Abstract
Membrane proteins are one of the most challenging and attractive objects in modern structural biology, as they are targets for the majority of medicines. However, studies of membrane proteins are hindered by several obstacles, including their low ability to crystallize, highly dynamic behavior of some of their domains, and need for membrane-like environment. Although solution nuclear magnetic resonance (NMR) is a very powerful technique of structural biology in terms of the amount of provided data, it imposes several limitations on the object under investigation, with the main constraint being related to the size of the object. For this reason, the membrane mimetic has to form particles of small size and simultaneously to properly simulate the bilayer membrane to be applicable for solution NMR spectroscopy. Here we review the recent advances in the field of membrane mimetics for solution NMR studies, discuss the advantages and drawbacks of specific membrane-like environments, and formulate the criteria for the selection of proper environment for a particular membrane protein or domain.
Abbreviations: MP, membrane protein; NMR, nuclear magnetic resonance; TM, transmembrane; p75NTR, p75 neurotrophin receptor; DAGK, diacylglycerol kinase; LPN, lipid-protein nanodisc; SMA, styrene and maleic acid; SMALP, SMA-lipid particle; CMC, critical micelle concentration; TROSY, transverse relaxation optimized spectroscopy; DM, decyl maltoside; DDM, dodecyl maltoside; DPC, dodecylphosphocholine; LDAO, lauryldimethylamine-N-oxide; SDS, sodium dodecyl sulfate; LMPG, 1-myristoyl-sn-glycero-3-phospoglycerol; LPPG, 1-palmitoyl- sn-glycero-3-phospoglycerol; DMPC, 1,2-dimyristoyl-sn- glycero-3-phosphocholine; DHPC, 1,2-dihexanoyl-sn-glycero-3-phosphocholine; DH7PC, 1,2-diheptanoyl-sn-glycero- 3-phosphocholine; CHAPS, 3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonate; CHAPSO, 3-[(3-cholamidopropyl)dimethylammonio]-2-hydroxy-1-propanesulfonate; TFE, trifluoroethanol.
1 Introduction
Membrane proteins (MPs) are one of the most attractive objects in modern structural biology. A total of 20–30% of human genome open reading frames encode the MPs [1], and MPs represent the vast majority of drug targets [2]. MPs take part in the development of many severe diseases, including cancer, neurodegenerative and autoimmune disorders, pain syndromes, etc. Altogether, the listed facts highlight the importance of structural studies of MPs. Solving the spatial structures of such proteins would allow the deep understanding of the structure-function relationship for MPs, rational protein engineering, and drug design. By contrast, of 100,000 structures that are available in the Protein Data Bank, only 3% are annotated as MPs, implying that the MPs are underinvestigated from the structural viewpoint for several reasons. Many MPs, including all type I or bitopic proteins, are highly dynamic and often contain intrinsically disordered regions, and it prevents their crystallization and high-resolution studies by cryo-electron microscopy. Recombinant MPs are also very difficult to produce: yields in eukaryotic cells are extremely low [3] and refolding of MPs that are obtained in inclusion bodies of bacterial cells is not straightforward and is often a state-of-the-art task. Some MPs contain both the extracellular and the intracellular globular domains that require the different redox properties of the environment – cysteine residues are engaged in disulfide bridges outside the cell and are reduced in the cytoplasm. Last, but not the least, MPs require the specific environment to retain their native structure. Almost all conventional approaches of structural biology are not applicable in cells and even in liposomes. Therefore, the special membrane mimetics are necessary, which contain the unnatural components affecting the properties of protein under investigation, and this is especially important in the case of solution nuclear magnetic resonance (NMR) spectroscopy.
