obtained in non-crystalline PR samples produced under alternative reconstitution protocol, which used the DMPC/DMPA lipid mixture (9:1 w/w), and a lower protein-to-lipid ratio (2:1 w/w) [Fig. 2(D)].93 Thus, homogeneity of 2D crystal formation, as well as the chemistry of lipids, protein local mobility, and uniform state of oligomerization are also important factors for spectral resolution.
Overall, MAS SSNMR has yielded well-resolved spectra for a large number of membrane proteins of different classes, secondary structure and architecture. While samples need to be optimized individually for each protein, the examples given above demonstrate that the preparation of structurally homogeneous samples does not appear to be a limitation for the structural characterization of membrane proteins by solid-state NMR, at least in the cases where intrinsic conformational heterogeneity is not present or can be minimized.
While traditional applications of MAS SSNMR to membrane proteins mainly relied on carbon detection in 15N, 13C-labeled samples, there is a growing number of studies where proton detection in back-H/D-exchanged perdeuterated samples is employed.94–97 Proton detection is attractive because it provides both the improvement in resolution, and higher sensitivity due to the large gyromagnetic ratio of protons (four times larger than that of carbons). The utility of this approach has principally been demonstrated in the initial applications to α-spectrin SH3 domain, where excellent spectral resolution rivaling that of solution NMR was obtained.98,99 Although excellent resolution was also achieved in the first application to a membrane protein bacteriorhodopsin, only about a third of the expected resonances could be observed because of the incomplete back-exchange.94 Similarly, only solvent-accessible sites could be observed in another 7TM membrane protein proteorhodopsin.95 Thus, application of proton detection in membrane proteins would require the use of unfolding-refolding protocols, or alternative methods for the introduction of the detectable protons in the perdeuterated protein.100,101 Overall, the use of proton detection is an attractive route and will continue to be the area of active development. Not only it gives access to proton chemical shifts, which are sensitive probes of protein structure and interactions, but it also permits using approximately an order of magnitude smaller amounts of sample, which is important for hard-to-express proteins.
Structure determination based on distance restraints
MAS solid-state NMR structure determination strategies follow in the footsteps of solution NMR spectroscopy. The majority of studies of membrane proteins to date used carbon detection, and site-specific assignments can be obtained using established suites of 13C-detected 3D or 4D chemical shift correlation experiments83,102,103 assisted by simplifying isotope labeling schemes.58,59,104 The analysis of experimental data and extraction of chemical shifts is assisted by the available software tools, which reduce the possibility of error.105–107 Chemical shifts of individual residues can be used to determine the backbone structure using either the established correlations between chemical shifts and secondary structure,108 or an empirical relation between the chemical shifts and backbone dihedral angles.109 Additionally, torsion angles defining relative orientation of chemical bonds can be determined from correlation experiments, which relate relative orientations of dipolar and/or chemical shift anisotropy interactions, which are fixed with respect to the protein backbone.110,111
Structure determination in the solid state is based on determining a large number of structurally constraining internuclear distances (typically up to ∼6 Å between carbon atoms, or between carbon and nitrogen atoms), which are normally obtained from the dipolar recoupling experiments combined with multidimensional chemical shift correlation experiments for optimal spectral resolution.24,112,113 Such approach turned out to be useful even for proteins of relatively large size, such as 7TM alpha-helical microbial rhodopsin ASR. Long-range distance restraints (approximately up to 20 Å) can also been obtained in the form of Paramagnetic Relaxation Enhancements (PRE)114,115 from samples with paramagnetic tags attached to cysteines introduced through site-directed mutagenesis. Owing to their longer-range order, PREs are ideally suited for the characterization of protein−protein interactions, and have been used to determine the oligomerization interface of ASR, resulting in a refined SSNMR structure of its 81 kDa trimer.116
Structure determination using rotational alignment
As pointed out above, the term “solid-state NMR” generally applies to the systems that lack rapid overall tumbling. There are, however, other types of global motions that may have a profound effect on the NMR spectra. In particular, it has been observed in the pioneering experiment in BR that rapid submicrosecond rotational diffusion in the bilayer results in partial averaging of dipolar couplings and chemical shift anisotropies, and produces NMR line shapes that are dependent on the orientation of an internuclear vector with respect to the magnetic field axis.117 Based on this observation, Opella and co-workers have proposed a method for membrane protein structure determination termed Rotational Alignment (RA), which relies on MAS118 performed on samples of membrane proteins that undergo rotational diffusion. While OS-NMR is primarily restricted to low sensitivity nitrogen detection, RA MAS NMR can be combined with higher sensitivity of carbon detection, and can be further combined with more general (e.g., not requiring rapid rotational diffusion) MAS approaches discussed above.
RA relies on fast rotational diffusion of a protein about the bilayer normal, and this imposes some restrictions on the sample condition. Any of the factors that limit the rate of diffusion, for example, large molecular weight and/or oligomerization state, interactions with other proteins, fluidity of the lipid bilayer, and so forth, will render the RA approach inapplicable. In the aforementioned study of BR,117 the protein had to be reconstituted in DMPC lipids in the fluid phase, where it exists as a monomer,119 while the rotational diffusion of BR in the purple membrane was much slower, likely limited by interprotein contacts in a trimer, and between trimers within the hexagonal lattice. Still, many important classes of membrane proteins fall into the regime of fast rotational diffusion, as was demonstrated on the monomeric G-protein coupled chemokine CXCR1 receptor, whose structure was determined using the RA technique.39
Solid-state NMR of membrane proteins in cellular membranes
While artificial phospholipid bilayer represents a good initial approximation of a realistic cell membrane, true cell membranes are much more complex and contain many more components, e.g., other integral and peripheral membrane proteins, various lipids, carbohydrates, etc., which can transiently or stably interact with a protein of interest and affect its structure. Solid-state NMR offers a unique capability to examine membrane protein structure site-specifically in the environment of a cell membrane. Principal limitations of SSNMR studies of proteins in cellular membranes are as follows: (i) limited content of the protein of interest and reduced relative and absolute amplitude of its NMR signals; (ii) abundance of other proteins in the membrane results in high spectral overlap; (iii) highly heterogeneous environment may result in structural heterogeneity and inhomogeneous line broadening of the spectra, further reducing both amplitude and resolution.
For a long time, bacteriorhodopsin had been the only protein subjected to detailed structural investigations in its native membrane.120–123 BR, however, represents a special case because it is localized in the 2D crystalline purple membrane (PM) and is highly abundant, constituting ∼75% of the PM content by weight. Several recent studies have considerably expanded the list of peptides and proteins studied in cell membranes, and demonstrated the feasibility of obtaining structural data in cell membranes or even in whole cells.42–48 The first case of in situ SSNMR was presented by Tian et al. who have studied a TM peptide LR11 in E. coli.42 LR11 interacts with Amyloid Precursor Protein and regulates its localization in cells. LR11 was expressed as MBP fusion construct, cleaved in the E. coli membrane, and constituted ∼80% of the total labeled protein in the sample. This greatly facilitated SSNMR assignments of ∼50% of residues of the protein and allowed characterization of its secondary structure.
