Comprehensive Physiology Wiley Online Library

Structure and Dynamic Properties of Membrane Proteins using NMR

Full Article on Wiley Online Library



Abstract

Integral membrane proteins are one of the most challenging groups of macromolecules despite their apparent conformational simplicity. They manage and drive transport, circulate information, and participate in cellular movements via interactions with other proteins and through intricate conformational changes. Their structural and functional decoding is challenging and has imposed demanding experimental development. Solution nuclear magnetic resonance (NMR) spectroscopy is one of the techniques providing the capacity to make a significant difference in the deciphering of the membrane protein structure‐function paradigm. The method has evolved dramatically during the last decade resulting in a plethora of new experiments leading to a significant increase in the scientific repertoire for studying membrane proteins. Besides solving the three‐dimensional structures using state‐of‐the‐art approaches, a large variety of developments of well‐established techniques are available providing insight into membrane protein flexibility, dynamics, and interactions. Inspired by the speed of development in the application of new strategies, by invention of methods to measure solvent accessibility and describe low‐populated states, this review seeks to introduce the vast possibilities solution NMR can offer to the study of membrane protein structure‐function analyses with special focus on applicability. © 2012 American Physiological Society. Compr Physiol 2:1491‐1539, 2012.

Comprehensive Physiology offers downloadable PowerPoint presentations of figures for non-profit, educational use, provided the content is not modified and full credit is given to the author and publication.

Download a PowerPoint presentation of all images


Figure 1. Figure 1.

Structures of 18 selected membrane protein solved by solution NMR spectroscopy. Since 2001, only around 20 structures of integral membrane proteins have been solved in solution and are publicly available. The figure has been divided into two parts, β‐barrel structures in the upper part and α‐helical structures in the lower part. The structures are (A) OmpX, PDB code 1Q9F 78, (B) OmpA, PDB code 2GEA 7, (C) PagP, PDB code 1MM4 111, (D) OmpG, PDB code 2JQY 75, (E) human VDAC‐1, PDB code 2JK4 102, (F) KpOmpA, PDB code 2K0L 119,205, (G) potassium channel KcsA, PDB code 2A9H 10, (H) diacylglycerol kinase, DAGK, PDB code 2KDC 259, (I) phospholamban, PDB code 1ZLL 192, (J) influenza A M2 proton channel, PDB code 2RLF 219, (K) influenza B M2 proton channel, PDB code 2KIX 267, (L) DsbB, PDB code 2K73 289, (M) sensory rhodopsin II, PDB code 2KSY 86, (N) n‐acetylcholine receptor, PDB code 2KSR 28, (O) KCNE1, PDB code 2K21 123, (P) FXYD1, PDB code 2JO1 246, (Q) Rv1761c, PDB code 2K3M 193, and (R) subunit c of the F0F1 ATPase synthase, PDB code 1A91 80. Images of structures generated in PyMol (DeLano Scientific).

Figure 2. Figure 2.

Choice of lipids for NMR characterization. At the top, a schematic illustration of membrane protein extraction and reconstitution into detergents suitable for structural analysis is shown. The detergent needs to preserve the native fold of the internal membrane part. However, some detergents will also affect external parts and even denature those, which is detrimental for structure analysis. Later, an overview of various commonly used detergents applied in solution nuclear magnetic resonance (NMR) studies of IMPs is provided. The chemical structures are grouped according to their nature into organic solvents, nonionic, ionic, zwitterionic, and sterol‐based detergents. For each chemical structure, the critical micelle concentration (cmc) and the aggregation number (AG number) are listed, which are taken either from 215 or from the suppliers.

Figure 3. Figure 3.

The NMR assignment process. Typically by selecting different types of NMR spectra, one selects intra‐residue signals versus inter‐residue signals in a 3D‐spectrum. By a combination of such spectra, which often is nine or more 3D spectra, and together with the knowledge of the sequence of the protein, one performs the job of linking the chemical shifts to distinct nuclei. This process is called assignment. In the assignment process, the NMR signals are linked to the correct nuclei in the protein. (A) A 1D1H‐NMR spectrum of a protein with the chemical shifts of the NMR signals on the x‐axis, approximately going from −1 ppm to 11 ppm. The large overlap between signals demands for extraction into more dimensions. (B) A 2D15N,1H‐HSQC spectrum of a protein with NMR signals from all protons covalently bound to a nitrogen. Here the overlap is minimized dramatically. (C) A 3D HNCO‐spectrum of a protein showing NMR signals from amide protons covalently bound to nitrogen, which again is covalently bound to a carbonyl atom. This type of spectrum requires stable isotope labeling with 15N and 13C. (D) A schematic description of the assignment process. From knowledge of the chemical structure of the peptide chain and the sequence of the protein (E), one uses a set of six to ten 3D heteronuclear NMR spectra to assign the individual NMR signals to the individual atoms of the protein. (F) A schematic representation of the sequential walk, where signals from six different strips of a theoretical HNCACB 3D spectrum are shown. One strip represents carbon signals correlating to a single N,HN. In each strip the N,HN correlates to four carbons, the Cα and Cβ from its own residue and the Cα and Cβ from the preceding residue. This directly links the strips (see arrows). The challenge in assignment is to identify which strip, i.e. which residue is a neighbor to which. (G) An example of the magnetization transfer steps exploited to generate two of the triple‐resonance spectra HNCACB (left) and CBCA(CO)NH (right) important for the assignment. The fist spectrum (left) provides both intraresidual and sequential signals and the second spectrum (right) selects only sequential signals. Arrows indicate the route of transfer and the number the order by which magnetization is transferred from one nucleus to the next. The connections are between residue i and the preceding residue i − 1. All carbons are 13C‐labeled and all nitrogens 15N‐labeled.

Figure 4. Figure 4.

HSQC spectra with and without TROSY‐a tool to circumvent the size problem. Comparison of (A) a conventional 2D [1H, 15N]‐HSQC and (B) a 2D [1H, 15N]‐TROSY spectrum. Both spectra were recorded on a 2 mmol/L OmpX solution in DHPC micelles at 30°C at a resonance frequency of 750 Mhz. The superior performance of the TROSY experiment clearly manifests in the better resolution of the individual cross peaks in the TROSY spectrum. (C) Solution nuclear magnetic resonance (NMR) structure of OmpX in DHPC micelles [20, PDB‐code 1Q9F]. The TROSY effect is largest at very strong magnetic fields. Parts (A) and (B) are modified, with permission, from reference 77. Copyright 2001 National Academy of Sciences USA.

Figure 5. Figure 5.

NMR and the size problem. (A) Molecules tumble freely in solution. However, larger proteins take longer time to tumble around their own axis than small molecules. Left: illustration of a large molecule with slow tumbling, and right: a small molecule with fast tumbling. (B) A nuclear magnetic resonance (NMR) signal is recorded after the application of a radiofrequency pulse that perturbs the equilibrium distribution of spins. One records the return to equilibrium that is an exponential process and is called the free‐induction decay, FID (in red waves). To obtain the NMR spectrum, the FID undergoes a Fourier transformation (FT)—a mathematical operation. For large molecules tumbling more slowly (left), the corresponding NMR signal disappears (relaxes) quickly and the intensity of the signal is dampened, whereas for small molecules (right), the NMR signal is more long‐lived, giving rise to more intense and sharper signals.

Figure 6. Figure 6.

Typical NMR‐derived structural restraints. A number of different conformational restraints can be derived from different types of nuclear magnetic resonance (NMR) experiments. Chemical shifts‐derived dihedral angles restraints: when chemical shifts are assigned to the nuclei, the values are compared to those of databases that link chemical shifts to structures. Hereby one obtains the likelihood for the backbone dihedral angles φ and ψ to be in specific conformations, that is, occupying specific regions of the Ramachandran plot. The dihedral angles are then included together with an allowed angle deviation as input to the structure calculation. NOE‐derived distance restraints: through cross‐relaxation through space, one obtains a distance dependence of the intensity of the NMR signal that declines with the six‐root of the distance. This is termed the NOE effect. This means that only short‐range distances (r), can be detected, typically up to 5 Å. Thus, the intensity of the NOE signal determines the range of the distance restraint, where a strong peak typically is interpreted as a distance being closer than 2.5 Å and weak peak as a distance closer than 5 Å. RDC‐derived global restraints: residual dipolar couplings (RDCs) are global restraints that relate the bond vector to the laboratory frame. From a change in the splitting of the NMR signal, either to larger or smaller splitting, the angles relating the N‐H vector to the laboratory frame can be determined for each vector of the molecule. PRE‐derived distance restraints: paramagnetic relaxation enhancements (PREs) are long‐range distance restraints. The presence of a paramagnetic center enhances the relaxation process and thus lowers the intensity of the NMR signals in its surroundings. For the three cases described here, atom a is closest to the paramagnetic center and thus its NMR signal will be most affected, whereas atom c is too far away to be affected and b is in between.

Figure 7. Figure 7.

Structure calculation scheme. Flowchart describing a possible procedure for protein structure determination by solution nuclear magnetic resonance (NMR) spectroscopy. After protein production and purification a large set of triple resonance, heteronuclear three‐dimensional (3D) NMR spectra are recorded. The chemical shifts are assigned and restraints are derived from chemical shifts, NOEs, PREs, RDCs, etc. Together with information of the chemical structure and sequence, these restraints are fed into a computer algorithm that iteratively calculates and refines the structure and finally yields an ensemble of structures such as shown here for OmpX, fulfilling the experimental restraints. The red arrow indicated the iterative cycle of refinements, which often result in numerous cycles.

Figure 8. Figure 8.

Paramagnetic relaxation enhancement (PRE) and its application in membrane protein structure characterization. PREs can be used to measure long‐range distance restraints in proteins as illustrated in (A) or determine solvent/detergent exposure as illustrated in (B). The 2D [1H, 15N]—HSQC spectrum of Rv1761c in the presence of a (A) reduced and an (B) oxidized MTSL‐label in position F30C. The circles in (B) highlight the residues whose signals are missing in the oxidized spectrum. A more quantitative illustration showing the intensity ratio of reduced (para) and oxidized (dia) cross peaks (Ired/Iox) is shown in (C). An asterisk marks the position of the spin label. From this pattern, additional structural restraints could be deduced. The inset shows the solution structure of Rv1761c with a schematic representation of the influence range of the spin label. (E, F) Schematic illustration of how specifically spin‐labeled detergents can be used to probe the insertion depth in the membrane. (E) Structure 1: 5‐DSA, Structure 2: 16‐DSA. Depending on the position of the spin label within the chemical structure of the detergent, different zones within the transmembrane region will be affected as highlighted by the red rings in (F). Thus, nuclear magnetic resonance (NMR) signals from residues occupying the upper layer are influenced more by 5‐DSA than by 16‐DSA and vice versa.

Parts (A)‐(C) reproduced, with permission, from reference 193. Copyright 2009 John Wiley and Sons.
Figure 9. Figure 9.

Residual dipolar coupling (RDC) and its application in membrane protein structure characterization. (A) Under isotropic conditions the protein is free to tumble in solution and the dipolar coupling is averaged to 0. In this case the splitting in the nuclear magnetic resonance (NMR) signal corresponds to the J‐coupling. (B) In an anisotropic scenario, where submolar fractions of a suitable alignment media has been added to the NMR sample, the splitting of the two peaks changes as a function of the sum of the J‐coupling and the residual dipolar coupling D. (C) With increased alignment the residual dipolar coupling increases and since the sign of the dipolar coupling may be both positive and negative, this changes the splitting in the NMR signal correspondingly. The numbers above the NMR signals are the 1JN‐HN in (A) and the sum of 1JN‐HN and 1DN‐HN in both (B) and (C). The numbers in parentheses are thus (D). All numbers are given in Hz. (D) Refinement of the OmpA structure using RDCs. By using RDC derived restraints, the structure of OmpA improved from a precision of 1.82 ± 0.22 Å, accuracy 2.94 ± 0.18 Å to a precision of 0.90 ± 0.20 Å, accuracy 1.92± 0.15 Å in the absence of additional hydrogen bond restraints and from precision of 1.10 ± 0.15 Å, accuracy 1.66 ± 0.11 Å to a precision of 0.48 ± 0.05 Å, and accuracy 1.02 ± 0.052 Å in the presence of additional hydrogen‐bond restraints. The crystal structure of OmpA is underlying in green, TM regions are shown in blue, and periplasmic loops in red. (E) FXYD1 has a TM helix and an EM helix. Using RDC‐derived restraints, four possible orientations of the EM relative to the TM helix were possible. With the aid of paramagnetic ions probing for solvent exposure, orientation Nr. 1 could be selected. Here, the gradient in red represents the reduction of the signal in the presence of paramagnetic Mn2+ ions in solution and the gray bare the membrane.

Parts (D) reprinted, with permission, from reference 51 and (E) from reference 246. Copyright 2006 and 2007 American Chemical Society.
Figure 10. Figure 10.

Schematic illustration of conformational events in membrane proteins. The motions that change a protein's conformation occur with time‐constants from ps to several minutes, indicated by the red arrow. The fastest motions, vibrations and librations around covalent bonds, result in atoms moving only a fraction of an Ångström. Ligand binding may involve only subtle motions like rearrangement of amino‐acid residues in the binding site or larger movements over several Ångströms‐like domain reorientations. Here, the NMR structure of the extracelluar domain of the 7 TM human corticotrophin‐releasing factor receptor 1 (in blue) complexed with a high affinity agonist in red (PDB code 2L27) 93. Similar movements may be observed for allosteric structural changes where binding of a ligand at one site of the protein changes the structure in another part. This is shown here for phospholamban, where assembly to the pentamer changes the helix curvature (PDB code 2KB7 and 2K0L) 192,248. Lastly, protein folding involves large movements of the whole protein chain as illustrated hypothetical here for the TM domain of the N‐actylcholine receptor B2‐subunit (PDB code 2KSR) 28. Images of structures generated in PyMol (DeLano Scientific).

Figure 11. Figure 11.