Solution NMR is one of the most powerful techniques of structural biology in terms of the amount of provided data. Apart from the determination of high-resolution spatial structures, NMR is used to study the intramolecular mobility of proteins, to monitor the conformational transitions, and to investigate the kinetic and thermodynamic parameters of various processes. However, the wide spread of NMR in structural studies is limited because of the several experimental problems. The major limiting restraint of NMR spectroscopy is the size of the object under investigation. Large molecules tumble slowly in solution, which results in the enhanced transverse relaxation, broad lines, loss of sensitivity, and resolution in NMR spectra. In addition, large molecules contain many nuclei that give rise to the signals in NMR spectra, which, in turn, becomes overcrowded and hard to interpret. This was in part overcome by the recent advances. Transverse relaxation optimized spectroscopy (TROSY) pulse sequences that were developed for the aromatic [4], amide [5], [6], and methyl [7] groups allow to decelerate the transverse relaxation and enhance the sensitivity, whereas the novel techniques of specific labeling of protein methyl groups and other moieties [8], [9], [10], [11], [12], [13] simplify the analysis of NMR spectra and abolish the dipole-dipole interactions between protons, which contribute a lot to the transverse relaxation. However, the size of molecules/molecular complexes studied by NMR in solution rarely exceeds 50–70 kDa. Investigation of larger objects is usually a state-of-the-art work [14], [15] and requires the great time and money expenses. Therefore, if an MP needs to be studied by NMR in solution, the membrane mimetic has to form particles that are relatively small and simultaneously be alike the lipid bilayer to adequately model the properties of the real cell membrane. Conventional membrane mimetics that are applicable for solution NMR spectroscopy are well described in several recent reviews [16], [17], [18], [19], [20], [21], [22], [23], [24], [25], [26]. For that reason, we will give a brief overview on the types of available membrane mimetics (Figure 1), including the most recent data, and then focus on the main problem of NMR studies of MPs – approaches to the rational selection and optimization of a membrane-like environment for the particular protein.

Shape and architecture of particles of membrane mimetics, applicable for the solution NMR spectroscopy. Orange cylinders represent the MSP or other belt-forming protein, blue band is the SMA molecule, and green ribbon is the amphipol. Gray are the molecules of detergent and lipids are shown in yellow.
2 General types of membrane mimetics for solution NMR studies
2.1 Organic solvents
MPs are usually not soluble in water because of the presence of large hydrophobic regions. One of the strategies to shield the hydrophobic parts of the MP is to add up to 100% of one or mixture of organic solvents such as methanol, ethanol, isopropanol, trifluoroethanol (TFE), chloroform, dimethyl sulfoxide, etc. For example, fragments of bacteriorhodopsin were studied in organic solvents by solution NMR [27], [28], [29], and the mixture of chloroform-methanol-water (4:4:1) was shown to mimic the membrane properties for the transmembrane (TM) H+-transporting subunit c of the F1Fo ATP synthase [30]. Although MPs usually adopt the proper secondary structure in such mixtures, the tertiary structure is not formed because of the absence of the expressed interface between the polar and the nonpolar portions of the solution. Therefore, the use of organic solvents is nowadays restricted to the studies of secondary structure of the single-TM or two-span helical proteins [31], [32], [33] and small membrane-active peptides [34], [35].
2.2 Detergents
Detergents are historically the first membrane mimetic, which is indeed membrane-like. One of the most important characteristics of detergents is the critical micelle concentration (CMC). Below the CMC, detergents are soluble in water, whereas above the CMC, the amphipathic properties of detergent molecules cause the formation of aggregates – detergent micelles with the hydrophobic core and hydrophilic outlet. Detergent micelles have a distinct border between the polar and the hydrophobic compartments, which makes the various parts of MPs to interact with one another within the micelle and to form the tertiary structure. Detergents are traditionally called “harsh” and “mild” based on their ability to denature the MPs [36]. Harsh detergents are commonly ionic and are used to dissolve the bacteria inclusion bodies or protein precipitates of other kinds, whereas mild detergents are uncharged, sometimes bear the hydrocarbon moieties, and are used to extract the proteins from the membrane, retaining their native structure and activity. The properties of lipid packing inside the particles and the curved shape of the micelle surface are quite far away from the characteristics of real cell membrane. This may cause the improper folding of the MP and is definitely a disadvantage of this membrane mimetic. By contrast, micelles have a relatively small size (20–100 kDa), which is extremely important for solution NMR spectroscopy. Besides, many detergents are now available in the deuterated form. As we will show in the next section, MPs may retain their native structure in a specific detergent or a mixture of detergents, which makes micelles the most widely used membrane mimetic for the solution NMR studies.