Low concentration of protein of interest in the “in-membrane” samples currently precludes the de novo structure determination. NMR can however be used to detect chemical shift perturbations and use them for comparison with the respective structures obtained in the liposome environment. Such an approach was recently used to validate the structure of the influenza A M2 proton channel which had been previously obtained in proteoliposomes.33 From the direct comparison of chemical shifts of the conducting part of the channel it was concluded that the channel conformation in the E. coli membrane is very similar to that in synthetic liposomes.45
The presence of background proteins and the associated spectral crowdedness constitutes another significant complication for in-cell and in-membrane SSNMR. To reduce the background signal, Baldus et al. have implemented a new sample preparation strategy, which enabled a study of a much larger 150-residue outer membrane enzyme PagL, which is responsible for the removal of fatty acid chains from lipopolysaccharides in Pseudomonas aeruginosa.47 PagL was expressed in the E. coli strain lacking genes for the two most abundant proteins of the outer membrane, OmpA and OmpF. They further used a labeling protocol, in which the target protein was preferentially labeled, thus reducing the overall background signal in the NMR spectra (Fig. 3).
Figure 3.

Sample preparation strategy for cellular NMR of PagL employed in Ref. 47. A: Schematic representation of the E. coli K-12 cell envelope. B: Preparation of whole cell, cell envelope, and proteoliposome samples. In the initial step, cells are first grown on the unlabeled medium, and then transferred to isotopically labeled medium for induction. Adapted with permission from Ref. 47, The National Academy of Sciences of the USA, 2012.
These improvements have allowed them to simplify spectra and to enhance the relative signal from PagL protein to a degree sufficient to transfer spectroscopic assignments obtained for PagL in proteoliposomes. Of the 13 residues that could be observed and reassigned in PagL in the environments of both cell envelopes and in whole cells, the majority belonged to the extramembranous water-exposed fragments, which were found to be in conformations similar to those in the proteoliposome environment (Fig. 4).
Figure 4.

13C-13C correlation spectra of whole cell (A), cell envelope (B) and proteoliposome (C) samples containing PagL. Encircled regions demonstrate characteristic alanines, serines and threonines of PagL. Adapted with permission from Ref. (47), The National Academy of Sciences of the USA, 2012.
One of the important observations in the aforementioned study was surprisingly high resolution of PagL spectra, despite the apparent heterogeneity of the environment. However, because of the presence of the background E. coli proteins and associated spectral complexity only a handful of resonances belonging to PagL could be identified.
In our investigations of ASR in E. coli membrane, we have implemented additional purification steps commonly used in membrane biochemistry, aimed at increasing the relative content of ASR (Fig. 5). As ASR preferentially partitions into the inner membrane, we applied sucrose gradient to isolate the ASR-containing inner membrane from the outer membrane,124 and to increase the relative content of the protein in the NMR sample. Further sample simplification was achieved by using metal affinity two-phase system,125 which allowed isolation of ASR-containing vesicles through preferential binding of the His-tagged ASR to the Ni2+ resin. This led to increase of the relative content of the protein of interest to ∼10% of the total protein, and was sufficient to record three-dimensional chemical shift correlation experiments, and transfer spectroscopic assignments previously obtained in proteoliposomes for 40% of residues, located both in the TM and solvent-exposed regions.48
Figure 5.

Sample preparation of ASR in E. coli membranes. Following growth in unlabeled media to the logarithmic phase (Step 1), cells are resuspended in isotopically enriched media at a high concentration, and protein expression is induced (Step 2).43 Cells are broken in Step 3, subjected to sucrose gradient to separate inner and outer membranes (Step 4)124 and to 2 phase purification system to isolate membrane fractions containing the His-tagged ASR (Step 5).125 Proteoliposomes are represented as green circles, ASR is shown as red insertions.
The detected chemical shifts of ASR in the E.coli membrane match very well those measured in the proteoliposomes, indicating, overall, a rigid structure of ASR. ASR remains trimeric in the cellular membrane, which is evident from the minimal perturbations of chemical shifts of residues at the intermonomer interface, as well as from complementary visible range CD spectra, which show a characteristic bilobe shape consistent with the trimer structure.48 Similarly, the all-trans conformation of retinal as well as the conformation of the retinal binding pocket are well conserved. At the same time, we could identify a number of small changes in chemical environment and mobility of certain regions, showing a subtle adaptation to a different environment of the E. coli membrane (Fig. 6).
Figure 6.

ASR in E. coli membrane. A: 16 peaks could be assigned from comparison of 2D NCA spectra of the proteoliposome and E. coli sample (the latter is shown in A). B−D: Much higher resolution was obtained using 3D CANCO and NCACB spectroscopy, which permitted assignments for ∼40% of the protein residues. Red and black contours in the representative 2D planes represent peaks detected in the proteoliposome and E. coli sample, respectively. Blue contours are from the control sample containing the background proteins, but not ASR. E,F: Structural conservation in ASR in E. coli vs. proteoliposomes. In E, residues surrounding retinal are well conserved and shown in green. F: Residues at the intermonomer interface are well conserved (shown in green). Small chemical shift perturbations (up to 1 ppm for 15N and up to 0.5 ppm for 13C) are detected in some of the residues in the helices forming oligomerization interface but facing interior of the protein or lipids (shown in yellow).48 Reprinted from Biophysical Journal, 108, 7, 1683–1696, M. Ward, S. Wang, R. Munro, E. Ritz, I. Hung, P. Gor'kov, Y. Jiang, H. Liang, L. Brown, V. Ladizhansky, “In situ Structural Studies of Anabaena Sensory Rhodopsin in the E. coli Membrane”, with permission from Elsevier.
Current status and future prospects of SSNMR structure determination of membrane proteins
The research described in this review illustrates the potential of solid-state NMR approaches for the characterization of membrane protein structure. High spectral resolution has been demonstrated for several classes of membrane proteins, and currently available sensitivity is already sufficient for providing detailed insights into structure and function of proteins and protein complexes with molecular weights of tens of kilodaltons. So far, SSNMR has resulted in about 70 SSNMR structures including those of large oligomeric membrane protein assemblies, including those of influenza A M2 proton channel,33 phospholamban,35 Yersinia adhesin A (YadA),37 Anabaena Sensory Rhodopsin,38 the chemokine receptor CXCR1,39 DsbB, and of DsbB-DsbA complex41,40 (Fig. 7).
Figure 7.

Selected structures of membrane proteins determined or refined using solid-state NMR restraints. Protein name, molecular weight, lipid mimetic used, method and PDB code are given for each structure. Different colors are used to differentiate individual subunits.
Most importantly, the initial studies of membrane proteins in the cellular environments discussed above indicate that the resolution detected under cellular conditions is comparable to that in proteoliposomes, while application of three-dimensional spectroscopy provides additional resolution sufficient to resolve many resonances. Sensitivity remains the most critical practical limitation. Further success of SSNMR strongly depends on progress in the development of high-field magnets, probe technologies, and proton detection methods. The introduction of ultra-high magnetic fields and cryogenic MAS probes, and the associated sensitivity and resolution improvement will allow an expansion of the list of molecular systems amenable to structural analysis by SSNMR.
An important development in this area is the introduction of Dynamic Nuclear Polarization (DNP) MAS NMR at high magnetic fields, which provides enhancements of SSNMR signals in the range of 10−60 fold. DNP makes use of paramagnetic agents, whose electronic spin polarization is transferred to nuclear spins in the NMR sample, resulting in enhanced NMR signal amplitudes. Pioneering DNP studies on BR allowed observing protein conformational substates not accessible to traditional SSNMR,121,122 and a number of other membrane proteins are being studied by the technique46,126 While this new technique is still under development, and search for new paramagnetic agents and better methods is ongoing,127–130 the most serious limitation for application of DNP for detailed structural studies of membrane proteins is a significant loss of spectral resolution, which complicates DNP studies of uniformly labeled samples. Nevertheless, it promises to be very useful with selectively labeled samples, and may be applied not only to purified proteins with low expression yields but also to in-cell and in-membrane samples.46,131
Acknowledgments
VL thanks Prof. Clemens Glaubitz (Goethe University Frankfurt) for useful discussions. The authors thank Meaghan Ward and Shenlin Wang for help with some of the figures, and Drs. M. Baldus (Utrecht University), B.-J. van Rossum (Leibniz-Institut für Molekulare Pharmakologie), R.G. Griffin (Massachusetts Institute of Technology), M. Eddy (Scripps Research Institute), C.M. Rienstra (University of Illinois at Urbana Champaign), and M. Tang (City University of New York) for providing high-resolution images used in this review.