Fast time scale dynamics in protein measured by nuclear magnetic resonance (NMR) spectroscopy. (A) The relaxation times T1 (spin‐lattice relaxation time) and T2 (spin‐spin relaxation time) are among other things strongly dependent on the rotational correlation time, τc, of the molecule. Thus for large molecules, the slower it tumbles in solution, the more dominant the influence of the relaxation on the NMR signal becomes. (B) The order parameter S2 is derived from the various different relaxation parameters and yields a measure for the dependent local mobility of the NMR probe. One qualitative way to describe the order parameter is to describe the motion the bond vector as a cone with an angular dependence θ (C). The smaller this cone, the larger the generalized order parameter S2. Thus for fully rigid structures, S2 will be 1, and for completely random structures S2 will be 0. Often S2 is decomposed into two separate contributions, S2f, which is the fast component representing the ps dynamics and S2s the slow component representing the ns dynamics. Membranes influence IMP dynamics. The water‐soluble analogue WSK3 to the potassium channel KcsA was analyzed by [1H, 15N]—relaxation parameters at 14.1 T(black) and 16.5 T (red). The experimental data are shown in panel (D). These were fitted to the extended model free approach yielding Sf2, SS2, and τs shown in (E). The water‐soluble WSK3 shows dramatically increased dynamics in the slow time regime as represented by SS2 when compared to the original values for the dynamics of KcsA. Especially the pore helix appears to be affected as highlighted in red in (F) (top: KcsA, bottom: WSK3).

Figures (D) and (E) reproduced, with permission, from reference 159. Copyright 2008 National Academy of Sciences USA.
Figure 12. Figure 12.

The influence of chemical exchange on the nuclear magnetic resonance (NMR) signal. Slow chemical exchange results in two distinct signals to be recorded (A top). Chemical exchange on the millisecond (ms) to microsecond (μs) timescale will result in line broadening of the NMR signal which in the extreme case results in line broadening beyond detection (A middle). For chemical exchange in the fast regime, an averaged signal will be recorded (A bottom). In the last case, the position of the signal will depend on the population ratio of the interconverting species and the interconversion rate as illustrated in (B) (C) Chemical exchange also contributes to an apparent faster loss of the signal (R2 + Rex) than can be attributed to the relaxation process described by R2 alone. (D) This loss can be partially recovered by applying a train of refocusing pulses, a so‐called refocusing CPMG pulse sequence in the preparation time. (E) The combined relaxation rate as a function of the frequency of the CPMG sequence. The CPMG will thus remove the contribution from Rex to R2and thereby yield information on the underlying chemical exchange rate Rex. (F) PagP exists in a relaxed (R) and a tensed (T) state, and shows two different NMR signals at 25°C. These two sets of peaks show distinct dynamic differences on the intermediate timescale as highlighted for the residues Y38 and Q139 in (G). (H) Chemical shift differences between the R and the T states color coded onto the cartoon structure of PagP (PDB code 1MM4). (I) The ensemble of the 20 lowest energy structures of PagP highlighting the flexibility of the EM parts.

Parts (F)‐(H) reproduced, with permission, from reference 110. Copyright 2004 National Academy of Sciences USA
Figure 13. Figure 13.

Amide hydrogen to deuterium exchange measured by solution nuclear magnetic resonance (NMR) spectroscopy. (A) Exchange of amides in a membrane protein requires access to bulk solvent, either directly or through conformational changes. An amide proton exchanges with an intrinsic rate constant dependent on the chemical structure of the peptide chain, kint, and with an observed exchange rate constant, kobs, which depends on the structural environment. (B) Since the deuteron has different NMR properties than the proton, it will not be detected in a proton NMR spectrum and as the amide at a certain position is exchanged with a deuteron, the signal gradually disappears following an exponential decay (D). In the lower square (E), two different amides with different rates of exchange are shown, one with fast (red) and one with slower (black) exchange. Each data point represents the intensity of an NMR peak in an HSQC spectrum. (C) A protection factor P can be described as the ratio of the intrinsic to the observed rate constant. The more protected the amide, the larger the protection factor. A log scaling is typically used to illustrate P. For membrane proteins an equilibrium factor is often used, ɛeq, which is the intensity ratio of the NMR signal in H2O to that of the NMR signal in a given amount of D2O. Here, the larger the equilibrium factor, the less protected the amide is. (D) Equilibrium factors used to describe the water accessibility of OmpX. The equilibrium exchange factor ɛeq represents the ratio of peak volumes for OmpX/APol complexes incubated in 100 versus 5% D2O. The ɛeq is normalized relative to the ratio observed for the most protected residue. Here diversity in exchange factors was seen within the barrel, with some strands in the middle of the membrane being most protected while other strands show increased exchange (lower) in the membrane center. This suggesting that dynamics in the barrel is important for the function of OmpX 39.

Part (F) reproduced, with permission, from reference 39. Copyright 2009 Springer.
Figure 14. Figure 14.

Single‐pass IMP structures solved by solution nuclear magnetic resonance (NMR) spectroscopy. (A) Glycoprotein E1 from hepatitis C (PDB code 1EMZ) 186; in red is shown three glycine residues, (B) Sarcolipin, a regulator of Ca‐ATPase (PDB code 1JDM) 165; in red is shown the only three polar residues, (C) KCNE1 a modulator of the KCNQ1 potassium channel (PDB code 2K21) 123, (D) The tetrameric M2‐channel from influenza A virus (PDB code 2RLF) 219; shown in red are four tryptophans and in green above four histidines and in green below four aspartic acids, (E) EphA1 (PDB code 1KIL) 23, in red is shown the tandem glycine motif, and in green a leucine zipper motif facing the lipids, (F) EphA2 (PDB code 2K9Y) 24, in red the extensive hydrophobic interface ending with a Phe‐stack, (G) neu/ErbB2 monomer (PDB code 1IIJ) 91, (H) ErbB2 homodimer (PDB code 2JWA) 25; in red the Ser/Thr‐node, in green valines and leucines as well as valines in blue sticks facing the lipids, (I) ErbB1:ErbB2 heterodimer (PDB code 2KS1) 175; in red the GXXXG:Ser/Thr packing, (J) αIIbβ3‐heterodimer (PDB code 2KNC) 281, in red glycines and in green a tryptophan, (K) Glycophorin A (PDB code1AFO) 160, in red the glycines and valines of the glycine zipper motif, (L) BNip3 homodimer (PDB code 2J5D) 26, in red at the top the Phe‐cluster, at bottom the GXXXG‐motif, in green the Ser‐His node, (M) ζζ‐homodimer (PDB code 2HAC) 37, in red the Tyr/Thr network, in green the Asp‐lock, (N) DAP12 homodimer (PDB code 2L34) 38, in red at the top the GXXXG‐motif, in bottom the threonines, in green the Asp‐lock, and (O) DAP12:NGK2C heterotrimer (PDB code 2L35) 38. In red the NGK2C helix, in green the Lys:Asp network. The three monomer structures were solved in TFE or sodium dodecyl sulphate (SDS). Images of structures generated in PyMol (DeLano Scientific).

Figure 15. Figure 15.

Fragments of IMPs can be fair structural representatives. The entire sequence of bacteriorhodopsin was synthesized as short, overlapping peptides, and their structures solved by solution NMR spectroscopy. These structures were compared to the crystal structure and showed surprising similarity throughout, also what concerned turn structures.

The figure is reproduced from the Biophysical Journal 124, with permission, from Elsevier.
Figure 16. Figure 16.

Interacellular juxtamembrane structures of integral membrane proteins (IMPs) solved by solution nuclear magnetic resonance (NMR) spectroscopy. (A) For some intracellular domains of IMPs, an equilibrium between a solvent resident, disordered state and a membrane embedded, helical state is suggested and its importance in receptor regulation presently discussed. (B) Structure of the αIIbβ3‐complex (PDB code 1M8O) 264 and those of the monomeric states of αIIb (PDB code 1DPK) and of β3 (pdb code 1S4X) 263, as well as of an extended version of β3 (PDB code 2KV9) 173. (C) The ICD of the epidermal growth receptor (EGF)‐receptor forms a similar EM structure as β3 (PDB code 1Z9I) 48. (D) Structure of human CD4mut with the TM helix and part of the ICD (PDB code 2KLU) 278. (E) The phospholamban monomer (PDB code 2KB7) 248. (F) The ICD of the T‐cell receptor CD3ɛ with two ITAM motifs (PDB code 2K4F) 280; the tyrosine and leucines are shown in sticks. (G) Mec4 (PDB code 2K2B). JM part of viruses and bacteriophages proteins (H) HIV‐1 virus protein U (VpU) with tertiary contact between EM helices (PDB code 2K7Y) 276, (I) Hepatitis C glycoprotein E1, HepC E1 (PDB code 2KNU) 234, (J) fd coat protein from filamentous bacteriophages (pdb code 1FDM) 4, and (K) tryptophan‐rich domain of HIV glycoprotein gp41 (PDB code 1JAV) 218. Images of structures generated in PyMol (DeLano Scientific).

Figure 17. Figure 17.

Structure‐activity relationship (SAR) by nuclear magnetic resonance (NMR). (A) Cartoon describing the pipeline of novel lead candidate generation. Small probes are used to screen the binding site and are subsequently combined to form the new lead compound. (B) A summary of the approach to find novel p‐arylthio cinnamides as lead candidates for the leukocyte function associated antigen 1 (LFA‐1) 152. The different chemical compounds identified in the various screening steps are drawn as chemical structures under the respective screen steps.

Figure 18. Figure 18.

Nuclear magnetic resonance (NMR) shift mapping—a tool to locate binding sites and interactions. Chemical shift perturbations of CR56 and RAPd1 by complex formation. (A) HSQC spectra of free (black peaks) and CR56‐bound (red) RAPd1. (B) Combined chemical shift changes ((|ΔδH| + 1/4|ΔδN|)0.5) are shown as bars versus residue number for 15N‐CR56 upon binding of unlabeled RAPd1. Proline residues (+) and nonassigned residues in the bound form (x) are indicated. The CR5 and CR6 domain boundaries are indicated with bars above the plots. (C) Surface plots showing in red those residues with significantly perturbed chemical shifts upon complex formation in CR56 (Ala944, Cys948, Trp953, Thr954, Asp956, Asp966, Ser968, Ser970, Ala977, Thr975, Phe977, Thr980, Gly987, Asp997, Asn998, and Gly1003). (D) Sections of 15N‐1H HSQC spectra showing the peaks from RAPd1 residue Glu40 during the titration of 15N‐labeled RAPd1 with increasing amounts of unlabeled CR56. Figure modified, with permission, from reference 120 and with permission from Elsevier.

Figure 19. Figure 19.

Ligand screening methods. A summary of the various NMR methods used to screen for ligands with moderately binding affinity, highlighting subtle differences in the modes of action. Paramagnetic relaxation enhancement (PRE): a label is attached to the macromolecule close to the ligand binding site and will modify the relaxation properties of nearby residues, including atoms of a weakly interacting ligand. Diffusion‐ordered spectroscopy (DOSY): the interaction of the ligand with the protein will result into slower diffusion properties of the ligand than expected from its actual size, resulting in modification of the NMR parameters of the ligand. Saturation transfer difference (STD) and trNOE: saturation will be transferred from the macromolecule to the ligand. The efficiency of this transfer will depend on the actual orientation of the ligand when bound to the protein. Interligand nuclear Overhauser effect (ILOE): the interligand NOE makes use of the strong negative ligand‐ligand NOE. The NOEs can be used to identify those different ligands that bind simultaneously and can help in the design of novel lead candidates based on the initial two weak binders and their orientation relative to each other.



Figure 1.

Structures of 18 selected membrane protein solved by solution NMR spectroscopy. Since 2001, only around 20 structures of integral membrane proteins have been solved in solution and are publicly available. The figure has been divided into two parts, β‐barrel structures in the upper part and α‐helical structures in the lower part. The structures are (A) OmpX, PDB code 1Q9F 78, (B) OmpA, PDB code 2GEA 7, (C) PagP, PDB code 1MM4 111, (D) OmpG, PDB code 2JQY 75, (E) human VDAC‐1, PDB code 2JK4 102, (F) KpOmpA, PDB code 2K0L 119,205, (G) potassium channel KcsA, PDB code 2A9H 10, (H) diacylglycerol kinase, DAGK, PDB code 2KDC 259, (I) phospholamban, PDB code 1ZLL 192, (J) influenza A M2 proton channel, PDB code 2RLF 219, (K) influenza B M2 proton channel, PDB code 2KIX 267, (L) DsbB, PDB code 2K73 289, (M) sensory rhodopsin II, PDB code 2KSY 86, (N) n‐acetylcholine receptor, PDB code 2KSR 28, (O) KCNE1, PDB code 2K21 123, (P) FXYD1, PDB code 2JO1 246, (Q) Rv1761c, PDB code 2K3M 193, and (R) subunit c of the F0F1 ATPase synthase, PDB code 1A91 80. Images of structures generated in PyMol (DeLano Scientific).



Figure 2.

Choice of lipids for NMR characterization. At the top, a schematic illustration of membrane protein extraction and reconstitution into detergents suitable for structural analysis is shown. The detergent needs to preserve the native fold of the internal membrane part. However, some detergents will also affect external parts and even denature those, which is detrimental for structure analysis. Later, an overview of various commonly used detergents applied in solution nuclear magnetic resonance (NMR) studies of IMPs is provided. The chemical structures are grouped according to their nature into organic solvents, nonionic, ionic, zwitterionic, and sterol‐based detergents. For each chemical structure, the critical micelle concentration (cmc) and the aggregation number (AG number) are listed, which are taken either from 215 or from the suppliers.



Figure 3.

The NMR assignment process. Typically by selecting different types of NMR spectra, one selects intra‐residue signals versus inter‐residue signals in a 3D‐spectrum. By a combination of such spectra, which often is nine or more 3D spectra, and together with the knowledge of the sequence of the protein, one performs the job of linking the chemical shifts to distinct nuclei. This process is called assignment. In the assignment process, the NMR signals are linked to the correct nuclei in the protein. (A) A 1D1H‐NMR spectrum of a protein with the chemical shifts of the NMR signals on the x‐axis, approximately going from −1 ppm to 11 ppm. The large overlap between signals demands for extraction into more dimensions. (B) A 2D15N,1H‐HSQC spectrum of a protein with NMR signals from all protons covalently bound to a nitrogen. Here the overlap is minimized dramatically. (C) A 3D HNCO‐spectrum of a protein showing NMR signals from amide protons covalently bound to nitrogen, which again is covalently bound to a carbonyl atom. This type of spectrum requires stable isotope labeling with 15N and 13C. (D) A schematic description of the assignment process. From knowledge of the chemical structure of the peptide chain and the sequence of the protein (E), one uses a set of six to ten 3D heteronuclear NMR spectra to assign the individual NMR signals to the individual atoms of the protein. (F) A schematic representation of the sequential walk, where signals from six different strips of a theoretical HNCACB 3D spectrum are shown. One strip represents carbon signals correlating to a single N,HN. In each strip the N,HN correlates to four carbons, the Cα and Cβ from its own residue and the Cα and Cβ from the preceding residue. This directly links the strips (see arrows). The challenge in assignment is to identify which strip, i.e. which residue is a neighbor to which. (G) An example of the magnetization transfer steps exploited to generate two of the triple‐resonance spectra HNCACB (left) and CBCA(CO)NH (right) important for the assignment. The fist spectrum (left) provides both intraresidual and sequential signals and the second spectrum (right) selects only sequential signals. Arrows indicate the route of transfer and the number the order by which magnetization is transferred from one nucleus to the next. The connections are between residue i and the preceding residue i − 1. All carbons are 13C‐labeled and all nitrogens 15N‐labeled.