On the dawn of the solution NMR studies, MPs were investigated mainly in harsh detergents, such as sodium dodecyl sulfate (SDS) [37], [38], [39], and even now some studies are performed in this mimetic [40], [41], [42], [43], [44], [45], [46]. Despite the variety of the detergents that are commercially available, only few are used in solution NMR studies and in mimetic screenings (Figure 1). Very mild detergents, decyl maltoside (DM) and dodecyl maltoside (DDM), can be used to extract the proteins from the membranes in the active form and are taken to investigate the 7-TM proteins, such as bacteriorhodopsin [47], [48] and G-protein-coupled receptors (GPCRs) [49], [50], [51], and other helical MPs, e.g. voltage-gated channels [52]. GPCRs are also active in the mixtures of DM and DDM with cholesterol hemisuccinate [53], [54]. These mimetics are known to support the native folding of many proteins but form very large micelles (~70 kDa), which prevents the high-resolution studies in such an environment. Dodecylphosphocholine (DPC) and lauryldimethylamine-N-oxide (LDAO) are harsh detergents with small micelles (20–25 kDa) that often maintain the native structure of MPs and provide the good quality of NMR spectra [55], [56], [57], [58], [59], [60], [61], [62], [63], [64], [65], [66], [67], [68], [69], [70], [71]. In some cases, the nonconventional analogs of DPC with methylated and hydroxylated fatty chains or with the altered number of carbon atoms, such as FOS-30, FOS-10, or FOS-14, reveal the better performance to dissolve the MPs [72], [73], [74], [75]. Anionic lysolipids (1-myristoyl-sn-glycero-3-phospoglycerol [LMPG] and 1-palmitoyl-sn-glycero-3-phospoglycerol [LPPG]) can solubilize the proteins directly from the cell-free reaction precipitates and are often used in NMR studies [32], [76], [77], [78], [79]; however, they were shown to cause the improper folding and inactivation of some MPs [19], [80]. In addition, recent studies revealed the prospects of the unnatural micelle-forming short-chain lipid 1,2-diheptanoyl-sn-glycero-3-phosphocholine (DH7PC) as a membrane mimetic for various MPs, including 7-TM [81], [82] and other α-helical proteins [83]. Similar short-chain lipid 1,2-dihexanoyl-sn-glycero-3-phosphocholine (DHPC) was often used in the studies of β-barrel bacterial outer MPs [84], [85], [86], [87], [88]. In many cases, the best results are obtained in mixed micelles, where the detergents with different length of fatty tails [58], [89], [90], [91] and/or charge on the headgroups are combined together [60], [65], [92], [93]. Detergents from the Brij and Tween families and Triton X-100, which are conventionally used for the extraction of MPs from the cell membranes, were never shown to provide the NMR spectra of MPs of reasonable quality [76], [94]. Similarly, no example of the successful use of the bile salt derivatives in micelles, such as 3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonate (CHAPS) and 3-[(3-cholamidopropyl)dimethylammonio]-2-hydroxy-1-propanesulfonate (CHAPSO), for solution NMR studies of MPs is reported [76]. The use of various detergents for the structure determination by solution NMR is summarized in Figure 2, and properties of common detergents are well described in the review [22].

Membrane mimetics for NMR structure determination in solution. Shown is the number of structures in Protein Data Bank (PDB) database determined in various membrane mimetics since 2010. A total of 114 spatial structures were gathered from the websites http://blanco.biomol.uci.edu/mpstruc/ and http://www.drorlist.com/nmr/MPNMR.html and manually found in PDB database among the entries that are annotated as “membrane protein” and are not mentioned on both websites. Left histogram describes the distribution of found structures between the general types of membrane mimetics, blue sector corresponds to the organic solvents. Right histogram describes the use of particular detergents in micelles. FOS-10 (2 structures), FOS-14 (1 structure), and FOS-30 (1 structure) are counted together with DPC. Blue sector corresponds to the mixtures of DPC with other detergent (SDS) or phospholipid.