Glossary
- ASR
Anabaena Sensory Rhodopsin
- BPTI
Bovine Pancreatic Trypsin Inhibitor
- BR
bacteriorhodopsin
- CrgA
mycobacterial cell division regulatory protein
- CD
circular dichroism
- DNP
Dynamic Nuclear Polarization
- DMPA
1,2-dimyristoyl-sn-glycero-3-phosphate
- DMPC
1,2-dimyristoyl-sn-glycero-3-phosphocholine
- DsbA
Disulfide bond formation protein A
- DsbB
Disulfide bond formation protein B
- MAS
magic angle spinning
- NMR
nuclear magnetic resonance
- OmpA
outer membrane protein A
- OmpF
outer membrane protein F
- OmpG
outer membrane protein G
- OmpX
outer membrane protein X
- OS-NMR
oriented sample NMR
- PISA
polarity index slant angle
- PR
proteorhodopsin
- PM
purple membrane
- PRE
paramagnetic relaxation enhancement
- RA
rotational alignment
- SSNMR
solid-state NMR
- VDAC1
voltage-dependent anion channel 1
- YadA
Yersinia enterocolitica adhesin A.
References
- 1.Kendrew JC, Bodo G, Dintzis HM, Parrish RG, Wyckoff H, Phillips DC. A three-dimensional model of the myoglobin molecule obtained by x-ray analysis. Nature. 1958;181:662–666. doi: 10.1038/181662a0. [DOI] [PubMed] [Google Scholar]
- 2.Bill RM, Henderson PJ, Iwata S, Kunji ER, Michel H, Neutze R, Newstead S, Poolman B, Tate CG, Vogel H. Overcoming barriers to membrane protein structure determination. Nat Biotechnol. 2011;29:335–340. doi: 10.1038/nbt.1833. [DOI] [PubMed] [Google Scholar]
- 3.Stroud RM. New tools in membrane protein determination. F1000 Biol Rep. 2011;3:8. doi: 10.3410/B3-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Bertheleme N, Chae PS, Singh S, Mossakowska D, Hann MM, Smith KJ, Hubbard JA, Dowell SJ, Byrne B. Unlocking the secrets of the gatekeeper: methods for stabilizing and crystallizing GPCRs. Biochim Biophys Acta. 2013;1828:2583–2591. doi: 10.1016/j.bbamem.2013.07.013. [DOI] [PubMed] [Google Scholar]
- 5.Bukowska MA, Grutter MG. New concepts and aids to facilitate crystallization. Curr Opin Struct Biol. 2013;23:409–416. doi: 10.1016/j.sbi.2013.03.003. [DOI] [PubMed] [Google Scholar]
- 6.Hiller S, Wagner G. The role of solution NMR in the structure determinations of VDAC-1 and other membrane proteins. Curr Opin Struct Biol. 2009;19:396–401. doi: 10.1016/j.sbi.2009.07.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Kim HJ, Howell SC, Van Horn WD, Jeon YH, Sanders CR. Recent advances in the application of solution NMR spectroscopy to multi-span integral membrane proteins. Prog Nucl Magn Reson Spectrosc. 2009;55:335–360. doi: 10.1016/j.pnmrs.2009.07.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Nietlispach D, Gautier A. Solution NMR studies of polytopic alpha-helical membrane proteins. Curr Opin Struct Biol. 2011;21:497–508. doi: 10.1016/j.sbi.2011.06.009. [DOI] [PubMed] [Google Scholar]
- 9.Maslennikov I, Choe S. Advances in NMR structures of integral membrane proteins. Curr Opin Struct Biol. 2013;23:555–562. doi: 10.1016/j.sbi.2013.05.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Zhou Y, Cierpicki T, Jimenez RH, Lukasik SM, Ellena JF, Cafiso DS, Kadokura H, Beckwith J, Bushweller JH. NMR solution structure of the integral membrane enzyme DsbB: functional insights into DsbB-catalyzed disulfide bond formation. Mol Cell. 2008;31:896–908. doi: 10.1016/j.molcel.2008.08.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Van Horn WD, Kim HJ, Ellis CD, Hadziselimovic A, Sulistijo ES, Karra MD, Tian C, Sonnichsen FD, Sanders CR. Solution nuclear magnetic resonance structure of membrane-integral diacylglycerol kinase. Science. 2009;324:1726–1729. doi: 10.1126/science.1171716. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Gautier A, Mott HR, Bostock MJ, Kirkpatrick JP, Nietlispach D. Structure determination of the seven-helix transmembrane receptor sensory rhodopsin II by solution NMR spectroscopy. Nat Struct Mol Biol. 2010;17:768–774. doi: 10.1038/nsmb.1807. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Berardi MJ, Shih WM, Harrison SC, Chou JJ. Mitochondrial uncoupling protein 2 structure determined by NMR molecular fragment searching. Nature. 2011;476:109–113. doi: 10.1038/nature10257. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Reckel S, Gottstein D, Stehle J, Lohr F, Verhoefen MK, Takeda M, Silvers R, Kainosho M, Glaubitz C, Wachtveitl J, Bernhard F, Schwalbe H, Guntert P, Dotsch V. Solution NMR structure of proteorhodopsin. Angew Chem Int Ed Engl. 2011;50:11942–11946. doi: 10.1002/anie.201105648. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Creuzet F, McDermott A, Gebhard R, van der Hoef K, Spijker-Assink MB, Herzfeld J, Lugtenburg J, Levitt MH, Griffin RG. Determination of membrane protein structure by rotational resonance NMR: bacteriorhodopsin. Science. 1991;251:783–786. doi: 10.1126/science.1990439. [DOI] [PubMed] [Google Scholar]
- 16.Creemers AFL, Kiihne S, Bovee-Geurts PHM, DeGrip WJ, Lugtenburg J, de Groot HJM. H-1 and C-13 MAS NMR evidence for pronounced ligand-protein interactions involving the ionone ring of the retinylidene chromophore in rhodopsin. Proc Natl Acad Sci USA. 2002;99:9101–9106. doi: 10.1073/pnas.112677599. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Luca S, White JF, Sohal AK, Filippov DV, van Boom JH, Grisshammer R, Baldus M. The conformation of neurotensin bound to its G protein-coupled receptor. Proc Natl Acad Sci USA. 2003;100:10706–10711. doi: 10.1073/pnas.1834523100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Lopez JJ, Shukla AK, Reinhart C, Schwalbe H, Michel H, Glaubitz C. The structure of the neuropeptide bradykinin bound to the human G-protein coupled receptor bradykinin B2 as determined by solid-state NMR spectroscopy. Angew Chem Int Ed Engl. 2008;47:1668–1671. doi: 10.1002/anie.200704282. [DOI] [PubMed] [Google Scholar]
- 19.Stringer JA, Bronnimann CE, Mullen CG, Zhou DH, Stellfox SA, Li Y, Williams EH, Rienstra CM. Reduction of RF-induced sample heating with a scroll coil resonator structure for solid-state NMR probes. J Magn Reson. 2005;173:40–48. doi: 10.1016/j.jmr.2004.11.015. [DOI] [PubMed] [Google Scholar]
- 20.Gor'kov PL, Chekmenev EY, Li C, Cotten M, Buffy JJ, Traaseth NJ, Veglia G, Brey WW. Using low-E resonators to reduce RF heating in biological samples for static solid-state NMR up to 900 MHz. J Magn Reson. 2007;185:77–93. doi: 10.1016/j.jmr.2006.11.008. [DOI] [PubMed] [Google Scholar]