Figure 4.

HSQC spectra with and without TROSY‐a tool to circumvent the size problem. Comparison of (A) a conventional 2D [1H, 15N]‐HSQC and (B) a 2D [1H, 15N]‐TROSY spectrum. Both spectra were recorded on a 2 mmol/L OmpX solution in DHPC micelles at 30°C at a resonance frequency of 750 Mhz. The superior performance of the TROSY experiment clearly manifests in the better resolution of the individual cross peaks in the TROSY spectrum. (C) Solution nuclear magnetic resonance (NMR) structure of OmpX in DHPC micelles [20, PDB‐code 1Q9F]. The TROSY effect is largest at very strong magnetic fields. Parts (A) and (B) are modified, with permission, from reference 77. Copyright 2001 National Academy of Sciences USA.



Figure 5.

NMR and the size problem. (A) Molecules tumble freely in solution. However, larger proteins take longer time to tumble around their own axis than small molecules. Left: illustration of a large molecule with slow tumbling, and right: a small molecule with fast tumbling. (B) A nuclear magnetic resonance (NMR) signal is recorded after the application of a radiofrequency pulse that perturbs the equilibrium distribution of spins. One records the return to equilibrium that is an exponential process and is called the free‐induction decay, FID (in red waves). To obtain the NMR spectrum, the FID undergoes a Fourier transformation (FT)—a mathematical operation. For large molecules tumbling more slowly (left), the corresponding NMR signal disappears (relaxes) quickly and the intensity of the signal is dampened, whereas for small molecules (right), the NMR signal is more long‐lived, giving rise to more intense and sharper signals.



Figure 6.

Typical NMR‐derived structural restraints. A number of different conformational restraints can be derived from different types of nuclear magnetic resonance (NMR) experiments. Chemical shifts‐derived dihedral angles restraints: when chemical shifts are assigned to the nuclei, the values are compared to those of databases that link chemical shifts to structures. Hereby one obtains the likelihood for the backbone dihedral angles φ and ψ to be in specific conformations, that is, occupying specific regions of the Ramachandran plot. The dihedral angles are then included together with an allowed angle deviation as input to the structure calculation. NOE‐derived distance restraints: through cross‐relaxation through space, one obtains a distance dependence of the intensity of the NMR signal that declines with the six‐root of the distance. This is termed the NOE effect. This means that only short‐range distances (r), can be detected, typically up to 5 Å. Thus, the intensity of the NOE signal determines the range of the distance restraint, where a strong peak typically is interpreted as a distance being closer than 2.5 Å and weak peak as a distance closer than 5 Å. RDC‐derived global restraints: residual dipolar couplings (RDCs) are global restraints that relate the bond vector to the laboratory frame. From a change in the splitting of the NMR signal, either to larger or smaller splitting, the angles relating the N‐H vector to the laboratory frame can be determined for each vector of the molecule. PRE‐derived distance restraints: paramagnetic relaxation enhancements (PREs) are long‐range distance restraints. The presence of a paramagnetic center enhances the relaxation process and thus lowers the intensity of the NMR signals in its surroundings. For the three cases described here, atom a is closest to the paramagnetic center and thus its NMR signal will be most affected, whereas atom c is too far away to be affected and b is in between.



Figure 7.

Structure calculation scheme. Flowchart describing a possible procedure for protein structure determination by solution nuclear magnetic resonance (NMR) spectroscopy. After protein production and purification a large set of triple resonance, heteronuclear three‐dimensional (3D) NMR spectra are recorded. The chemical shifts are assigned and restraints are derived from chemical shifts, NOEs, PREs, RDCs, etc. Together with information of the chemical structure and sequence, these restraints are fed into a computer algorithm that iteratively calculates and refines the structure and finally yields an ensemble of structures such as shown here for OmpX, fulfilling the experimental restraints. The red arrow indicated the iterative cycle of refinements, which often result in numerous cycles.



Figure 8.

Paramagnetic relaxation enhancement (PRE) and its application in membrane protein structure characterization. PREs can be used to measure long‐range distance restraints in proteins as illustrated in (A) or determine solvent/detergent exposure as illustrated in (B). The 2D [1H, 15N]—HSQC spectrum of Rv1761c in the presence of a (A) reduced and an (B) oxidized MTSL‐label in position F30C. The circles in (B) highlight the residues whose signals are missing in the oxidized spectrum. A more quantitative illustration showing the intensity ratio of reduced (para) and oxidized (dia) cross peaks (Ired/Iox) is shown in (C). An asterisk marks the position of the spin label. From this pattern, additional structural restraints could be deduced. The inset shows the solution structure of Rv1761c with a schematic representation of the influence range of the spin label. (E, F) Schematic illustration of how specifically spin‐labeled detergents can be used to probe the insertion depth in the membrane. (E) Structure 1: 5‐DSA, Structure 2: 16‐DSA. Depending on the position of the spin label within the chemical structure of the detergent, different zones within the transmembrane region will be affected as highlighted by the red rings in (F). Thus, nuclear magnetic resonance (NMR) signals from residues occupying the upper layer are influenced more by 5‐DSA than by 16‐DSA and vice versa.

Parts (A)‐(C) reproduced, with permission, from reference 193. Copyright 2009 John Wiley and Sons.


Figure 9.

Residual dipolar coupling (RDC) and its application in membrane protein structure characterization. (A) Under isotropic conditions the protein is free to tumble in solution and the dipolar coupling is averaged to 0. In this case the splitting in the nuclear magnetic resonance (NMR) signal corresponds to the J‐coupling. (B) In an anisotropic scenario, where submolar fractions of a suitable alignment media has been added to the NMR sample, the splitting of the two peaks changes as a function of the sum of the J‐coupling and the residual dipolar coupling D. (C) With increased alignment the residual dipolar coupling increases and since the sign of the dipolar coupling may be both positive and negative, this changes the splitting in the NMR signal correspondingly. The numbers above the NMR signals are the 1JN‐HN in (A) and the sum of 1JN‐HN and 1DN‐HN in both (B) and (C). The numbers in parentheses are thus (D). All numbers are given in Hz. (D) Refinement of the OmpA structure using RDCs. By using RDC derived restraints, the structure of OmpA improved from a precision of 1.82 ± 0.22 Å, accuracy 2.94 ± 0.18 Å to a precision of 0.90 ± 0.20 Å, accuracy 1.92± 0.15 Å in the absence of additional hydrogen bond restraints and from precision of 1.10 ± 0.15 Å, accuracy 1.66 ± 0.11 Å to a precision of 0.48 ± 0.05 Å, and accuracy 1.02 ± 0.052 Å in the presence of additional hydrogen‐bond restraints. The crystal structure of OmpA is underlying in green, TM regions are shown in blue, and periplasmic loops in red. (E) FXYD1 has a TM helix and an EM helix. Using RDC‐derived restraints, four possible orientations of the EM relative to the TM helix were possible. With the aid of paramagnetic ions probing for solvent exposure, orientation Nr. 1 could be selected. Here, the gradient in red represents the reduction of the signal in the presence of paramagnetic Mn2+ ions in solution and the gray bare the membrane.

Parts (D) reprinted, with permission, from reference 51 and (E) from reference 246. Copyright 2006 and 2007 American Chemical Society.


Figure 10.

Schematic illustration of conformational events in membrane proteins. The motions that change a protein's conformation occur with time‐constants from ps to several minutes, indicated by the red arrow. The fastest motions, vibrations and librations around covalent bonds, result in atoms moving only a fraction of an Ångström. Ligand binding may involve only subtle motions like rearrangement of amino‐acid residues in the binding site or larger movements over several Ångströms‐like domain reorientations. Here, the NMR structure of the extracelluar domain of the 7 TM human corticotrophin‐releasing factor receptor 1 (in blue) complexed with a high affinity agonist in red (PDB code 2L27) 93. Similar movements may be observed for allosteric structural changes where binding of a ligand at one site of the protein changes the structure in another part. This is shown here for phospholamban, where assembly to the pentamer changes the helix curvature (PDB code 2KB7 and 2K0L) 192,248. Lastly, protein folding involves large movements of the whole protein chain as illustrated hypothetical here for the TM domain of the N‐actylcholine receptor B2‐subunit (PDB code 2KSR) 28. Images of structures generated in PyMol (DeLano Scientific).



Figure 11.

Fast time scale dynamics in protein measured by nuclear magnetic resonance (NMR) spectroscopy. (A) The relaxation times T1 (spin‐lattice relaxation time) and T2 (spin‐spin relaxation time) are among other things strongly dependent on the rotational correlation time, τc, of the molecule. Thus for large molecules, the slower it tumbles in solution, the more dominant the influence of the relaxation on the NMR signal becomes. (B) The order parameter S2 is derived from the various different relaxation parameters and yields a measure for the dependent local mobility of the NMR probe. One qualitative way to describe the order parameter is to describe the motion the bond vector as a cone with an angular dependence θ (C). The smaller this cone, the larger the generalized order parameter S2. Thus for fully rigid structures, S2 will be 1, and for completely random structures S2 will be 0. Often S2 is decomposed into two separate contributions, S2f, which is the fast component representing the ps dynamics and S2s the slow component representing the ns dynamics. Membranes influence IMP dynamics. The water‐soluble analogue WSK3 to the potassium channel KcsA was analyzed by [1H, 15N]—relaxation parameters at 14.1 T(black) and 16.5 T (red). The experimental data are shown in panel (D). These were fitted to the extended model free approach yielding Sf2, SS2, and τs shown in (E). The water‐soluble WSK3 shows dramatically increased dynamics in the slow time regime as represented by SS2 when compared to the original values for the dynamics of KcsA. Especially the pore helix appears to be affected as highlighted in red in (F) (top: KcsA, bottom: WSK3).

Figures (D) and (E) reproduced, with permission, from reference 159. Copyright 2008 National Academy of Sciences USA.


Figure 12.

The influence of chemical exchange on the nuclear magnetic resonance (NMR) signal. Slow chemical exchange results in two distinct signals to be recorded (A top). Chemical exchange on the millisecond (ms) to microsecond (μs) timescale will result in line broadening of the NMR signal which in the extreme case results in line broadening beyond detection (A middle). For chemical exchange in the fast regime, an averaged signal will be recorded (A bottom). In the last case, the position of the signal will depend on the population ratio of the interconverting species and the interconversion rate as illustrated in (B) (C) Chemical exchange also contributes to an apparent faster loss of the signal (R2 + Rex) than can be attributed to the relaxation process described by R2 alone. (D) This loss can be partially recovered by applying a train of refocusing pulses, a so‐called refocusing CPMG pulse sequence in the preparation time. (E) The combined relaxation rate as a function of the frequency of the CPMG sequence. The CPMG will thus remove the contribution from Rex to R2and thereby yield information on the underlying chemical exchange rate Rex. (F) PagP exists in a relaxed (R) and a tensed (T) state, and shows two different NMR signals at 25°C. These two sets of peaks show distinct dynamic differences on the intermediate timescale as highlighted for the residues Y38 and Q139 in (G). (H) Chemical shift differences between the R and the T states color coded onto the cartoon structure of PagP (PDB code 1MM4). (I) The ensemble of the 20 lowest energy structures of PagP highlighting the flexibility of the EM parts.

Parts (F)‐(H) reproduced, with permission, from reference 110. Copyright 2004 National Academy of Sciences USA


Figure 13.

Amide hydrogen to deuterium exchange measured by solution nuclear magnetic resonance (NMR) spectroscopy. (A) Exchange of amides in a membrane protein requires access to bulk solvent, either directly or through conformational changes. An amide proton exchanges with an intrinsic rate constant dependent on the chemical structure of the peptide chain, kint, and with an observed exchange rate constant, kobs, which depends on the structural environment. (B) Since the deuteron has different NMR properties than the proton, it will not be detected in a proton NMR spectrum and as the amide at a certain position is exchanged with a deuteron, the signal gradually disappears following an exponential decay (D). In the lower square (E), two different amides with different rates of exchange are shown, one with fast (red) and one with slower (black) exchange. Each data point represents the intensity of an NMR peak in an HSQC spectrum. (C) A protection factor P can be described as the ratio of the intrinsic to the observed rate constant. The more protected the amide, the larger the protection factor. A log scaling is typically used to illustrate P. For membrane proteins an equilibrium factor is often used, ɛeq, which is the intensity ratio of the NMR signal in H2O to that of the NMR signal in a given amount of D2O. Here, the larger the equilibrium factor, the less protected the amide is. (D) Equilibrium factors used to describe the water accessibility of OmpX. The equilibrium exchange factor ɛeq represents the ratio of peak volumes for OmpX/APol complexes incubated in 100 versus 5% D2O. The ɛeq is normalized relative to the ratio observed for the most protected residue. Here diversity in exchange factors was seen within the barrel, with some strands in the middle of the membrane being most protected while other strands show increased exchange (lower) in the membrane center. This suggesting that dynamics in the barrel is important for the function of OmpX 39.

Part (F) reproduced, with permission, from reference 39. Copyright 2009 Springer.


Figure 14.