2.3 Bicelles
Bicelles are one of the most promising mimetic to study the MPs by solution NMR. Bicelles contain the patch of a planar lipid bilayer surrounded by the rim of the detergent [95], [96]. Lipids with various length and saturation of fatty chains [97], [98], [99], headgroups [100], [101], [102], [103], [104], cholesterol [105], and gangliosides [106] were shown to be capable of bicelle formation alone or in the mixtures with other lipids. This makes bicelles a convenient environment to study the effect of the membrane lipid composition on the structural properties of the MPs and to investigate the specific lipid/protein interactions [107], [108]. In addition, lipid analogs with the ester bonds can be used instead of the conventional phospholipids, to exclude the lipid hydrolysis that can occur at low pH [109]. It is also known that any arbitrary detergent would not always form bicelles, being mixed with lipids. The ability to assemble into the discoidal particles was documented for the DHPC [110], [111], bile salt [112], and its derivatives CHAPS, CHAPSO, and Façade detergents [99], [113], [114]. DH7PC/1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) solution was also called “bicelles” in the NMR structural studies [115]; however, the shape of particles in such a mixture was not characterized. Rim-forming detergents can be polymerized in bicelles to enhance the stability of the MP under investigation [116]. Out of all detergents, CHAPS and CHAPSO are preferred if the MP is sensitive to the presence of the detergents, and DHPC should be selected if the deuteration of fatty chains is required for the needs of the experiment [99].
The size of bicelles can be controlled in quite a wide range, starting from approximately 40 kDa. By varying the lipid/detergent ratio (q), the character of the size dependence is well established for many rim-forming agents [95], [99], [117]. However, it was shown that bicelles increase their size upon dilution and heating [99], [118], [119], [120]. The first effect can be excluded keeping in mind that a fraction of the detergent is soluble in the monomeric form [99], whereas the temperature-dependent growth is not observed for bicelles with radii less than 3 nm [99], [119]. Bicelles formed with large q orient spontaneously in strong magnetic fields and are utilized in solid-state NMR spectroscopy of MPs [121] and in solution to measure the residual dipolar couplings of soluble proteins [122]. Small bicelles that are used in solution NMR studies of MPs are not oriented by the magnetic field and are called “isotropic”. Lipids in larger bicelles (q>0.75 for DMPC/DHPC) can undergo the phase transition at temperatures, close to the observed for the lipid bilayers [118]. However, the correspondence of the lipid packing parameters to the real bilayer membranes is not established, the mobility of the MPs is enhanced in some kinds of bicelles, and packing of lipids around the protein is not tight [98], [99].
The benefits of bicelles are obvious – they contain the portion of lipid bilayer, can mimic the lipid composition of the cell membrane, and retain the activity and native structure of many MPs [19], [115], [123]. However, the use of bicelles in solution NMR is limited – spatial structure of one β-barrel [96] protein, several dimers of single-TM α-helices [124], [125], [126], [127], [128], [129], [130], [131], [132], ArfGTP [133], complex of two cytochrome P-450 subunits [134], and Smr [135] were determined or characterized in this mimetic (Figure 2). In the most recent study, bicelles were successfully utilized to reconstruct and determine the spatial structure of the HIV Env trimer [136]. Altogether, less than 10% of NMR spatial structures of MPs were obtained in bicelles (Figure 2). It may be the consequence of the relatively large minimal size of bicelles and difficulties with the transfer of the MP of interest into bicelles from the detergent that was initially used to extract the MP from cell membranes or to solubilize the MP from the inclusion bodies or cell-free precipitates.