- 21.Pauli J, van Rossum B, Forster H, de Groot HJ, Oschkinat H. Sample optimization and identification of signal patterns of amino acid side chains in 2D RFDR spectra of the alpha-spectrin SH3 domain. J Magn Reson. 2000;143:411–416. doi: 10.1006/jmre.2000.2029. [DOI] [PubMed] [Google Scholar]
- 22.McDermott A, Polenova T, Bockmann A, Zilm KW, Paulsen EK, Martin RW, Montelione GT. Partial NMR assignments for uniformly (C-13, N-15)-enriched BPTI in the solid state. J Biomol NMR. 2000;16:209–219. doi: 10.1023/a:1008391625633. [DOI] [PubMed] [Google Scholar]
- 23.Pauli J, Baldus M, van Rossum B, de Groot H, Oschkinat H. Backbone and side-chain 13C and 15N signal assignments of the alpha-spectrin SH3 domain by magic angle spinning solid-state NMR at 17.6 Tesla. Chembiochem. 2001;2:272–281. doi: 10.1002/1439-7633(20010401)2:4<272::AID-CBIC272>3.0.CO;2-2. [DOI] [PubMed] [Google Scholar]
- 24.Castellani F, van Rossum B, Diehl A, Schubert M, Rehbein K, Oschkinat H. Structure of a protein determined by solid-state magic-angle-spinning NMR spectroscopy. Nature. 2002;420:98–102. doi: 10.1038/nature01070. [DOI] [PubMed] [Google Scholar]
- 25.Igumenova TI, McDermott AE, Zilm KW, Martin RW, Paulson EK, Wand AJ. Assignments of carbon NMR resonances for microcrystalline ubiquitin. J Am Chem Soc. 2004;126:6720–6727. doi: 10.1021/ja030547o. [DOI] [PubMed] [Google Scholar]
- 26.Zech SG, Wand AJ, McDermott AE. Protein structure determination by high-resolution solid-state NMR spectroscopy: application to microcrystalline ubiquitin. J Am Chem Soc. 2005;127:8618–8626. doi: 10.1021/ja0503128. [DOI] [PubMed] [Google Scholar]
- 27.Franks WT, Zhou DH, Wylie BJ, Money BG, Graesser DT, Frericks HL, Sahota G, Rienstra CM. Magic-angle spinning solid-state NMR spectroscopy of the beta1 immunoglobulin binding domain of protein G (GB1): 15N and 13C chemical shift assignments and conformational analysis. J Am Chem Soc. 2005;127:12291–12305. doi: 10.1021/ja044497e. [DOI] [PubMed] [Google Scholar]
- 28.Marulanda D, Tasayco ML, Cataldi M, Arriaran V, Polenova T. Resonance assignments and secondary structure analysis of E-coli thioredoxin by magic angle spinning solid-state NMR spectroscopy. J Phys Chem B. 2005;109:18135–18145. doi: 10.1021/jp052774d. [DOI] [PubMed] [Google Scholar]
- 29.Jaroniec CP, MacPhee CE, Bajaj VS, McMahon MT, Dobson CM, Griffin RG. High-resolution molecular structure of a peptide in an amyloid fibril determined by magic angle spinning NMR spectroscopy. Proc Natl Acad Sci USA. 2004;101:711–716. doi: 10.1073/pnas.0304849101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Wasmer C, Lange A, Van Melckebeke H, Siemer AB, Riek R, Meier BH. Amyloid fibrils of the HET-s(218–289) prion form a beta solenoid with a triangular hydrophobic core. Science. 2008;319:1523–1526. doi: 10.1126/science.1151839. [DOI] [PubMed] [Google Scholar]
- 31.Fitzpatrick AW, Debelouchina GT, Bayro MJ, Clare DK, Caporini MA, Bajaj VS, Jaroniec CP, Wang L, Ladizhansky V, Muller SA, MacPhee CE, Waudby CA, Mott HR, De Simone A, Knowles TP, Saibil HR, Vendruscolo M, Orlova EV, Griffin RG, Dobson CM. Atomic structure and hierarchical assembly of a cross-beta amyloid fibril. Proc Natl Acad Sci USA. 2013;110:5468–5473. doi: 10.1073/pnas.1219476110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Lu JX, Qiang W, Yau WM, Schwieters CD, Meredith SC, Tycko R. Molecular structure of beta-amyloid fibrils in Alzheimer's disease brain tissue. Cell. 2013;154:1257–1268. doi: 10.1016/j.cell.2013.08.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Sharma M, Yi M, Dong H, Qin H, Peterson E, Busath DD, Zhou HX, Cross TA. Insight into the mechanism of the influenza A proton channel from a structure in a lipid bilayer. Science. 2010;330:509–512. doi: 10.1126/science.1191750. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Cady SD, Schmidt-Rohr K, Wang J, Soto CS, Degrado WF, Hong M. Structure of the amantadine binding site of influenza M2 proton channels in lipid bilayers. Nature. 2010;463:689–692. doi: 10.1038/nature08722. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Verardi R, Shi L, Traaseth NJ, Walsh N, Veglia G. Structural topology of phospholamban pentamer in lipid bilayers by a hybrid solution and solid-state NMR method. Proc Natl Acad Sci USA. 2011;108:9101–9106. doi: 10.1073/pnas.1016535108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Das N, Dai J, Hung I, Rajagopalan MR, Zhou HX, Cross TA. Structure of CrgA, a cell division structural and regulatory protein from Mycobacterium tuberculosis, in lipid bilayers. Proc Natl Acad Sci USA. 2015;112:E119–E126. doi: 10.1073/pnas.1415908112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Shahid SA, Bardiaux B, Franks WT, Krabben L, Habeck M, van Rossum BJ, Linke D. Membrane-protein structure determination by solid-state NMR spectroscopy of microcrystals. Nat Methods. 2012;9:1212–1217. doi: 10.1038/nmeth.2248. [DOI] [PubMed] [Google Scholar]
- 38.Wang S, Munro RA, Shi L, Kawamura I, Okitsu T, Wada A, Kim SY, Jung KH, Brown LS, Ladizhansky V. Solid-state NMR spectroscopy structure determination of a lipid-embedded heptahelical membrane protein. Nature Methods. 2013;10:1007–1012. doi: 10.1038/nmeth.2635. [DOI] [PubMed] [Google Scholar]
- 39.Park SH, Das BB, Casagrande F, Tian Y, Nothnagel HJ, Chu M, Kiefer H, Maier K, De Angelis AA, Marassi FM, Opella SJ. Structure of the chemokine receptor CXCR1 in phospholipid bilayers. Nature. 2012;491:779–783. doi: 10.1038/nature11580. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Tang M, Nesbitt AE, Sperling LJ, Berthold DA, Schwieters CD, Gennis RB, Rienstra CM. Structure of the disulfide bond generating membrane protein DsbB in the lipid bilayer. J Mol Biol. 2013;425:1670–1682. doi: 10.1016/j.jmb.2013.02.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Tang M, Sperling LJ, Berthold DA, Schwieters CD, Nesbitt AE, Nieuwkoop AJ, Gennis RB, Rienstra CM. High-resolution membrane protein structure by joint calculations with solid-state NMR and X-ray experimental data. J Biomol NMR. 2011;51:227–233. doi: 10.1007/s10858-011-9565-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Fu R, Wang X, Li C, Santiago-Miranda AN, Pielak GJ, Tian F. In situ structural characterization of a recombinant protein in native Escherichia coli membranes with solid-state magic-angle-spinning NMR. J Am Chem Soc. 2011;133:12370–12373. doi: 10.1021/ja204062v. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Jacso T, Franks WT, Rose H, Fink U, Broecker J, Keller S, Oschkinat H, Reif B. Characterization of membrane proteins in isolated native cellular membranes by dynamic nuclear polarization solid-state NMR spectroscopy without purification and reconstitution. Angew Chem Int Ed Engl. 2012;51:432–435. doi: 10.1002/anie.201104987. [DOI] [PubMed] [Google Scholar]