Single‐pass IMP structures solved by solution nuclear magnetic resonance (NMR) spectroscopy. (A) Glycoprotein E1 from hepatitis C (PDB code 1EMZ) 186; in red is shown three glycine residues, (B) Sarcolipin, a regulator of Ca‐ATPase (PDB code 1JDM) 165; in red is shown the only three polar residues, (C) KCNE1 a modulator of the KCNQ1 potassium channel (PDB code 2K21) 123, (D) The tetrameric M2‐channel from influenza A virus (PDB code 2RLF) 219; shown in red are four tryptophans and in green above four histidines and in green below four aspartic acids, (E) EphA1 (PDB code 1KIL) 23, in red is shown the tandem glycine motif, and in green a leucine zipper motif facing the lipids, (F) EphA2 (PDB code 2K9Y) 24, in red the extensive hydrophobic interface ending with a Phe‐stack, (G) neu/ErbB2 monomer (PDB code 1IIJ) 91, (H) ErbB2 homodimer (PDB code 2JWA) 25; in red the Ser/Thr‐node, in green valines and leucines as well as valines in blue sticks facing the lipids, (I) ErbB1:ErbB2 heterodimer (PDB code 2KS1) 175; in red the GXXXG:Ser/Thr packing, (J) αIIbβ3‐heterodimer (PDB code 2KNC) 281, in red glycines and in green a tryptophan, (K) Glycophorin A (PDB code1AFO) 160, in red the glycines and valines of the glycine zipper motif, (L) BNip3 homodimer (PDB code 2J5D) 26, in red at the top the Phe‐cluster, at bottom the GXXXG‐motif, in green the Ser‐His node, (M) ζζ‐homodimer (PDB code 2HAC) 37, in red the Tyr/Thr network, in green the Asp‐lock, (N) DAP12 homodimer (PDB code 2L34) 38, in red at the top the GXXXG‐motif, in bottom the threonines, in green the Asp‐lock, and (O) DAP12:NGK2C heterotrimer (PDB code 2L35) 38. In red the NGK2C helix, in green the Lys:Asp network. The three monomer structures were solved in TFE or sodium dodecyl sulphate (SDS). Images of structures generated in PyMol (DeLano Scientific).



Figure 15.

Fragments of IMPs can be fair structural representatives. The entire sequence of bacteriorhodopsin was synthesized as short, overlapping peptides, and their structures solved by solution NMR spectroscopy. These structures were compared to the crystal structure and showed surprising similarity throughout, also what concerned turn structures.

The figure is reproduced from the Biophysical Journal 124, with permission, from Elsevier.


Figure 16.

Interacellular juxtamembrane structures of integral membrane proteins (IMPs) solved by solution nuclear magnetic resonance (NMR) spectroscopy. (A) For some intracellular domains of IMPs, an equilibrium between a solvent resident, disordered state and a membrane embedded, helical state is suggested and its importance in receptor regulation presently discussed. (B) Structure of the αIIbβ3‐complex (PDB code 1M8O) 264 and those of the monomeric states of αIIb (PDB code 1DPK) and of β3 (pdb code 1S4X) 263, as well as of an extended version of β3 (PDB code 2KV9) 173. (C) The ICD of the epidermal growth receptor (EGF)‐receptor forms a similar EM structure as β3 (PDB code 1Z9I) 48. (D) Structure of human CD4mut with the TM helix and part of the ICD (PDB code 2KLU) 278. (E) The phospholamban monomer (PDB code 2KB7) 248. (F) The ICD of the T‐cell receptor CD3ɛ with two ITAM motifs (PDB code 2K4F) 280; the tyrosine and leucines are shown in sticks. (G) Mec4 (PDB code 2K2B). JM part of viruses and bacteriophages proteins (H) HIV‐1 virus protein U (VpU) with tertiary contact between EM helices (PDB code 2K7Y) 276, (I) Hepatitis C glycoprotein E1, HepC E1 (PDB code 2KNU) 234, (J) fd coat protein from filamentous bacteriophages (pdb code 1FDM) 4, and (K) tryptophan‐rich domain of HIV glycoprotein gp41 (PDB code 1JAV) 218. Images of structures generated in PyMol (DeLano Scientific).



Figure 17.

Structure‐activity relationship (SAR) by nuclear magnetic resonance (NMR). (A) Cartoon describing the pipeline of novel lead candidate generation. Small probes are used to screen the binding site and are subsequently combined to form the new lead compound. (B) A summary of the approach to find novel p‐arylthio cinnamides as lead candidates for the leukocyte function associated antigen 1 (LFA‐1) 152. The different chemical compounds identified in the various screening steps are drawn as chemical structures under the respective screen steps.



Figure 18.

Nuclear magnetic resonance (NMR) shift mapping—a tool to locate binding sites and interactions. Chemical shift perturbations of CR56 and RAPd1 by complex formation. (A) HSQC spectra of free (black peaks) and CR56‐bound (red) RAPd1. (B) Combined chemical shift changes ((|ΔδH| + 1/4|ΔδN|)0.5) are shown as bars versus residue number for 15N‐CR56 upon binding of unlabeled RAPd1. Proline residues (+) and nonassigned residues in the bound form (x) are indicated. The CR5 and CR6 domain boundaries are indicated with bars above the plots. (C) Surface plots showing in red those residues with significantly perturbed chemical shifts upon complex formation in CR56 (Ala944, Cys948, Trp953, Thr954, Asp956, Asp966, Ser968, Ser970, Ala977, Thr975, Phe977, Thr980, Gly987, Asp997, Asn998, and Gly1003). (D) Sections of 15N‐1H HSQC spectra showing the peaks from RAPd1 residue Glu40 during the titration of 15N‐labeled RAPd1 with increasing amounts of unlabeled CR56. Figure modified, with permission, from reference 120 and with permission from Elsevier.



Figure 19.

Ligand screening methods. A summary of the various NMR methods used to screen for ligands with moderately binding affinity, highlighting subtle differences in the modes of action. Paramagnetic relaxation enhancement (PRE): a label is attached to the macromolecule close to the ligand binding site and will modify the relaxation properties of nearby residues, including atoms of a weakly interacting ligand. Diffusion‐ordered spectroscopy (DOSY): the interaction of the ligand with the protein will result into slower diffusion properties of the ligand than expected from its actual size, resulting in modification of the NMR parameters of the ligand. Saturation transfer difference (STD) and trNOE: saturation will be transferred from the macromolecule to the ligand. The efficiency of this transfer will depend on the actual orientation of the ligand when bound to the protein. Interligand nuclear Overhauser effect (ILOE): the interligand NOE makes use of the strong negative ligand‐ligand NOE. The NOEs can be used to identify those different ligands that bind simultaneously and can help in the design of novel lead candidates based on the initial two weak binders and their orientation relative to each other.