- 44.Kulminskaya NV, Pedersen MO, Bjerring M, Underhaug J, Miller M, Frigaard NU, Nielsen JT, Nielsen NC. In situ solid-state NMR spectroscopy of protein in heterogeneous membranes: the baseplate antenna complex of Chlorobaculum tepidum. Angew Chem Int Ed Engl. 2012;51:6891–6895. doi: 10.1002/anie.201201160. [DOI] [PubMed] [Google Scholar]
- 45.Miao Y, Qin H, Fu R, Sharma M, Can TV, Hung I, Luca S, Gor'kov PL, Brey WW, Cross TA. M2 proton channel structural validation from full-length protein samples in synthetic bilayers and E. coli membranes. Angew Chem Int Ed Engl. 2012;51:8383–8386. doi: 10.1002/anie.201204666. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Renault M, Pawsey S, Bos MP, Koers EJ, Nand D, Tommassen-van Boxtel R, Rosay M, Tommassen J, Maas WE, Baldus M. Solid-state NMR spectroscopy on cellular preparations enhanced by dynamic nuclear polarization. Angew Chem Int Ed Engl. 2012;51:2998–3001. doi: 10.1002/anie.201105984. [DOI] [PubMed] [Google Scholar]
- 47.Renault M, Tommassen-van Boxtel R, Bos MP, Post JA, Tommassen J, Baldus M. Cellular solid-state nuclear magnetic resonance spectroscopy. Proc Natl Acad Sci USA. 2012;109:4863–4868. doi: 10.1073/pnas.1116478109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Ward ME, Wang S, Munro RA, Ritz E, Hung I, Gor'kov PL, Jiang Y, Liang H, Brown LS, Ladizhansky V. In situ structural studies of anabaena sensory rhodopsin in the E. coli membrane. Biophys J. 2015;108:1683–1696. doi: 10.1016/j.bpj.2015.02.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Schmidt P, Lindner D, Montag C, Berndt S, Beck-Sickinger AG, Rudolph R, Huster D. Prokaryotic expression, in vitro folding, and molecular pharmacological characterization of the neuropeptide Y receptor type 2. Biotechnol Prog. 2009;25:1732–1739. doi: 10.1002/btpr.266. [DOI] [PubMed] [Google Scholar]
- 50.Kimura T, Vukoti K, Lynch DL, Hurst DP, Grossfield A, Pitman MC, Reggio PH, Yeliseev AA, Gawrisch K. Global fold of human cannabinoid type 2 receptor probed by solid-state (13) C-, (15) N-MAS NMR and molecular dynamics simulations. Proteins. 2014;82:452–465. doi: 10.1002/prot.24411. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Park SH, Casagrande F, Chu M, Maier K, Kiefer H, Opella SJ. Optimization of purification and refolding of the human chemokine receptor CXCR1 improves the stability of proteoliposomes for structure determination. Biochim Biophys Acta. 2012;1818:584–591. doi: 10.1016/j.bbamem.2011.10.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Fan Y, Shi L, Ladizhansky V, Brown LS. Uniform isotope labeling of a eukaryotic seven-transmembrane helical protein in yeast enables high-resolution solid-state NMR studies in the lipid environment. J Biomol NMR. 2011;49:151–161. doi: 10.1007/s10858-011-9473-9. [DOI] [PubMed] [Google Scholar]
- 53.Emami S, Fan Y, Munro R, Ladizhansky V, Brown LS. Yeast-expressed human membrane protein aquaporin-1 yields excellent resolution of solid-state MAS NMR spectra. J Biomol NMR. 2013;55:147–155. doi: 10.1007/s10858-013-9710-5. [DOI] [PubMed] [Google Scholar]
- 54.LeMaster DM, Kushlan DM. Dynamical mapping of E-coli thioredoxin via C-13 NMR relaxation analysis. J Am Chem Soc. 1996;118:9255–9264. [Google Scholar]
- 55.Hong M, Jakes K. Selective and extensive C-13 labeling of a membrane protein for solid-state NMR investigations. J Biomol NMR. 1999;14:71–74. doi: 10.1023/a:1008334930603. [DOI] [PubMed] [Google Scholar]
- 56.Loquet A, Giller K, Becker S, Lange A. Supramolecular interactions probed by 13C-13C solid-state NMR spectroscopy. J Am Chem Soc. 2010;132:15164–15166. doi: 10.1021/ja107460j. [DOI] [PubMed] [Google Scholar]
- 57.Loquet A, Lv G, Giller K, Becker S, Lange A. 13C spin dilution for simplified and complete solid-state NMR resonance assignment of insoluble biological assemblies. J Am Chem Soc. 2011;133:4722–4725. doi: 10.1021/ja200066s. [DOI] [PubMed] [Google Scholar]
- 58.Higman VA, Flinders J, Hiller M, Jehle S, Markovic S, Fiedler S, van Rossum BJ, Oschkinat H. Assigning large proteins in the solid state: a MAS NMR resonance assignment strategy using selectively and extensively 13C-labelled proteins. J Biomol NMR. 2009;44:245–260. doi: 10.1007/s10858-009-9338-7. [DOI] [PubMed] [Google Scholar]
- 59.Sperling LJ, Berthold DA, Sasser TL, Jeisy-Scott V, Rienstra CM. Assignment strategies for large proteins by magic-angle spinning NMR: the 21-kDa disulfide-bond-forming enzyme DsbA. J Mol Biol. 2010;399:268–282. doi: 10.1016/j.jmb.2010.04.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Sugiki T, Ichikawa O, Miyazawa-Onami M, Shimada I, Takahashi H. Isotopic labeling of heterologous proteins in the yeast Pichia pastoris and Kluyveromyces lactis. Methods Mol Biol. 2012;831:19–36. doi: 10.1007/978-1-61779-480-3_2. [DOI] [PubMed] [Google Scholar]
- 61.Miyazawa-Onami M, Takeuchi K, Takano T, Sugiki T, Shimada I, Takahashi H. Perdeuteration and methyl-selective (1)H, (13)C-labeling by using a Kluyveromyces lactis expression system. J Biomol NMR. 2013;57:297–304. doi: 10.1007/s10858-013-9789-8. [DOI] [PubMed] [Google Scholar]
- 62.Egorova-Zachernyuk TA, Bosman GJ, Degrip WJ, Shvets VI. Stable isotope labelling of human histamine receptor H1R: prospects for structure-based drug design. Dokl Biochem Biophys. 2010;433:164–167. doi: 10.1134/S160767291004006X. [DOI] [PubMed] [Google Scholar]
- 63.Egorova-Zachernyuk TA, Bosman GJ, Degrip WJ. Uniform stable-isotope labeling in mammalian cells: formulation of a cost-effective culture medium. Appl Microbiol Biotechnol. 2011;89:397–406. doi: 10.1007/s00253-010-2896-5. [DOI] [PubMed] [Google Scholar]
- 64.Ahuja S, Hornak V, Yan EC, Syrett N, Goncalves JA, Hirshfeld A, Ziliox M, Sakmar TP, Sheves M, Reeves PJ, Smith SO, Eilers M. Helix movement is coupled to displacement of the second extracellular loop in rhodopsin activation. Nat Struct Mol Biol. 2009;16:168–175. doi: 10.1038/nsmb.1549. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Goncalves JA, South K, Ahuja S, Zaitseva E, Opefi CA, Eilers M, Vogel R, Reeves PJ, Smith SO. Highly conserved tyrosine stabilizes the active state of rhodopsin. Proc Natl Acad Sci USA. 2010;107:19861–19866. doi: 10.1073/pnas.1009405107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Salom D, Cao P, Yuan Y, Miyagi M, Feng Z, Palczewski K. Isotopic labeling of mammalian G protein-coupled receptors heterologously expressed in Caenorhabditis elegans. Anal Biochem. 2015;472:30–36. doi: 10.1016/j.ab.2014.11.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Sobhanifar S, Reckel S, Junge F, Schwarz D, Kai L, Karbyshev M, Lohr F, Bernhard F, Dotsch V. Cell-free expression and stable isotope labelling strategies for membrane proteins. J Biomol NMR. 2010;46:33–43. doi: 10.1007/s10858-009-9364-5. [DOI] [PubMed] [Google Scholar]
- 68.Klammt C, Maslennikov I, Bayrhuber M, Eichmann C, Vajpai N, Chiu EJ, Blain KY, Esquivies L, Kwon JH, Balana B, Pieper U, Sali A, Slesinger PA, Kwiatkowski W, Riek R, Choe S. Facile backbone structure determination of human membrane proteins by NMR spectroscopy. Nat Methods. 2012;9:834–839. doi: 10.1038/nmeth.2033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Linser R, Gelev V, Hagn F, Arthanari H, Hyberts SG, Wagner G. Selective methyl labeling of eukaryotic membrane proteins using cell-free expression. J Am Chem Soc. 2014;136:11308–11310. doi: 10.1021/ja504791j. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Abdine A, Verhoeven MA, Park KH, Ghazi A, Guittet E, Berrier C, Van Heijenoort C, Warschawski DE. Structural study of the membrane protein MscL using cell-free expression and solid-state NMR. J Magn Reson. 2010;204:155–159. doi: 10.1016/j.jmr.2010.02.003. [DOI] [PubMed] [Google Scholar]
- 71.Kang SJ, Todokoro Y, Yumen I, Shen B, Iwasaki I, Suzuki T, Miyagi A, Yoshida M, Fujiwara T, Akutsu H. Active-site structure of the thermophilic Foc-subunit ring in membranes elucidated by solid-state NMR. Biophys J. 2014;106:390–398. doi: 10.1016/j.bpj.2013.12.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Sanders CR, Hare BJ, Howard KP, Prestegard JH. Magnetically oriented phospholipid micelles as a tool for the study of membrane associated molecules. Prog Nucl Magn Reson Spectrosc. 1994;26:421–444. [Google Scholar]
- 73.De Angelis AA, Opella SJ. Bicelle samples for solid-state NMR of membrane proteins. Nat Protoc. 2007;2:2332–2338. doi: 10.1038/nprot.2007.329. [DOI] [PubMed] [Google Scholar]
- 74.Park SH, Casagrande F, Cho L, Albrecht L, Opella SJ. Interactions of interleukin-8 with the human chemokine receptor CXCR1 in phospholipid bilayers by NMR spectroscopy. J Mol Biol. 2011;414:194–203. doi: 10.1016/j.jmb.2011.08.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Mahalakshmi R, Marassi FM. Orientation of the Escherichia coli outer membrane protein OmpX in phospholipid bilayer membranes determined by solid-State NMR. Biochemistry. 2008;47:6531–6538. doi: 10.1021/bi800362b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Durr UH, Yamamoto K, Im SC, Waskell L, Ramamoorthy A. Solid-state NMR reveals structural and dynamical properties of a membrane-anchored electron-carrier protein, cytochrome b5. J Am Chem Soc. 2007;129:6670–6671. doi: 10.1021/ja069028m. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Marassi FM, Opella SJ. A solid-state NMR index of helical membrane protein structure and topology. J Magn Reson. 2000;144:150–155. doi: 10.1006/jmre.2000.2035. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Wang J, Denny J, Tian C, Kim S, Mo Y, Kovacs F, Song Z, Nishimura K, Gan Z, Fu R, Quine JR, Cross TA. Imaging membrane protein helical wheels. J Magn Reson. 2000;144:162–167. doi: 10.1006/jmre.2000.2037. [DOI] [PubMed] [Google Scholar]
- 79.Gayen A, Banigan JR, Traaseth NJ. Ligand-induced conformational changes of the multidrug resistance transporter EmrE probed by oriented solid-state NMR spectroscopy. Angew Chem Int Ed Engl. 2013;52:10321–10324. doi: 10.1002/anie.201303091. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Park SH, Mrse AA, Nevzorov AA, Mesleh MF, Oblatt-Montal M, Montal M, Opella SJ. Three-dimensional structure of the channel-forming trans-membrane domain of virus protein "u" (Vpu) from HIV-1. J Mol Biol. 2003;333:409–424. doi: 10.1016/j.jmb.2003.08.048. [DOI] [PubMed] [Google Scholar]
- 81.Shi L, Kawamura I, Jung KH, Brown LS, Ladizhansky V. Conformation of a seven-helical transmembrane photosensor in the lipid environment. Angew Chem Int Ed Engl. 2011;50:1302–1305. doi: 10.1002/anie.201004422. [DOI] [PubMed] [Google Scholar]
- 82.Shahid SA, Markovic S, Linke D, van Rossum BJ. Assignment and secondary structure of the YadA membrane protein by solid-state MAS NMR. Sci Rep. 2012;2:803. doi: 10.1038/srep00803. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Li Y, Berthold DA, Gennis RB, Rienstra CM. Chemical shift assignment of the transmembrane helices of DsbB, a 20-kDa integral membrane enzyme, by 3D magic-angle spinning NMR spectroscopy. Protein Sci. 2008;17:199–204. doi: 10.1110/ps.073225008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Martin RW, Zilm KW. Preparation of protein nanocrystals and their characterization by solid state NMR. J Magn Reson. 2003;165:162–174. doi: 10.1016/s1090-7807(03)00253-2. [DOI] [PubMed] [Google Scholar]
- 85.Schneider R, Ader C, Lange A, Giller K, Hornig S, Pongs O, Becker S, Baldus M. Solid-state NMR spectroscopy applied to a chimeric potassium channel in lipid bilayers. J Am Chem Soc. 2008;130:7427–7435. doi: 10.1021/ja800190c. [DOI] [PubMed] [Google Scholar]
- 86.Eddy MT, Ong TC, Clark L, Teijido O, van der Wel PC, Garces R, Wagner G, Rostovtseva TK, Griffin RG. Lipid dynamics and protein-lipid interactions in 2D crystals formed with the beta-barrel integral membrane protein VDAC1. J Am Chem Soc. 2012;134:6375–6387. doi: 10.1021/ja300347v. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Hiller M, Krabben L, Vinothkumar KR, Castellani F, van Rossum BJ, Kuhlbrandt W, Oschkinat H. Solid-state magic-angle spinning NMR of outer-membrane protein G from Escherichia coli. Chembiochem. 2005;6:1679–1684. doi: 10.1002/cbic.200500132. [DOI] [PubMed] [Google Scholar]