References
 1. Aisenbrey C, Bertani P, Henklein P, Bechinger B. Structure, dynamics and topology of membrane polypeptides by oriented 2H solid‐state NMR spectroscopy. Eur Biophys J 36: 451‐460, 2007.
 2. Aivazian D, Stern LJ. Phosphorylation of T cell receptor zeta is regulated by a lipid dependent folding transition. Nat Struct Biol 7: 1023‐1026, 2000.
 3. Alipanahi B, Gao X, Karakoc E, Li SC, Balbach F, Feng G, Donaldson L, Li M. Error tolerant NMR backbone resonance assignment and automated structure generation. J Bioinf Comp Biol 9: 15‐41, 2011.
 4. Almeida FC, Opella SJ. fd coat protein structure in membrane environments: Structural dynamics of the loop between the hydrophobic trans‐membrane helix and the amphipathic in‐plane helix. J Mol Biol 270: 481‐495, 1997.
 5. Altenbach C, Marti T, Khorana HG, Hubbell WL. Transmembrane protein structure: Spin labeling of bacteriorhodopsin mutants. Science 248: 1088‐1092, 1990.
 6. Amani M, Yousefi R, Moosavi‐Movahedi AA, Pintus F, Mura A, Floris G, Kurganov BI, Saboury AA. Structural changes and aggregation process of Cu/containing amine oxidase in the presence of 2,2,2′‐trifluoroethanol. Prot Pept Lett 15: 521‐527, 2008.
 7. Arora A, Abildgaard F, Bushweller JH, Tamm LK. Structure of outer membrane protein A transmembrane domain by NMR spectroscopy. Nat Struct Biol 8: 334‐338, 2001.
 8. Ashwell JD, Klusner RD. Genetic and mutational analysis of the T‐cell antigen receptor. Ann Rev Immunol 8: 139‐167, 1990.
 9. Ayala I, Sounier R, Use N, Gans P, Boisbouvier J. An efficient protocol for the complete incorporation of methyl‐protonated alanine in perdeuterated protein. J Biomol NMR 43: 111‐119, 2009.
 10. Baker KA, Tzitzilonis C, Kwiatkowski W, Choe S, Riek R. Conformational dynamics of the KcsA potassium channel governs gating properties. Nat Struct Mol Biol 14: 1089‐1095, 2007.
 11. Baldermann C, Lupas A, Lubieniecki J, Engelhardt H. The regulated outer membrane protein Omp21 from Comamonas acidovorans is identified as a member of a new family of eight‐stranded beta‐sheet proteins by its sequence and properties. J Bacteriology 180: 3741‐3749, 1998.
 12. Baldwin S. Membrane Transport, a Practical Approach. Oxford University Press, 2000.
 13. Baskaran K, Brunner K, Munte CE, Kalbitzer HR. Mapping of protein structural ensembles by chemical shifts. J Biomol NMR 48: 71‐83, 2010.
 14. Battiste JL, Wagner G. Utilization of site‐directed spin labeling and high‐resolution heteronuclear nuclear magnetic resonance for global fold determination of large proteins with limited nuclear overhauser effect data. Biochemistry 39: 5355‐5365, 2000.
 15. Bax A, Tjandra N. High‐resolution heteronuclear NMR of human ubiquitin in an aqueous liquid crystalline medium. J Biomol NMR 10: 289‐292, 1997.
 16. Bayburt TH, Sligar SG. Single‐molecule height measurements on microsomal cytochrome P450 in nanometer‐scale phospholipid bilayer disks. Proc Natl Acad Sci U S A 99: 6725‐6730, 2002.
 17. Bayrhuber M, Meins T, Habeck M, Becker S, Giller K, Villinger S, Vonrhein C, Griesinger C, Zweckstetter M, Zeth K. Structure of the human voltage‐dependent anion channel. Proc Natl Acad Sci U S A 105: 15370‐15375, 2008.
 18. Becattini B, Pellecchia M. SAR by ILOEs: An NMR‐based approach to reverse chemical genetics. Chemistry 12: 2658‐2662, 2006.
 19. Becucci L, Cembran A, Karim CB, Thomas DD, Guidelli R, Gao J, Veglia G. On the function of pentameric phospholamban: Ion channel or storage form? Biophys J 96: L60‐L62, 2009.
 20. Bellot G, Granier S, Bourguet W, Seyer R, Rahmeh R, Mouillac B, Pascal R, Mendre C, Demene H. Structure of the third intracellular loop of the vasopressin V2 receptor and conformational changes upon binding to gC1qR. J Mol Biol 388: 491‐507, 2009.
 21. Benga G. Water channel proteins (later called aquaporins) and relatives: Past, present, and future. IUBMB Life 61: 112‐133, 2009.
 22. Berman HM, Bhat TN, Bourne PE, Feng Z, Gilliland G, Weissig H, Westbrook J. The Protein Data Bank and the challenge of structural genomics. Nat Struct Biol 7: 957‐959, 2000.
 23. Bocharov EV, Mayzel ML, Volynsky PE, Goncharuk MV, Ermolyuk YS, Schulga AA, Artemenko EO, Efremov RG, Arseniev AS. Spatial structure and pH‐dependent conformational diversity of dimeric transmembrane domain of the receptor tyrosine kinase EphA1. J Biol Chem 283: 29385‐29395, 2008.
 24. Bocharov EV, Mayzel ML, Volynsky PE, Mineev KS, Tkach EN, Ermolyuk YS, Schulga AA, Efremov RG, Arseniev AS. Left‐handed dimer of EphA2 transmembrane domain: Helix packing diversity among receptor tyrosine kinases. Biophys J 98: 881‐889, 2010.
 25. Bocharov EV, Mineev KS, Volynsky PE, Ermolyuk YS, Tkach EN, Sobol AG, Chupin VV, Kirpichnikov MP, Efremov RG, Arseniev AS. Spatial structure of the dimeric transmembrane domain of the growth factor receptor ErbB2 presumably corresponding to the receptor active state. J Biol Chem 283: 6950‐6956, 2008.
 26. Bocharov EV, Pustovalova YE, Pavlov KV, Volynsky PE, Goncharuk MV, Ermolyuk YS, Karpunin DV, Schulga AA, Kirpichnikov MP, Efremov RG, Maslennikov IV, Arseniev AS. Unique dimeric structure of BNip3 transmembrane domain suggests membrane permeabilization as a cell death trigger. J Biol Chem 282: 16256‐16266, 2007.
 27. Bohm HJ, Banner D, Bendels S, Kansy M, Kuhn B, Muller K, Obst‐Sander U, Stahl M. Fluorine in medicinal chemistry. Chembiochem 5: 637‐643, 2004.
 28. Bondarenko V, Tillman T, Xu Y, Tang P. NMR structure of the transmembrane domain of the n‐acetylcholine receptor beta2 subunit. Biochim Biophys Acta 1798: 1608‐1614, 2010.
 29. Booth PJ, Clarke J. Membrane protein folding makes the transition. Proc Natl Acad Sci U S A 107: 3947‐3948, 2010.
 30. Borch J, Hamann T. The nanodisc: A novel tool for membrane protein studies. Biol Chem 390: 805‐814, 2009.
 31. Boudker O, Verdon G. Structural perspectives on secondary active transporters. Trends Pharm Sci 31: 418‐426, 2010.
 32. Bouvignies G, Korzhnev DM, Neudecker P, Hansen DF, Cordes MH, Kay LE. A simple method for measuring signs of (1)H (N) chemical shift differences between ground and excited protein states. J Biomol NMR 47: 135‐141, 2010.
 33. Bowie JU. Helix packing angle preferences. Nat Struct Biol 4: 915‐917, 1997a.
 34. Bowie JU. Helix packing in membrane proteins. J Mol Biol 272: 780‐789, 1997b.
 35. Bowie JU. Solving the membrane protein folding problem. Nature 438: 581‐589, 2005.
 36. Call ME, Chou JJ. A view into the blind spot: Solution NMR provides new insights into signal transduction across the lipid bilayer. Structure 18: 1559‐1569, 2010.
 37. Call ME, Schnell JR, Xu C, Lutz RA, Chou JJ, Wucherpfennig KW. The structure of the zetazeta transmembrane dimer reveals features essential for its assembly with the T cell receptor. Cell 127: 355‐368, 2006.
 38. Call ME, Wucherpfennig KW, Chou JJ. The structural basis for intramembrane assembly of an activating immunoreceptor complex. Nat Immunnol 11: 1023‐1029, 2010.
 39. Catoire LJ, Zoonens M, van Heijenoort C, Giusti F, Guittet E, Popot JL. Solution NMR mapping of water‐accessible residues in the transmembrane beta‐barrel of OmpX. Eur Biophys J 39: 623‐630, 2010.
 40. Cavalli A, Salvatella X, Dobson CM, Vendruscolo M. Protein structure determination from NMR chemical shifts. Proc Natl Acad Sci U S A 104: 9615‐9620, 2007.
 41. Chamani J, Moosavi‐Movahedi AA, Rajabi O, Gharanfoli M, Momen‐Heravi M, Hakimelahi GH, Neamati‐Baghsiah A, Varasteh AR. Cooperative alpha‐helix formation of beta‐lactoglobulin induced by sodium n‐alkyl sulfates. J Coll Interface Sci 293: 52‐60, 2006.
 42. Champeil P, Menguy T, Tribet C, Popot JL, le Maire M. Interaction of amphipols with sarcoplasmic reticulum Ca2+‐ATPase. J Biol Chem 275: 18623‐18637, 2000.
 43. Chandrasekhar K, Profy AT, Dyson HJ. Solution conformational preferences of immunogenic peptides derived from the principal neutralizing determinant of the HIV‐1 envelope glycoprotein gp120. Biochemistry 30: 9187‐9194, 1991.
 44. Chen H, Ji F, Olman V, Mobley CK, Liu Y, Zhou Y, Bushweller JH, Prestegard JH, Xu Y. Optimal mutation aites for PRE sata collection and membrane protein structure prediction. Structure 19: 484‐495, 2011.
 45. Chen K, Tjandra N. Top‐down approach in protein RDC data analysis: De novo estimation of the alignment tensor. J Biomol NMR 38: 303‐313, 2007.
 46. Chevelkov V, Xue Y, Rao DK, Forman‐Kay JD, Skrynnikov NR. 15N H/D‐SOLEXSY experiment for accurate measurement of amide solvent exchange rates: Application to denatured drkN SH3. J Biomol NMR 46: 227‐244, 2010.
 47. Chill JH, Louis JM, Miller C, Bax A. NMR study of the tetrameric KcsA potassium channel in detergent micelles. Protein Sci 15: 684‐698, 2006.
 48. Choowongkomon K, Carlin CR, Sonnichsen FD. A structural model for the membrane‐bound form of the juxtamembrane domain of the epidermal growth factor receptor. J Biol Chem 280: 24043‐24052, 2005.
 49. Chou JJ, Gaemers S, Howder B, Louis JM, Bax A. A simple apparatus for generating stretched polyacrylamide gels, yielding uniform alignment of proteins and detergent micelles. J Biomol NMR 21: 377‐382, 2001.
 50. Chu S, Maltsev S, Emwas AH, Lorigan GA. Solid‐state NMR paramagnetic relaxation enhancement immersion depth studies in phospholipid bilayers. J Magn Reson 207: 89‐94, 2010.
 51. Cierpicki T, Liang B, Tamm LK, Bushweller JH. Increasing the accuracy of solution NMR structures of membrane proteins by application of residual dipolar couplings. High‐resolution structure of outer membrane protein A. J Am Chem Soc 128: 6947‐6951, 2006.
 52. Claasen B, Axmann M, Meinecke R, Meyer B. Direct observation of ligand binding to membrane proteins in living cells by a saturation transfer double difference (STDD) NMR spectroscopy method shows a significantly higher affinity of integrin alpha(IIb)beta3 in native platelets than in liposomes. J Am Chem Soc 127: 916‐919, 2005.
 53. Clore GM, Driscoll PC, Wingfield PT, Gronenborn AM. Analysis of the backbone dynamics of interleukin‐1 beta using two‐dimensional inverse detected heteronuclear 15N‐1H NMR spectroscopy. Biochemistry 29: 7387‐7401, 1990.
 54. Colman PM, Lawrence MC. The structural biology of type I viral membrane fusion. Nat Rev Mol Cell Biol 4: 309‐319, 2003.
 55. Czerski L, Vinogradova O, Sanders CR. NMR‐Based amide hydrogen‐deuterium exchange measurements for complex membrane proteins: Development and critical evaluation. J Magn Reson 142: 111‐119, 2000.
 56. Dalmas O, Cuello LG, Jogini V, Cortes DM, Roux B, Perozo E. Structural dynamics of the magnesium‐bound conformation of CorA in a lipid bilayer. Structure 18: 868‐878, 2010.
 57. Dalvit C. 1H to 15N polarization transfer via 1H chemical‐shift anisotropy‐1H–15N dipole dipole cross corelation. J Magn Reson 97: 645‐650, 1992.
 58. Dalvit C, Mongelli N, Papeo G, Giordano P, Veronesi M, Moskau D, Kummerle R. Sensitivity improvement in 19F NMR‐based screening experiments: Theoretical considerations and experimental applications. J Am Chem Soc 127: 13380‐13385, 2005.
 59. Danielsson J, Liljedahl L, Barany‐Wallje E, Sonderby P, Kristensen LH, Martinez‐Yamout MA, Dyson HJ, Wright PE, Poulsen FM, Maler L, Graslund A, Kragelund BB. The intrinsically disordered RNR inhibitor Sml1 is a dynamic dimer. Biochemistry 47: 13428‐13437, 2008.
 60. Daragan VA, Mayo KH. Analysis of internally restricted correlated rotations in peptides and proteins using 13C and 15N NMR relaxation data. J Phys Chem 100: 8278‐8288, 1996.
 61. Daragan VA, Mayo KH. Motional model analyses of protein and peptide dynamics using 13C and 15N NMR. Prog NMR Spect 21: 63‐105, 1997.
 62. Das TK, Mazumdar S, Mitra S. Characterization of a partially unfolded structure of cytochrome c induced by sodium dodecyl sulphate and the kinetics of its refolding. Eur J Biochem 254: 662‐670, 1998.
 63.De Simone A, Cavalli A, Hsu ST, Vranken W, Vendruscolo M. Accurate random coil chemical shifts from an analysis of loop regions in native states of proteins. J Am Chem Soc 131: 16332‐16333, 2009.
 64. Dempsey CE, Handcock LJ. Hydrogen bond stabilities in membrane‐reconstituted alamethicin from amide‐resolved hydrogen‐exchange measurements. Biophys J 70: 1777‐1788, 1996.
 65. Denisov IG, Grinkova YV, Lazarides AA, Sligar SG. Directed self‐assembly of monodisperse phospholipid bilayer Nanodiscs with controlled size. J Am Chem Soc 126: 3477‐3487, 2004.
 66. Douglas SM, Chou JJ, Shih WM. DNA‐nanotube‐induced alignment of membrane proteins for NMR structure determination. Proc Natl Acad Sci U S A 104: 6644‐6648, 2007.
 67. Drin G, Casella JF, Gautier R, Boehmer T, Schwartz TU, Antonny B. A general amphipathic alpha‐helical motif for sensing membrane curvature. Nat Struct Mol Biol 14: 138‐146, 2007.
 68. Drummer HE, Boo I, Poumbourios P. Mutagenesis of a conserved fusion peptide‐like motif and membrane‐proximal heptad‐repeat region of hepatitis C virus glycoprotein E1. J Gen Virol 88: 1144‐1148, 2007.
 69. Dunker AK, Obradovic Z, Romero P, Garner EC, Brown CJ. Intrinsic protein disorder in complete genomes. Genome Inform Ser Workshop Genome Inform 11: 161‐171, 2000.
 70. Dyson HJ, Wright PE. Intrinsically unstructured proteins and their functions. Nat Rev Mol Cell Biol 6: 197‐208, 2005.
 71. Englander SW, Mayne L, Krishna MM. Protein folding and misfolding: Mechanism and principles. Q Rev Biophys 40: 287‐326, 2007.
 72. Engler A, Stangler T, Willbold D. Solution structure of human immunodeficiency virus type 1 Vpr(13‐33) peptide in micelles. Eur J Biochem 268: 389‐395, 2001.
 73. Engler A, Stangler T, Willbold D. Structure of human immunodeficiency virus type 1 Vpr(34‐51) peptide in micelle containing aqueous solution. Eur J Biochem 269: 3264‐3269, 2002.
 74. Eriksson HM, Persson K, Zhang S, Wieslander K. High‐yield expression and purification of a monotopic membrane glycosyltransferase. Protein Expres Pur 66: 143‐148, 2009.
 75. Fajardo DA, Cheung J, Ito C, Sugawara E, Nikaido H, Misra R. Biochemistry and regulation of a novel escherichia coli K‐12 porin protein, OmpG, which produces unusually large channels. J Bacteriology 180: 4452‐4459, 1998.
 76. Fendler JH. Interactions and reactions in reversed micellar systems. Accts Chem Res 9: 153‐161, 1976.
 77. Fernandez C, Hilty C, Bonjour S, Adeishvili K, Pervushin K, Wuthrich K. Solution NMR studies of the integral membrane proteins OmpX and OmpA from Escherichia coli. FEBS Lett 504: 173‐178, 2001.
 78. Fernandez C, Hilty C, Wider G, Guntert P, Wuthrich K. NMR structure of the integral membrane protein OmpX. J Mol Biol 336: 1211‐1221, 2004.
 79. Fiaux J, Bertelsen EB, Horwich AL, Wuthrich K. Uniform and residue‐specific 15N‐labeling of proteins on a highly deuterated background. J Biomol NMR 29: 289‐297, 2004.
 80. Fillingame RH, Girvin ME, Jiang W, Valiyaveetil F, Hermolin J. Subunit interactions coupling H +transport and ATP synthesis in F1F0 ATP synthase. Acta Phys Scand 643: 163‐168, 1998.
 81. Fleming K, Matthews S. Media for studies of partially aligned states. Methods Mol Biol 278: 79‐88, 2004.
 82. Franzin CM, Teriete P, Marassi FM. Structural similarity of a membrane protein in micelles and membranes. J Am Chem Soc 129: 8078‐8079, 2007.
 83. Gaponenko V, Sarma SP, Altieri AS, Horita DA, Li J, Byrd RA. Improving the accuracy of NMR structures of large proteins using pseudocontact shifts as long‐range restraints. J Biomol NMR 28: 205‐212, 2004.
 84. Garza AM, Khan SH, Kumar R. Site‐specific phosphorylation induces functionally active conformation in the intrinsically disordered N‐terminal activation function (AF1) domain of the glucocorticoid receptor. Mol Cell Biol 30: 220‐230, 2010.
 85. Gautier A, Kirkpatrick JP, Nietlispach D. Solution‐state NMR spectroscopy of a seven‐helix transmembrane protein receptor: Backbone assignment, secondary structure, and dynamics. Angewandte Chem 47: 7297‐7300, 2008.
 86. 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 17: 768‐774, 2010.
 87. Gemmercker G, Jahnke W, Kessler H. Measurement of fast proton exchange rates in isotopically lebaled compounds. J Am Chem Soc 115: 11620‐11621, 1993.
 88. Gerlach H, Laumann V, Martens S, Becker CF, Goody RS, Geyer M. HIV‐1 Nef membrane association depends on charge, curvature, composition and sequence. Nat Chem Biol 6: 46‐53, 2010.
 89. Girvin ME, Rastogi VK, Abildgaard F, Markley JL, Fillingame RH. Solution structure of the transmembrane H+‐transporting subunit c of the F1F0 ATP synthase. Biochemistry 37: 8817‐8824, 1998.
 90. Glover JN, Harrison SC. Crystal structure of the heterodimeric bZIP transcription factor c‐Fos‐c‐Jun bound to DNA. Nature 373: 257‐261, 1995.
 91. Goetz M, Carlotti C, Bontems F, Dufourc EJ. Evidence for an alpha‐helix – pi‐bulge helicity modulation for the neu/erbB‐2 membrane‐spanning segment. A 1H NMR and circular dichroism study. Biochemistry 40: 6534‐6540, 2001.
 92. Goto NK, Gardner KH, Mueller GA, Willis RC, Kay LE. A robust and cost‐effective method for the production of Val, Leu, Ile (delta 1) methyl‐protonated 15N‐, 13C‐, 2H‐labeled proteins. J Biomol NMR 13: 369‐374, 1999.
 93. Grace CR, Perrin MH, Gulyas J, Rivier JE, Vale WW, Riek R. NMR structure of the first extracellular domain of corticotropin‐releasing factor receptor 1 (ECD1‐CRF‐R1) complexed with a high affinity agonist. J Biol Chem 285: 38580‐38589, 2010.
 94. Hajduk PJ. SAR by NMR: Putting the pieces together. Mol Interv 6: 266‐272, 2006.
 95. Hajduk PJ, Dinges J, Miknis GF, Merlock M, Middleton T, Kempf DJ, Egan DA, Walter KA, Robins TS, Shuker SB, Holzman TF, Fesik SW. NMR‐based discovery of lead inhibitors that block DNA binding of the human papillomavirus E2 protein. J Med Chem 40: 3144‐3150, 1997.
 96. Hammill JT, Miyake‐Stoner S, Hazen JL, Jackson JC, Mehl RA. Preparation of site‐specifically labeled fluorinated proteins for 19F‐NMR structural characterization. Nat Protoc 2: 2601‐2607, 2007.
 97. Hansen DF, Vallurupalli P, Lundstrom P, Neudecker P, Kay LE. Probing chemical shifts of invisible states of proteins with relaxation dispersion NMR spectroscopy: How well can we do? J Am Chem Soc 130: 2667‐2675, 2008.
 98. Hansen MR, Mueller L, Pardi A. Tunable alignment of macromolecules by filamentous phage yields dipolar coupling interactions. Nat Struct Biol 5: 1065‐1074, 1998.
 99. Haq SK, Khan RH. Effect of detergents and hexafluoroisopropanol on the conformation of a non‐helical and a helical plant protease inhibitor. Int J Biol Macromol 36: 47‐53, 2005.
 100. Haussinger D, Huang JR, Grzesiek S. DOTA‐M8: An extremely rigid, high‐affinity lanthanide chelating tag for PCS NMR spectroscopy. J Am Chem Soc 131: 14761‐14767, 2009.
 101. Hefke F, Bagaria A, Reckel S, Ullrich SJ, Dotsch V, Glaubitz C, Guntert P. Optimization of amino acid type‐specific (13)C and (15)N labeling for the backbone assignment of membrane proteins by solution‐ and solid‐state NMR with the UPLABEL algorithm. J Biomol NMR 49: 75‐84, 2011.
 102. Hiller S, Garces RG, Malia TJ, Orekhov VY, Colombini M, Wagner G. Solution structure of the integral human membrane protein VDAC‐1 in detergent micelles. Science 321: 1206‐1210, 2008.
 103. Hilty C, Wider G, Fernandez C, Wuthrich K. Membrane protein‐lipid interactions in mixed micelles studied by NMR spectroscopy with the use of paramagnetic reagents. Chembiochem 5: 467‐473, 2004.
 104. Howell SC, Mesleh MF, Opella SJ. NMR structure determination of a membrane protein with two transmembrane helices in micelles: MerF of the bacterial mercury detoxification system. Biochemistry 44: 5196‐5206, 2005.
 105. Huang JR, Grzesiek S. Ensemble calculations of unstructured proteins constrained by RDC and PRE data: A case study of urea‐denatured ubiquitin. J Am Chem Soc 132: 694‐705, 2010.
 106. Hubbell WL, Cafiso DS, Altenbach C. Identifying conformational changes with site‐directed spin labeling. Nat Struct Biol 7: 735‐739, 2000.
 107. Huysmans GH, Radford SE, Brockwell DJ, Baldwin SA. The N‐terminal helix is a post‐assembly clamp in the bacterial outer membrane protein PagP. J Mol Biol 373: 529‐540, 2007.
 108. Hvidt A, Linderstrom‐Lang K. Exchange of deuterium and 18O between water and other substances. III. Deuterium exchange of short peptides, Sanger's A‐chain and insulin. C R Trav Lab Carlsberg Chim 29: 385‐402, 1955.
 109. Hvidt A, Linderstrom‐Lang K. The kinetics of the deuterium exchange of insulin with D2O; an amendment. Biochim Biophys Acta 16: 168‐169, 1955.
 110. Hwang PM, Bishop RE, Kay LE. The integral membrane enzyme PagP alternates between two dynamically distinct states. Proc Natl Acad Sci U S A 101: 9618‐9623, 2004.
 111. Hwang PM, Choy WY, Lo EI, Chen L, Forman‐Kay JD, Raetz CR, Prive GG, Bishop RE, Kay LE. Solution structure and dynamics of the outer membrane enzyme PagP by NMR. Proc Natl Acad Sci U S A 99: 13560‐13565, 2002.
 112. Inaba K, Takahashi YH, Ito K, Hayashi S. Critical role of a thiolate‐quinone charge transfer complex and its adduct form in de novo disulfide bond generation by DsbB. Proc Natl Acad Sci U S A 103: 287‐292, 2006.
 113. Isaacson RL, Simpson PJ, Liu M, Cota E, Zhang X, Freemont P, Matthews S. A new labeling method for methyl transverse relaxation‐optimized spectroscopy NMR spectra of alanine residues. J Am Chem Soc 129: 15428‐15429, 2007.
 114. Jackson JC, Hammill JT, Mehl RA. Site‐specific incorporation of a (19)F‐amino acid into proteins as an NMR probe for characterizing protein structure and reactivity. J Am Chem Soc 129: 1160‐1166, 2007.
 115. Jahnke W, Rudisser S, Zurini M. Spin label enhanced NMR screening. J Am Chem Soc 123: 3149‐3150, 2001.
 116. Jaroniec CP, Kaufman JD, Stahl SJ, Viard M, Blumenthal R, Wingfield PT, Bax A. Structure and dynamics of micelle‐associated human immunodeficiency virus gp41 fusion domain. Biochemistry 44: 16167‐16180, 2005.
 117. Jarymowycz VA, Stone MJ. Fast time scale dynamics of protein backbones: NMR relaxation methods, applications, and functional consequences. Chem Rev 106: 1624‐1671, 2006.
 118. Javadpour MM, Eilers M, Groesbeek M, Smith SO. Helix packing in polytopic membrane proteins: Role of glycine in transmembrane helix association. Biophys J 77: 1609‐1618, 1999.
 119. Jeannin P, Bottazzi B, Sironi M, Doni A, Rusnati M, Presta M, Maina V, Magistrelli G, Haeuw JF, Hoeffel G, Thieblemont N, Corvaia N, Garlanda C, Delneste Y, Mantovani A. Complexity and complementarity of outer membrane protein A recognition by cellular and humoral innate immunity receptors. Immunity 22: 551‐560, 2005.
 120. Jensen GA, Andersen OM, Bonvin AM, Bjerrum‐Bohr I, Etzerodt M, Thogersen HC, O'Shea C, Poulsen FM, Kragelund BB. Binding site structure of one LRP‐RAP complex: Implications for a common ligand‐receptor binding motif. J Mol Biol 362: 700‐716, 2006.
 121. Jensen MR, Markwick PR, Meier S, Griesinger C, Zweckstetter M, Grzesiek S, Bernado P, Blackledge M. Quantitative determination of the conformational properties of partially folded and intrinsically disordered proteins using NMR dipolar couplings. Structure 17: 1169‐1185, 2009.
 122. Jonas A. Reconstitution of high‐density lipoproteins. Methods Enzym 128: 553‐582, 1986.
 123. Kang C, Tian C, Sonnichsen FD, Smith JA, Meiler J, George AL, Jr, Vanoye CG, Kim HJ, Sanders CR. Structure of KCNE1 and implications for how it modulates the KCNQ1 potassium channel. Biochemistry 47: 7999‐8006, 2008.
 124. Katragadda M, Alderfer JL, Yeagle PL. Solution structure of the loops of bacteriorhodopsin closely resembles the crystal structure. Biochim Biophys Acta 1466: 1‐6, 2000.
 125. Katragadda M, Alderfer JL, Yeagle PL. Assembly of a polytopic membrane protein structure from the solution structures of overlapping peptide fragments of bacteriorhodopsin. Biophys J 81: 1029‐1036, 2001.
 126. Katragadda M, Chopra A, Bennett M, Alderfer JL, Yeagle PL, Albert AD. Structures of the transmembrane helices of the G‐protein coupled receptor, rhodopsin. J Pept Res 58: 79‐89, 2001.
 127. Kay LE. Protein dynamics from NMR. Nat Struct Biol 5: 513‐517, 1998.
 128. Kielec JM, Valentine KG, Babu CR, Wand AJ. Reverse micelles in integral membrane protein structural biology by solution NMR spectroscopy. Structure 17: 345‐351, 2009.
 129. Kielec JM, Valentine KG, Wand AJ. A method for solution NMR structural studies of large integral membrane proteins: Reverse micelle encapsulation. Biochim Biophys Acta 1798: 150‐160, 2010.
 130. 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 Mag Res spec 55: 335‐360, 2009.
 131. Kim S, Jeon TJ, Oberai A, Yang D, Schmidt JJ, Bowie JU. Transmembrane glycine zippers: Physiological and pathological roles in membrane proteins. Proc Natl Acad Sci U S A 102: 14278‐14283, 2005.
 132. Kjaergaard M, Brander S, Poulsen FM. Random coil chemical shift for intrinsically disordered proteins: Effects of temperature and pH. J Biomol NMR 49: 139‐149, 2011.
 133. Kjaergaard M, Norholm AB, Hendus‐Altenburger R, Pedersen SF, Poulsen FM, Kragelund BB. Temperature‐dependent structural changes in intrinsically disordered proteins: Formation of alpha‐helices or loss of polyproline II? Protein Sci 19: 1555‐1564, 2011.
 134. Koenig BW, Mitchell DC, Konig S, Grzesiek S, Litman BJ, Bax A. Measurement of dipolar couplings in a transducin peptide fragment weakly bound to oriented photo‐activated rhodopsin. J Biomol NMR 16: 121‐125, 2000.
 135. Kontaxis G, Delaglio F, Bax A. Molecular fragment replacement approach to protein structure determination by chemical shift and dipolar homology database mining. Methods Enzymol 394: 42‐78, 2005.
 136. Kovacs RJ, Nelson MT, Simmerman HK, Jones LR. Phospholamban forms Ca2+‐selective channels in lipid bilayers. J Biol Chem 263: 18364‐18368, 1988.
 137. Kragelund BB, Knudsen J, Poulsen FM. Local perturbations by ligand binding of hydrogen deuterium exchange kinetics in a four‐helix bundle protein, acyl coenzyme A binding protein (ACBP). J Mol Biol 250: 695‐706, 1995.
 138. Krueger‐Koplin RD, Sorgen PL, Krueger‐Koplin ST, Rivera‐Torres IO, Cahill SM, Hicks DB, Grinius L, Krulwich TA, Girvin ME. An evaluation of detergents for NMR structural studies of membrane proteins. J Biomol NMR 28: 43‐57, 2004.
 139. Lau TL, Kim C, Ginsberg MH, Ulmer TS. The structure of the integrin alphaIIbbeta3 transmembrane complex explains integrin transmembrane signalling. EMBO J 28: 1351‐1361, 2009.
 140. Lawrie CM, Sulistijo ES, MacKenzie KR. Intermonomer hydrogen bonds enhance GxxxG‐driven dimerization of the BNIP3 transmembrane domain: Roles for sequence context in helix‐helix association in membranes. J Mol Biol 396: 924‐936, 2010.
 141. Lee BL, Sykes BD, Fliegel L. Structural analysis of the Na+/H+ exchanger isoform 1 (NHE1) using the divide and conquer approach. Biochem Cell Biol 89: 189‐199, 2011.
 142. Lee S, Mesleh MF, Opella SJ. Structure and dynamics of a membrane protein in micelles from three solution NMR experiments. J Biomol NMR 26: 327‐334, 2003.
 143. Li H, Li F, Qian ZM, Sun H. Structure and topology of the transmembrane domain 4 of the divalent metal transporter in membrane‐mimetic environments. Eur J Biochem 271: 1938‐1951, 2004.
 144. Liang B, Arora A, Tamm LK. Fast‐time scale dynamics of outer membrane protein A by extended model‐free analysis of NMR relaxation data. Biochim Biophys Acta 1798: 68‐76, 2010.
 145. Liang B, Bushweller JH, Tamm LK. Site‐directed parallel spin‐labeling and paramagnetic relaxation enhancement in structure determination of membrane proteins by solution NMR spectroscopy. J Am Chem Soc 128: 4389‐4397, 2006.