- 88.Varga K, Aslimovska L, Watts A. Advances towards resonance assignments for uniformly–13C, 15N enriched bacteriorhodopsin at 18.8 T in purple membranes. J Biomol NMR. 2008;41:1–4. doi: 10.1007/s10858-008-9235-5. [DOI] [PubMed] [Google Scholar]
- 89.Liang H, Whited G, Nguyen C, Stucky GD. The directed cooperative assembly of proteorhodopsin into 2D and 3D polarized arrays. Proc Natl Acad Sci USA. 2007;104:8212–8217. doi: 10.1073/pnas.0702336104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Shastri S, Vonck J, Pfleger N, Haase W, Kuehlbrandt W, Glaubitz C. Proteorhodopsin: characterisation of 2D crystals by electron microscopy and solid state NMR. Biochim Biophys Acta. 2007;1768:3012–3019. doi: 10.1016/j.bbamem.2007.10.001. [DOI] [PubMed] [Google Scholar]
- 91.Klyszejko AL, Shastri S, Mari SA, Grubmuller H, Muller DJ, Glaubitz C. Folding and assembly of proteorhodopsin. J Mol Biol. 2008;376:35–41. doi: 10.1016/j.jmb.2007.11.030. [DOI] [PubMed] [Google Scholar]
- 92.Pfleger N, Worner AC, Yang J, Shastri S, Hellmich UA, Aslimovska L, Maier MS, Glaubitz C. Solid-state NMR and functional studies on proteorhodopsin. Biochim Biophys Acta. 2009;1787:697–705. doi: 10.1016/j.bbabio.2009.02.022. [DOI] [PubMed] [Google Scholar]
- 93.Shi L, Ahmed MA, Zhang W, Whited G, Brown LS, Ladizhansky V. Three-dimensional solid-state NMR study of a seven-helical integral membrane proton pump–structural insights. J Mol Biol. 2009;386:1078–1093. doi: 10.1016/j.jmb.2009.01.011. [DOI] [PubMed] [Google Scholar]
- 94.Linser R, Dasari M, Hiller M, Higman V, Fink U, Lopez del Amo JM, Markovic S, Handel L, Kessler B, Schmieder P, Oesterhelt D, Oschkinat H, Reif B. Proton-detected solid-state NMR spectroscopy of fibrillar and membrane proteins. Angew Chem Int Ed Engl. 2011;50:4508–4512. doi: 10.1002/anie.201008244. [DOI] [PubMed] [Google Scholar]
- 95.Ward ME, Shi L, Lake E, Krishnamurthy S, Hutchins H, Brown LS, Ladizhansky V. Proton-detected solid-state NMR reveals intramembrane polar networks in a seven-helical transmembrane protein proteorhodopsin. J Am Chem Soc. 2011;133:17434–17443. doi: 10.1021/ja207137h. [DOI] [PubMed] [Google Scholar]
- 96.Zhou DH, Nieuwkoop AJ, Berthold DA, Comellas G, Sperling LJ, Tang M, Shah GJ, Brea EJ, Lemkau LR, Rienstra CM. Solid-state NMR analysis of membrane proteins and protein aggregates by proton detected spectroscopy. J Biomol NMR. 2012;54:291–305. doi: 10.1007/s10858-012-9672-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Barbet-Massin E, Pell AJ, Retel JS, Andreas LB, Jaudzems K, Franks WT, Nieuwkoop AJ, Hiller M, Higman V, Guerry P, Bertarello A, Knight MJ, Felletti M, Le Marchand T, Kotelovica S, Akopjana I, Tars K, Stoppini M, Bellotti V, Bolognesi M, Ricagno S, Chou JJ, Griffin RG, Oschkinat H, Lesage A, Emsley L, Herrmann T, Pintacuda G. Rapid proton-detected NMR assignment for proteins with fast magic angle spinning. J Am Chem Soc. 2014;136:12489–12497. doi: 10.1021/ja507382j. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Chevelkov V, van Rossum BJ, Castellani F, Rehbein K, Diehl A, Hohwy M, Steuernagel S, Engelke F, Oschkinat H, Reif B. H-1 detection in MAS solid-state NMR Spectroscopy of biomacromolecules employing pulsed field gradients for residual solvent suppression. J Am Chem Soc. 2003;125:7788–7789. doi: 10.1021/ja029354b. [DOI] [PubMed] [Google Scholar]
- 99.Chevelkov V, Rehbein K, Diehl A, Reif B. Ultrahigh resolution in proton solid-state NMR spectroscopy at high levels of deuteration. Angew Chem Int Ed. 2006;45:3878–3881. doi: 10.1002/anie.200600328. [DOI] [PubMed] [Google Scholar]
- 100.Goto NK, Kay LE. New developments in isotope labeling strategies for protein solution NMR spectroscopy. Curr Opin Struct Biol. 2000;10:585–592. doi: 10.1016/s0959-440x(00)00135-4. [DOI] [PubMed] [Google Scholar]
- 101.Asami S, Schmieder P, Reif B. High resolution 1H-detected solid-state NMR spectroscopy of protein aliphatic resonances: access to tertiary structure information. J Am Chem Soc. 2010;132:15133–15135. doi: 10.1021/ja106170h. [DOI] [PubMed] [Google Scholar]
- 102.Shi L, Lake EM, Ahmed MA, Brown LS, Ladizhansky V. Solid-state NMR study of proteorhodopsin in the lipid environment: secondary structure and dynamics. Biochim Biophys Acta. 2009;1788:2563–2574. doi: 10.1016/j.bbamem.2009.09.011. [DOI] [PubMed] [Google Scholar]
- 103.Wylie BJ, Bhate MP, McDermott AE. Transmembrane allosteric coupling of the gates in a potassium channel. Proc Natl Acad Sci USA. 2014;111:185–190. doi: 10.1073/pnas.1319577110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Etzkorn M, Martell S, Andronesi OC, Seidel K, Engelhard M, Baldus M. Secondary structure, dynamics, and topology of a seven-helix receptor in native membranes, studied by solid-state NMR spectroscopy. Angew Chem Int Ed Engl. 2007;46:459–462. doi: 10.1002/anie.200602139. [DOI] [PubMed] [Google Scholar]
- 105.Moseley HN, Sperling LJ, Rienstra CM. Automated protein resonance assignments of magic angle spinning solid-state NMR spectra of beta1 immunoglobulin binding domain of protein G (GB1) J Biomol NMR. 2010;48:123–128. doi: 10.1007/s10858-010-9448-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Hu KN, Qiang W, Tycko R. A general Monte Carlo/simulated annealing algorithm for resonance assignment in NMR of uniformly labeled biopolymers. J Biomol NMR. 2011;50:267–276. doi: 10.1007/s10858-011-9517-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Helmus JJ, Jaroniec CP. Nmrglue: an open source Python package for the analysis of multidimensional NMR data. J Biomol NMR. 2013;55:355–367. doi: 10.1007/s10858-013-9718-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Wishart DS, Sykes BD, Richards FM. The chemical-shift index—a fast and simple method for the assignment of protein secondary structure through Nmr-Spectroscopy. Biochemistry. 1992;31:1647–1651. doi: 10.1021/bi00121a010. [DOI] [PubMed] [Google Scholar]