 146. Liang B, Tamm LK. Structure of outer membrane protein G by solution NMR spectroscopy. Proc Natl Acad Sci U S A 104: 16140‐16145, 2007.
 147. Lin SH, Guidotti G. Purification of membrane proteins. Methods Enzymol 463: 619‐629, 2009.
 148. Lindberg M, Graslund A. The position of the cell penetrating peptide penetratin in SDS micelles determined by NMR. FEBS Lett 497: 39‐44, 2001.
 149. Linderstrom‐Lang K. The pH‐dependence of the deuterium exchange of insulin. Biochim Biophys Acta 18: 308, 1955.
 150. Lipari G, Szabo A. Model‐free approach to the interpretation of nuclear magnetic resonance relaxation in macromolecules. I. Theory and range of validity. J Am Chem Soc 104: 4546‐4559, 1982.
 151. Lipari G, Szabo A. Model‐free approach to the interpretation of nuclear magnetic resonance relaxation in macromolecules. II. Analysis of experimental results. J Am Chem Soc 104: 4559‐4570, 1982.
 152. Liu Y, Bacon ER, Ballinger K, Black CD, Illig K, McIntire GL, Wang PP, O'Neil N, Kinter L, Desai VC. Pharmacokinetics and hepatic disposition of bis[1‐(ethoxycarbonyl)propyl]5‐acetylamino‐2,4,6‐triiodoisophthalate in rats and isolated perfused rat livers. Drug Metab Dispos 28: 731‐736, 2000.
 153. Losonczi JA, Prestegard JH. Nuclear magnetic resonance characterization of the myristoylated, N‐terminal fragment of ADP‐ribosylation factor 1 in a magnetically oriented membrane array. Biochemistry 37: 706‐716, 1998.
 154. Low C, Weininger U, Lee H, Schweimer K, Neundorf I, Beck‐Sickinger AG, Pastor RW, Balbach J. Structure and dynamics of helix‐0 of the N‐BAR domain in lipid micelles and bilayers. Biophys J 95: 4315‐4323, 2008.
 155. Lu H, Buck M, Radford SE, Dobson CM. Acceleration of the folding of hen lysozyme by trifluoroethanol. J Mol Biol 265: 112‐117, 1997.
 156. Luchette PA, Prosser RS, Sanders CR. Oxygen as a paramagnetic probe of membrane protein structure by cysteine mutagenesis and (19)F NMR spectroscopy. J Am Chem Soc 124: 1778‐1781, 2002.
 157. Lugovskoy AA, Maslennikov IV, Utkin YN, Tsetlin VI, Cohen JB, Arseniev AS. Spatial structure of the M3 transmembrane segment of the nicotinic acetylcholine receptor alpha subunit. Eur J Biochem 255: 455‐461, 1998.
 158. Ma C, Opella SJ. Lanthanide ions bind specifically to an added “EF‐hand” and orient a membrane protein in micelles for solution NMR spectroscopy. J Magn Reson 146: 381‐384, 2000.
 159. Ma D, Tillman TS, Tang P, Meirovitch E, Eckenhoff R, Carnini A, Xu Y. NMR studies of a channel protein without membranes: Structure and dynamics of water‐solubilized KcsA. Proc Natl Acad Sci U S A 105: 16537‐16542, 2008.
 160. MacKenzie KR, Prestegard JH, Engelman DM. A transmembrane helix dimer: Structure and implications. Science 276: 131‐133, 1997.
 161. Magidovich E, Orr I, Fass D, Abdu U, Yifrach O. Intrinsic disorder in the C‐terminal domain of the Shaker voltage‐activated K+ channel modulates its interaction with scaffold proteins. Proc Natl Acad Sci U S A 104: 13022‐13027, 2007.
 162. Malia TJ, Wagner G. NMR structural investigation of the mitochondrial outer membrane protein VDAC and its interaction with antiapoptotic Bcl‐xL. Biochemistry 46: 514‐525, 2007.
 163. Marassi FM, Opella SJ. A solid‐state NMR index of helical membrane protein structure and topology. J Magn Reson 144: 150‐155, 2000.
 164. Margottin F, Benichou S, Durand H, Richard V, Liu LX, Gomas E, Benarous R. Interaction between the cytoplasmic domains of HIV‐1 Vpu and CD4: Role of Vpu residues involved in CD4 interaction and in vitro CD4 degradation. Virology 223: 381‐386, 1996.
 165. Mascioni A, Karim C, Barany G, Thomas DD, Veglia G. Structure and orientation of sarcolipin in lipid environments. Biochemistry 41: 475‐482, 2002.
 166. Maslennikov I, Klammt C, Hwang E, Kefala G, Okamura M, Esquivies L, Mors K, Glaubitz C, Kwiatkowski W, Jeon YH, Choe S. Membrane domain structures of three classes of histidine kinase receptors by cell‐free expression and rapid NMR analysis. Proc Natl Acad Sci U S A 107: 10902‐10907, 2010.
 167. Mayer M, Meyer B. Group epitope mapping by saturation transfer difference NMR to identify segments of a ligand in direct contact with a protein receptor. J Am Chem Soc 123: 6108‐6117, 2001.
 168. McDermott A. Structure and dynamics of membrane proteins by magic angle spinning solid‐state NMR. Ann Rev Biophys 38: 385‐403, 2009.
 169. McLuskey K, Roszak AW, Zhu Y, Isaacs NW. Crystal structures of all‐alpha type membrane proteins. Eur Biophys J 39: 723‐755, 2010.
 170. Meinecke R, Meyer B. Determination of the binding specificity of an integral membrane protein by saturation transfer difference NMR: RGD peptide ligands binding to integrin alphaIIbbeta3. J Med Chem 44: 3059‐3065, 2001.
 171. Melnyk RA, Kim S, Curran AR, Engelman DM, Bowie JU, Deber CM. The affinity of GXXXG motifs in transmembrane helix‐helix interactions is modulated by long‐range communication. J Biol Chem 279: 16591‐16597, 2004.
 172. Mesleh MF, Veglia G, DeSilva TM, Marassi FM, Opella SJ. Dipolar waves as NMR maps of protein structure. J Am Chem Soc 124: 4206‐4207, 2002.
 173. Metcalf DG, Moore DT, Wu Y, Kielec JM, Molnar K, Valentine KG, Wand AJ, Bennett JS, DeGrado WF. NMR analysis of the alphaIIb beta3 cytoplasmic interaction suggests a mechanism for integrin regulation. Proc Natl Acad Sci U S A 107: 22481‐22486, 2010.
 174. Metcalfe EE, Zamoon J, Thomas DD, Veglia G. (1)H/(15)N heteronuclear NMR spectroscopy shows four dynamic domains for phospholamban reconstituted in dodecylphosphocholine micelles. Biophys J 87: 1205‐1214, 2004.
 175. Mineev KS, Bocharov EV, Pustovalova YE, Bocharova OV, Chupin VV, Arseniev AS. Spatial structure of the transmembrane domain heterodimer of ErbB1 and ErbB2 receptor tyrosine kinases. J Mol Biol 400: 231‐243, 2010.
 176. Mittermaier A, Kay LE. New tools provide new insights in NMR studies of protein dynamics. Science 312: 224‐228, 2006.
 177. Morth JP, Pedersen BP, Buch‐Pedersen MJ, Andersen JP, Vilsen B, Palmgren MG, Nissen P. A structural overview of the plasma membrane Na+,K+‐ATPase and H+‐ATPase ion pumps. Nat Rev Mol Cell Biol 12: 60‐70, 2011.
 178. Mulder FA, Filatov M. NMR chemical shift data and ab initio shielding calculations: Emerging tools for protein structure determination. Chem Soc Rev 39: 578‐590, 2010.
 179. Nambudripad R, Stark W, Opella SJ, Makowski L. Membrane‐mediated assembly of filamentous bacteriophage Pf1 coat protein. Science 252: 1305‐1308, 1991.
 180. Neumoin A, Cohen LS, Arshava B, Tantry S, Becker JM, Zerbe O, Naider F. Structure of a double transmembrane fragment of a G‐protein‐coupled receptor in micelles. Biophys J 96: 3187‐3196, 2009.
 181. Neves P, Lopes SC, Sousa I, Garcia S, Eaton P, Gameiro P. Characterization of membrane protein reconstitution in LUVs of different lipid composition by fluorescence anisotropy. J Pharm Biomed Anal 49: 276‐281, 2009.
 182. Nielsen N, Malmendal A, Vosegaard T. Techniques and applications of NMR to membrane proteins. Mol Membrane Biol 21: 129‐141, 2004.
 183. Ninomiya R, Matsuoka K, Moroi Y. Micelle formation of sodium chenodeoxycholate and solubilization into the micelles: Comparison with other unconjugated bile salts. Biochim Biophys Acta 1634: 116‐125, 2003.
 184. Nygaard EB, Lagerstedt JO, Bjerre G, Shi B, Budamagunta M, Poulsen KA, Meinild S, Rigor RR, Voss JC, Cala PM, Pedersen SF. Structural modeling and electron paramagnetic resonance spectroscopy of the human Na+/H+ exchanger isoform 1, NHE1. J Biol Chem 286: 634‐648, 2010.
 185. Nørholm A, Hendus‐Altenburger R, Bjerre G, Kjaergaard M, Petersen SF, Kragelund BB. The intracellular distal tail of the Na+/K+ exchanger NHE1 is intrinsically disordered; implications for NHE1 trafficing. Biochemistry 50: 3469‐3480, 2011.
 186. Op De Beeck A, Montserret R, Duvet S, Cocquerel L, Cacan R, Barberot B, Le Maire M, Penin F, Dubuisson J. The transmembrane domains of hepatitis C virus envelope glycoproteins E1 and E2 play a major role in heterodimerization. J Biol Chem 275: 31428‐31437, 2000.
 187. O'Shea EK, Rutkowski R, Kim PS. Evidence that the leucine zipper is a coiled coil. Science 243: 538‐542, 1989.
 188. Opella SJ, Marassi FM. Structure determination of membrane proteins by NMR spectroscopy. Chem Rev 104: 3587‐3606, 2004.
 189. Otten R, Villali J, Kern D, Mulder FA. Probing microsecond time scale dynamics in proteins by methyl (1)H carr‐purcell‐meiboom‐gill relaxation dispersion NMR measurements. Application to activation of the signaling protein NtrC(r). J Am Chem Soc 132: 17004‐17014, 2010.
 190. Ottiger M, Bax A. Characterization of magnetically oriented phospholipid micelles for measurement of dipolar couplings in macromolecules. J Biomol NMR 12: 361‐372, 1998.
 191. Otting G. Protein NMR using paramagnetic ions. Ann Rev Biophys 39: 387‐405, 2010.
 192. Oxenoid K, Chou JJ. The structure of phospholamban pentamer reveals a channel‐like architecture in membranes. Proc Natl Acad Sci U S A 102: 10870‐10875, 2005.
 193. Page RC, Lee S, Moore JD, Opella SJ, Cross TA. Backbone structure of a small helical integral membrane protein: A unique structural characterization. Protein Sci 18: 134‐146, 2009.
 194. Parker W, Song PS. Protein structures in SDS micelle‐protein complexes. Biophys J 61: 1435‐1439, 1992.
 195. Paulick MG, Bertozzi CR. The glycosylphosphatidylinositol anchor: A complex membrane‐anchoring structure for proteins. Biochemistry 47: 6991‐7000, 2008.
 196. Pervushin K, Riek R, Wider G, Wuthrich K. Attenuated T2 relaxation by mutual cancellation of dipole‐dipole coupling and chemical shift anisotropy indicates an avenue to NMR structures of very large biological macromolecules in solution. Proc Natl Acad Sci U S A 94: 12366‐12371, 1997.
 197. Prestegard JH, al‐Hashimi HM, Tolman JR. NMR structures of biomolecules using field oriented media and residual dipolar couplings. Q Rev Biophys 33: 371‐424, 2000.
 198. Prive GG. Detergents for the stabilization and crystallization of membrane proteins. Methods 41: 388‐397, 2007.
 199. Prosser RS, Evanics F, Kitevski JL, Patel S. The measurement of immersion depth and topology of membrane proteins by solution state NMR. Biochim Biophys Acta 1768: 3044‐3051, 2007.
 200. Ptitsyn OB, Uversky VN. The molten globule is a third thermodynamical state of protein molecules. FEBS Lett 341: 15‐18, 1994.
 201. Rader DJ. High‐density lipoproteins and atherosclerosis. Am J Cardiology 90: 62i‐70i, 2002.
 202. Rathinavelan T, Im W. A novel strategy to determine protein structures using exclusively residual dipolar coupling. J Comp Chem 29: 1640‐1649, 2008.
 203. Reddy T, Rainey JK. Interpretation of biomolecular NMR spin relaxation parameters. Biochem Cell Biol 88: 131‐142, 2010.
 204. Renault M, Cukkemane A, Baldus M. Solid‐state NMR spectroscopy on complex biomolecules. Angewandte Chem 49: 8346‐8357, 2010.
 205. Renault M, Saurel O, Czaplicki J, Demange P, Gervais V, Lohr F, Reat V, Piotto M, Milon A. Solution state NMR structure and dynamics of KpOmpA, a 210 residue transmembrane domain possessing a high potential for immunological applications. J Mol Biol 385: 117‐130, 2009.
 206. Riek R, Wider G, Pervushin K, Wuthrich K. Polarization transfer by cross‐correlated relaxation in solution NMR with very large molecules. Proc Natl Acad Sci U S A 96: 4918‐4923, 1999.
 207. Rodriguez‐Castaneda F, Haberz P, Leonov A, Griesinger C. Paramagnetic tagging of diamagnetic proteins for solution NMR. Magn Reson Chem 44: S10‐S16, 2006.
 208. Rosenbaum DM, Rasmussen SG, Kobilka BK. The structure and function of G‐protein‐coupled receptors. Nature 459: 356‐363, 2009.
 209. Ruan K, Tolman JR. Composite alignment media for the measurement of independent sets of NMR residual dipolar couplings. J Am Chem Soc 127: 15032‐15033, 2005.
 210. Rückert M, Otting G. Alignment of biological macromolecules in novel nonionic liquid crystalline media for NMR experiments. J Am Chem Soc 122: 7793‐7797, 2000.
 211. Salmon L, Nodet G, Ozenne V, Yin G, Jensen MR, Zweckstetter M, Blackledge M. NMR characterization of long‐range order in intrinsically disordered proteins. J Am Chem Soc 132: 8407‐8418, 2010.
 212. Salnikov E, Aisenbrey C, Vidovic V, Bechinger B. Solid‐state NMR approaches to measure topological equilibria and dynamics of membrane polypeptides. Biochim Biophys Acta 1798: 258‐265, 2010.
 213. Salzmann M, Pervushin K, Wider G, Senn H, Wuthrich K. TROSY in triple‐resonance experiments: New perspectives for sequential NMR assignment of large proteins. Proc Natl Acad Sci U S A 95: 13585‐13590, 1998.
 214. Sanders CR, II, Landis GC. Reconstitution of membrane proteins into lipid‐rich bilayered mixed micelles for NMR studies. Biochemistry 34: 4030‐4040, 1995.
 215. Sanders CR, Sonnichsen F. Solution NMR of membrane proteins: Practice and challenges. Magn Reson Chem 44: S24‐40, 2006.
 216. Sarramegna V, Talmont F, Demange P, Milon A. Heterologous expression of G‐protein‐coupled receptors: comparison of expression systems from the standpoint of large‐scale production and purification. Cell Mol Life Sci 60: 1529‐1546, 2003.
 217. Scarborough GA. Large single crystals of the Neurospora crassa plasma membrane H+‐ATPase: An approach to the crystallization of integral membrane proteins. Acta Cryst 50: 643‐649, 1994.
 218. Schibli DJ, Montelaro RC, Vogel HJ. The membrane‐proximal tryptophan‐rich region of the HIV glycoprotein, gp41, forms a well‐defined helix in dodecylphosphocholine micelles. Biochemistry 40: 9570‐9578, 2001.
 219. Schnell JR, Chou JJ. Structure and mechanism of the M2 proton channel of influenza A virus. Nature 451: 591‐595, 2008.
 220. Schwalbe H, Fiebig KM, Buck M, Jones JA, Grimshaw SB, Spencer A, Glaser SJ, Smith LJ, Dobson CM. Structural and dynamical properties of a denatured protein. Heteronuclear 3D NMR experiments and theoretical simulations of lysozyme in 8 M urea. Biochemistry 36: 8977‐8991, 1997.
 221. Schwarzinger S, Kroon GJ, Foss TR, Wright PE, Dyson HJ. Random coil chemical shifts in acidic 8 M urea: implementation of random coil shift data in NMRView. J Biomol NMR 18: 43‐48, 2000.
 222. Senes A, Gerstein M, Engelman DM. Statistical analysis of amino acid patterns in transmembrane helices: The GxxxG motif occurs frequently and in association with beta‐branched residues at neighboring positions. J Mol Biol 296: 921‐936, 2000.
 223. 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 44: 213‐223, 2009.
 224. Shen Y, Lange O, Delaglio F, Rossi P, Aramini JM, Liu G, Eletsky A, Wu Y, Singarapu KK, Lemak A, Ignatchenko A, Arrowsmith CH, Szyperski T, Montelione GT, Baker D, Bax A. Consistent blind protein structure generation from NMR chemical shift data. Proc Natl Acad Sci U S A 105: 4685‐4690, 2008.
 225. Shenkarev ZO, Lyukmanova EN, Paramonov AS, Shingarova LN, Chupin VV, Kirpichnikov MP, Blommers MJ, Arseniev AS. Lipid‐protein nanodiscs as reference medium in detergent screening for high‐resolution NMR studies of integral membrane proteins. J Am Chem Soc 132: 5628‐5629.
 226. Shenkarev ZO, Paramonov AS, Lyukmanova EN, Shingarova LN, Yakimov SA, Dubinnyi MA, Chupin VV, Kirpichnikov MP, Blommers MJ, Arseniev AS. NMR structural and dynamical investigation of the isolated voltage‐sensing domain of the potassium channel KvAP: Implications for voltage gating. J Am Chem Soc 132: 5630‐5637, 2010.
 227. Sheppard D, Guo C, Tugarinov V. 4D 1H‐13C NMR spectroscopy for assignments of alanine methyls in large and complex protein structures. J Am Chem Soc 131: 1364‐1365, 2009.
 228. Shon KJ, Kim Y, Colnago LA, Opella SJ. NMR studies of the structure and dynamics of membrane‐bound bacteriophage Pf1 coat protein. Science 252: 1303‐1305, 1991.
 229. Shuker SB, Hajduk PJ, Meadows RP, Fesik SW. Discovering high‐affinity ligands for proteins: SAR by NMR. Science 274: 1531‐1534, 1996.
 230. Sigalov AB. Protein intrinsic disorder and oligomericity in cell signaling. Mol Bio Sys 6: 451‐461, 2010.
 231. Slepkov ER, Rainey JK, Li X, Liu Y, Cheng FJ, Lindhout DA, Sykes BD, Fliegel L. Structural and functional characterization of transmembrane segment IV of the NHE1 isoform of the Na+/H+ exchanger. J Biol Chem 280: 17863‐17872, 2005.
 232. Smith SG, Mahon V, Lambert MA, Fagan RP. A molecular Swiss army knife: OmpA structure, function and expression. FEMS Microbiol Lett 273: 1‐11, 2007.
 233. Soulie S, Neumann JM, Berthomieu C, Moller JV, le Maire M, Forge V. NMR conformational study of the sixth transmembrane segment of sarcoplasmic reticulum Ca2+‐ATPase. Biochemistry 38: 5813‐5821, 1999.
 234. Spadaccini R, D'Errico G, D'Alessio V, Notomista E, Bianchi A, Merola M, Picone D. Structural characterization of the transmembrane proximal region of the hepatitis C virus E1 glycoprotein. Biochim Biophys Acta 1798: 344‐353, 2010.
 235. Sprangers R, Velyvis A, Kay LE. Solution NMR of supramolecular complexes: Providing new insights into function. Nat Methods 4: 697‐703, 2007.
 236. Sticht H, Willbold D, Ejchart A, Rosin‐Arbesfeld R, Yaniv A, Gazit A, Rosch P. Trifluoroethanol stabilizes a helix‐turn‐helix motif in equine infectious‐anemia‐virus trans‐activator protein. Eur J Biochem 225: 855‐861, 1994.
 237. Strandberg E, Esteban‐Martin S, Salgado J, Ulrich AS. Orientation and dynamics of peptides in membranes calculated from 2H‐NMR data. Biophys J 96: 3223‐3232, 2009.
 238. Su XC, Otting G. Paramagnetic labelling of proteins and oligonucleotides for NMR. J Biomol NMR 46: 101‐112, 2010.
 239. Sulistijo ES, Mackenzie KR. Structural basis for dimerization of the BNIP3 transmembrane domain. Biochemistry 48: 5106‐5120, 2009.
 240. Surrey T, Jahnig F. Refolding and oriented insertion of a membrane protein into a lipid bilayer. Proc Natl Acad Sci U S A 89: 7457‐7461, 1992.
 241. Tabuchi M, Yoshimori T, Yamaguchi K, Yoshida T, Kishi F. Human NRAMP2/DMT1, which mediates iron transport across endosomal membranes, is localized to late endosomes and lysosomes in HEp‐2 cells. J Biol Chem 275: 22220‐22228, 2000.
 242. Tamm LK, Liang B. NMR of membrane proteins in solution. Prog NMR Spec 48: 201‐210, 2006.
 243. Teilum K, Kragelund BB, Poulsen FM. Transient structure formation in unfolded acyl‐coenzyme A‐binding protein observed by site‐directed spin labelling. J Mol Biol 324: 349‐357, 2002.
 244. Teilum K, Olsen JG, Kragelund BB. Functional aspects of protein flexibility. Cell Mol Life Sci 66: 2231‐2247, 2009.
 245. Teilum K, Smith MH, Schulz E, Christensen LC, Solomentsev G, Oliveberg M, Akke M. Transient structural distortion of metal‐free Cu/Zn superoxide dismutase triggers aberrant oligomerization. Proc Natl Acad Sci U S A 106: 18273‐18278, 2009.
 246. Teriete P, Franzin CM, Choi J, Marassi FM. Structure of the Na,K‐ATPase regulatory protein FXYD1 in micelles. Biochemistry 46: 6774‐6783, 2007.
 247. Topiol S, Sabio M. X‐ray structure breakthroughs in the GPCR transmembrane region. Biochem Pharm 78: 11‐20, 2009.
 248. Traaseth NJ, Shi L, Verardi R, Mullen DG, Barany G, Veglia G. Structure and topology of monomeric phospholamban in lipid membranes determined by a hybrid solution and solid‐state NMR approach. Proc Natl Acad Sci U S A 106: 10165‐10170, 2009.
 249. Tribet C, Audebert R, Popot JL. Amphipols: Polymers that keep membrane proteins soluble in aqueous solutions. Proc Natl Acad Sci U S A 93: 15047‐15050, 1996.
 250. Tsai CJ, Kumar S, Ma B, Nussinov R. Folding funnels, binding funnels, and protein function. Protein Sci 8: 1181‐1190, 1999.
 251. Tsao DH, Sutherland AG, Jennings LD, Li Y, Rush TS, III, Alvarez JC, Ding W, Dushin EG, Dushin RG, Haney SA, Kenny CH, Malakian AK, Nilakantan R, Mosyak L. Discovery of novel inhibitors of the ZipA/FtsZ complex by NMR fragment screening coupled with structure‐based design. Bioorg Med Chem 14: 7953‐7961, 2006.
 252. Tugarinov V, Muhandiram R, Ayed A, Kay LE. Four‐dimensional NMR spectroscopy of a 723‐residue protein: Chemical shift assignments and secondary structure of malate synthase g. J Am Chem Soc 124: 10025‐10035, 2002.
 253. Tycko R, Blanco FJ, Ishii Y. Alignment of bioploymers in strained gels: A new way to create detrecetable dipole‐dipole couplings in high‐resolution biomolecular NMR. J Am Chem Soc 9340‐9341, 2000.
 254. Tyukhtenko S, Tiburu EK, Deshmukh L, Vinogradova O, Janero DR, Makriyannis A. NMR solution structure of human cannabinoid receptor‐1 helix 7/8 peptide: Candidate electrostatic interactions and microdomain formation. Biochem Biophys Res Comm 390: 441‐446, 2009.
 255. Tzeng J, Lee BL, Sykes BD, Fliegel L. Structural and functional analysis of transmembrane segment VI of the NHE1 isoform of the Na+/H+ exchanger. J Biol Chem 285: 36656‐36665, 2010.
 256. Uversky VN. Cracking the folding code. Why do some proteins adopt partially folded conformations, whereas other don't? FEBS Lett 514: 181‐183, 2002.
 257. Uversky VN, Oldfield CJ, Dunker AK. Intrinsically disordered proteins in human diseases: Introducing the D2 concept. Ann Rev Biophys 37: 215‐246, 2008.
 258. Valentine KG, Peterson RW, Saad JS, Summers MF, Xu X, Ames JB, Wand AJ. Reverse micelle encapsulation of membrane‐anchored proteins for solution NMR studies. Structure 18: 9‐16, 2010.
 259. 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 324: 1726‐1729, 2009.
 260. Van Horn WD, Ogilvie ME, Flynn PF. Use of reverse micelles in membrane protein structural biology. J Biomol NMR 40: 203‐211, 2008.
 261. Veglia G, Zeri AC, Ma C, Opella SJ. Deuterium/hydrogen exchange factors measured by solution nuclear magnetic resonance spectroscopy as indicators of the structure and topology of membrane proteins. Biophys J 82: 2176‐2183, 2002.
 262. Vermeer LS, Reat V, Hemminga MA, Milon A. Structural properties of a peptide derived from H+ ‐V‐ATPase subunit a. Biochim Biophys Acta 1788: 1204‐1212, 2009.
 263. Vinogradova O, Vaynberg J, Kong X, Haas TA, Plow EF, Qin J. Membrane‐mediated structural transitions at the cytoplasmic face during integrin activation. Proc Natl Acad Sci U S A 101: 4094‐4099, 2004.
 264. Vinogradova O, Velyvis A, Velyviene A, Hu B, Haas T, Plow E, Qin J. A structural mechanism of integrin alpha(IIb)beta(3) “inside‐out” activation as regulated by its cytoplasmic face. Cell 110: 587‐597, 2002.
 265. Walters RF, DeGrado WF. Helix‐packing motifs in membrane proteins. Proc Natl Acad Sci U S A 103: 13658‐13663, 2006.
 266. Wang J, Cady SD, Balannik V, Pinto LH, DeGrado WF, Hong M. Discovery of spiro‐piperidine inhibitors and their modulation of the dynamics of the M2 proton channel from influenza A virus. J Am Chem Soc 131: 8066‐8076, 2009.
 267. Wang J, Pielak RM, McClintock MA, Chou JJ. Solution structure and functional analysis of the influenza B proton channel. Nat Struct Mol Biol 16: 1267‐1271, 2009.
 268. Wang JF, Wei DQ, Chou KC. Insights from investigating the interactions of adamantane‐based drugs with the M2 proton channel from the H1N1 swine virus. Biochem Biophys Res Comm 388: 413‐417, 2009.
 269. Wang Z, Jones JD, Rizo J, Gierasch LM. Membrane‐bound conformation of a signal peptide: A transferred nuclear Overhauser effect analysis. Biochemistry 32: 13991‐13999, 1993.
 270. Warne T, Serrano‐Vega MJ, Baker JG, Moukhametzianov R, Edwards PC, Henderson R, Leslie AG, Tate CG, Schertler GF. Structure of a beta1‐adrenergic G‐protein‐coupled receptor. Nature 454: 486‐491, 2008.
 271. Watson RM, Woody RW, Lewis RV, Bohle DS, Andreotti AH, Ray B, Miller KW. Conformational changes in pediocin AcH upon vesicle binding and approximation of the membrane‐bound structure in detergent micelles. Biochemistry 40: 14037‐14046, 2001.
 272. Watts A, Straus SK, Grage SL, Kamihira M, Lam YH, Zhao X. Membrane protein structure determination using solid‐state NMR. Methods Mol Biol 278: 403‐473, 2004.
 273. Whitehead B, Craven CJ, Waltho JP. Double and triple resonance NMR methods for protein assignment. Methods Mol Biol 60: 29‐52, 1997.
 274. Wiener MC. A pedestrian guide to membrane protein crystallization. Methods 34: 364‐372, 2004.
 275. Wimmer R, Andersen KK, Vad B, Davidsen M, Molgaard S, Nesgaard LW, Kristensen HH, Otzen DE. Versatile interactions of the antimicrobial peptide novispirin with detergents and lipids. Biochemistry 45: 481‐497, 2006.
 276. Wittlich M, Koenig BW, Stoldt M, Schmidt H, Willbold D. NMR structural characterization of HIV‐1 virus protein U cytoplasmic domain in the presence of dodecylphosphatidylcholine micelles. FEBS J 276: 6560‐6575, 2009.
 277. Wittlich M, Koenig BW, Willbold D. Structural consequences of phosphorylation of two serine residues in the cytoplasmic domain of HIV‐1 VpU. J Pept Sci 14: 804‐810, 2008.
 278. Wittlich M, Thiagarajan P, Koenig BW, Hartmann R, Willbold D. NMR structure of the transmembrane and cytoplasmic domains of human CD4 in micelles. Biochim Biophys Acta 1798: 122‐127, 2010.
 279. Wuthrich K. Protein structure determination in solution by NMR spectroscopy. J Biol Chem 265: 22059‐22062, 1990.
 280. Xu C, Gagnon E, Call ME, Schnell JR, Schwieters CD, Carman CV, Chou JJ, Wucherpfennig KW. Regulation of T cell receptor activation by dynamic membrane binding of the CD3epsilon cytoplasmic tyrosine‐based motif. Cell 135: 702‐713, 2008.
 281. Yang J, Ma YQ, Page RC, Misra S, Plow EF, Qin J. Structure of an integrin alphaIIb beta3 transmembrane‐cytoplasmic heterocomplex provides insight into integrin activation. Proc Natl Acad Sci U S A 106: 17729‐17734, 2009.
 282. Yeagle PL, Alderfer JL, Albert AD. Structure of the third cytoplasmic loop of bovine rhodopsin. Biochemistry 34: 14621‐14625, 1995.
 283. Yildiz O, Vinothkumar KR, Goswami P, Kuhlbrandt W. Structure of the monomeric outer‐membrane porin OmpG in the open and closed conformation. EMBO J 25: 3702‐3713, 2006.
 284. Yu L, Sun C, Song D, Shen J, Xu N, Gunasekera A, Hajduk PJ, Olejniczak ET. Nuclear magnetic resonance structural studies of a potassium channel‐charybdotoxin complex. Biochemistry 44: 15834‐15841, 2005.
 285. Zangger K, Respondek M, Gobl C, Hohlweg W, Rasmussen K, Grampp G, Madl T. Positioning of micelle‐bound peptides by paramagnetic relaxation enhancements. J Phys Chem 113: 4400‐4406, 2009.
 286. Zhao H, Pykalainen A, Lappalainen P. I‐BAR domain proteins: Linking actin and plasma membrane dynamics. Curr Opin Cell Biol 23: 14‐21, 2011.
 287. Zheng H, Zhao J, Sheng W, Xie XQ. A transmembrane helix‐bundle from G‐protein coupled receptor CB2: Biosynthesis, purification, and NMR characterization. Biopolymers 83: 46‐61, 2006.
 288. Zhong L, Johnson WC Jr. Environment affects amino acid preference for secondary structure. Proc Natl Acad Sci U S A 89: 4462‐4465, 1992.
 289. 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 31: 896‐908, 2008.
 290. Zimmerman T, Oyarzabal J, Sebastian ES, Majumdar S, Tejo BA, Siahaan TJ, Blanco FJ. ICAM‐1 peptide inhibitors of T‐cell adhesion bind to the allosteric site of LFA‐1. An NMR characterization. Chem Biol Drug Des 70: 347‐353, 2007.
 291. Zoonens M, Catoire LJ, Giusti F, Popot JL. NMR study of a membrane protein in detergent‐free aqueous solution. Proc Natl Acad Sci U S A 102: 8893‐8898, 2005.

Related Articles:

Membrane Structure/Proteins
Organization and Dynamics of the Lipid Components of Biological Membranes

Contact Editor

Submit a note to the editor about this article by filling in the form below.

* Required Field

How to Cite

Heike I. Rösner, Birthe B. Kragelund. Structure and Dynamic Properties of Membrane Proteins using NMR. Compr Physiol 2012, 2: 1491-1539. doi: 10.1002/cphy.c110036