- 109.Shen Y, Delaglio F, Cornilescu G, Bax A. TALOS+: a hybrid method for predicting protein backbone torsion angles from NMR chemical shifts. J Biomol NMR. 2009;44:213–223. doi: 10.1007/s10858-009-9333-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Franks WT, Wylie BJ, Schmidt HL, Nieuwkoop AJ, Mayrhofer RM, Shah GJ, Graesser DT, Rienstra CM. Dipole tensor-based atomic-resolution structure determination of a nanocrystalline protein by solid-state NMR. Proc Natl Acad Sci USA. 2008;105:4621–4626. doi: 10.1073/pnas.0712393105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Wylie BJ, Sperling LJ, Nieuwkoop AJ, Franks WT, Oldfield E, Rienstra CM. Ultrahigh resolution protein structures using NMR chemical shift tensors. Proc Natl Acad Sci USA. 2011;108:16974–16979. doi: 10.1073/pnas.1103728108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Helmus JJ, Nadaud PS, Hofer N, Jaroniec CP. Determination of methyl 13C-15N dipolar couplings in peptides and proteins by three-dimensional and four-dimensional magic-angle spinning solid-state NMR spectroscopy. J Chem Phys. 2008;128:052314. doi: 10.1063/1.2817638. [DOI] [PubMed] [Google Scholar]
- 113.Nieuwkoop AJ, Wylie BJ, Franks WT, Shah GJ, Rienstra CM. Atomic resolution protein structure determination by three-dimensional transferred echo double resonance solid-state nuclear magnetic resonance spectroscopy. J Chem Phys. 2009;131:095101. doi: 10.1063/1.3211103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Nadaud PS, Helmus JJ, Hofer N, Jaroniec CP. Long-range structural restraints in spin-labeled proteins probed by solid-state nuclear magnetic resonance spectroscopy. J Am Chem Soc. 2007;129:7502–7503. doi: 10.1021/ja072349t. [DOI] [PubMed] [Google Scholar]
- 115.Nadaud PS, Helmus JJ, Kall SL, Jaroniec CP. Paramagnetic ions enable tuning of nuclear relaxation rates and provide long-range structural restraints in solid-state NMR of proteins. J Am Chem Soc. 2009;131:8108–8120. doi: 10.1021/ja900224z. [DOI] [PubMed] [Google Scholar]
- 116.Wang S, Munro RA, Kim SY, Jung KH, Brown LS, Ladizhansky V. Paramagnetic relaxation enhancement reveals oligomerization interface of a membrane protein. J Am Chem Soc. 2012;134:16995–16998. doi: 10.1021/ja308310z. [DOI] [PubMed] [Google Scholar]
- 117.Lewis BA, Harbison GS, Herzfeld J, Griffin RG. NMR structural analysis of a membrane protein: bacteriorhodopsin peptide backbone orientation and motion. Biochemistry. 1985;24:4671–4679. doi: 10.1021/bi00338a029. [DOI] [PubMed] [Google Scholar]
- 118.Das BB, Nothnagel HJ, Lu GJ, Son WS, Tian Y, Marassi FM, Opella SJ. Structure determination of a membrane protein in proteoliposomes. J Am Chem Soc. 2012;134:2047–2056. doi: 10.1021/ja209464f. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Heyn MP, Cherry RJ, Dencher NA. Lipid–protein interactions in bacteriorhodopsin−dimyristoylphosphatidylcholine vesicles. Biochemistry. 1981;20:840–849. doi: 10.1021/bi00507a029. [DOI] [PubMed] [Google Scholar]
- 120.Saito H, Naito A. NMR studies on fully hydrated membrane proteins, with emphasis on bacteriorhodopsin as a typical and prototype membrane protein. Biochim Biophys Acta. 2007;1768:3145–3161. doi: 10.1016/j.bbamem.2007.08.026. [DOI] [PubMed] [Google Scholar]
- 121.Mak-Jurkauskas ML, Bajaj VS, Hornstein MK, Belenky M, Griffin RG, Herzfeld J. Energy transformations early in the bacteriorhodopsin photocycle revealed by DNP-enhanced solid-state NMR. Proc Natl Acad Sci USA. 2008;105:883–888. doi: 10.1073/pnas.0706156105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Bajaj VS, Mak-Jurkauskas ML, Belenky M, Herzfeld J, Griffin RG. Functional and shunt states of bacteriorhodopsin resolved by 250 GHz dynamic nuclear polarization-enhanced solid-state NMR. Proc Natl Acad Sci USA. 2009;106:9244–9249. doi: 10.1073/pnas.0900908106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Higman VA, Varga K, Aslimovska L, Judge PJ, Sperling LJ, Rienstra CM, Watts A. The conformation of bacteriorhodopsin loops in purple membranes resolved by solid-state MAS NMR spectroscopy. Angew Chem Int Ed Engl. 2011;50:8432–8435. doi: 10.1002/anie.201100730. [DOI] [PubMed] [Google Scholar]
- 124.Marani P, Wagner S, Baars L, Genevaux P, de Gier JW, Nilsson I, Casadio R, von Heijne G. New Escherichia coli outer membrane proteins identified through prediction and experimental verification. Protein Sci. 2006;15:884–889. doi: 10.1110/ps.051889506. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Everberg H, Clough J, Henderson P, Jergil B, Tjerneld F, Ramirez IB. Isolation of Escherichia coli inner membranes by metal affinity two-phase partitioning. J Chromat A. 2006;1118:244–252. doi: 10.1016/j.chroma.2006.03.123. [DOI] [PubMed] [Google Scholar]
- 126.Koers EJ, van der Cruijsen EA, Rosay M, Weingarth M, Prokofyev A, Sauvee C, Ouari O, van der Zwan J, Pongs O, Tordo P, Maas WE, Baldus M. NMR-based structural biology enhanced by dynamic nuclear polarization at high magnetic field. J Biomol NMR. 2014;60:157–168. doi: 10.1007/s10858-014-9865-8. [DOI] [PubMed] [Google Scholar]
- 127.Song CS, Hu KN, Joo CG, Swager TM, Griffin RG. TOTAPOL: a biradical polarizing agent for dynamic nuclear polarization experiments in aqueous media. J Am Chem Soc. 2006;128:11385–11390. doi: 10.1021/ja061284b. [DOI] [PubMed] [Google Scholar]
- 128.Matsuki Y, Maly T, Ouari O, Karoui H, Le Moigne F, Rizzato E, Lyubenova S, Herzfeld J, Prisner T, Tordo P, Griffin RG. Dynamic nuclear polarization with a rigid biradical. Angew Chem Int Ed Engl. 2009;48:4996–5000. doi: 10.1002/anie.200805940. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Sauvee C, Rosay M, Casano G, Aussenac F, Weber RT, Ouari O, Tordo P. Highly efficient, water-soluble polarizing agents for dynamic nuclear polarization at high frequency. Angew Chem Int Ed Engl. 2013;52:10858–10861. doi: 10.1002/anie.201304657. [DOI] [PubMed] [Google Scholar]
- 130.Zagdoun A, Casano G, Ouari O, Schwarzwalder M, Rossini AJ, Aussenac F, Yulikov M, Jeschke G, Coperet C, Lesage A, Tordo P, Emsley L. Large molecular weight nitroxide biradicals providing efficient dynamic nuclear polarization at temperatures up to 200 K. J Am Chem Soc. 2013;135:12790–12797. doi: 10.1021/ja405813t. [DOI] [PubMed] [Google Scholar]
- 131.Yamamoto K, Caporini MA, Im SC, Waskell L, Ramamoorthy A. Cellular solid-state NMR investigation of a membrane protein using dynamic nuclear polarization. Biochim Biophys Acta. 2015;1848:342–349. doi: 10.1016/j.bbamem.2014.07.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