2.4 Lipid-protein nanodiscs
Lipid-protein nanodiscs (LPNs) are, like bicelles, the mimetic that is extremely membrane-like. LPNs were also shown to contain a patch of lipid bilayer surrounded by the belt formed by special proteins. Several belt proteins were suggested for LPNs [137], [138], [139], the most widely spread originate from the membrane scaffold protein (MSP), which is a part of the apolipoprotein A-I. MSP consists of amphipathic a-helices, the nonpolar side of the helices covering the hydrophobic acyl chains of the lipids. The radius of the classic MSP1 LPNs is equal to approximately 5 nm [140], [141]; however, recently the size of LPNs was reduced to 3.5–4.0 nm after the development of shorter MSP versions [55]. LPNs undergo the phase transition at temperature close to the critical temperatures measured for lipid bilayers [140]; however, lipids are packed more tightly in LPNs than in liposomes [141], [142], [143] and in bicelles [99], suggesting that the state of the lipid bilayer in LPNs does not correspond to the liquid-crystalline membrane. The most useful advantages of LPNs are their homogeneity and monodispersity, stability to disruption and aggregation, ability for an experimenter to choose among the different lipid compositions to mimic the native membrane, and ability to obtain the suitable thickness of the lipid bilayer [144]. These special properties of the LPNs enable to maintain the stability and integrity of the protein under investigation and reach its high concentrations required for the structural studies by NMR. Compared with bicelles, LPNs are characterized by the absence of detergents, discrete set of possible sizes, and prohibition against the matter exchange between the particles. This prevents the oligomerization of MPs under investigation but makes the studies of interaction between MPs in LPNs impossible.
In the initial works, LPNs were used to study the topology of the membrane-active peptides on the membrane [145], [146], [147] and as a reference medium in the detergent screenings [80], [115], [148], [149], [150]. The principal possibility of NMR structural studies in large MSP-based LPNs was demonstrated by the works with VDAC-1 channel and CD4 protein [151], [152], and the introduction of smaller LPNs resulted in the structure determination of two β-barrels [153], [154], [155] and of the construct, containing the TM and cytoplasmic domains of the p75 neurotrophin receptor (p75NTR) [156]. LPNs formed by the other belt proteins, such as 22A peptide, were also shown to be applicable for NMR structural studies in solution: the complex of cytochrome b5 and cytochrome P-450 was reconstructed in these LPNs, and high-resolution TROSY spectrum was obtained [157]. The examples of LPN implementations in other fields of structural biology are beyond the scope of this article and can be found in the excellent reviews [20], [158].
Despite many benefits, LPNs have several considerable drawbacks that restrict their applicability in solution NMR studies. The mass of the smallest LPNs composed of truncated variants of MSP with the embedded protein is approximately 60 kDa, which is far too big for conventional NMR techniques and results in the NMR spectra of poor quality for the majority of the MPs [149]. One should use the deuteration of the target protein accompanied by the specific labeling techniques, which, in turn, significantly elaborates the expression protocols and increases the sample cost.
2.5 SMA-lipid particles
The next membrane mimetic is very much like LPNs and contains the portion of lipid bilayer surrounded by the styrene and maleic acid (SMA) copolymer (3:1) [159]. SMA is an amphiphilic molecule capable of forming relatively small disk-shaped nanoparticles from lipid vesicles. This novel membrane mimetic was called SMA/lipid particles or SMA-lipid particles (SMALPs). SMALPs were used to solubilize the 7-TM α-helical bacteriorhodopsin [159], [160], the 8-stranded b-barrel lipid A palmitoyltransferase PagP [159], and the function modulator of voltage-gated potassium channels protein KCNE1 [161]. MPs were reported to retain their integrity, stability, and function if incorporated into SMALPs. A particle with an embedded 7-TM protein is ~11 nm in diameter and contains approximately 11 lipid molecules [159]. Thus, SMALPs are monodispersed, thermostable, and soluble nanoparticles applicable for solution NMR investigations. The size of the nanoparticle could be easily controlled by manipulation of the lipid–SMA polymer ratio, as was demonstrated in [162], [163]. One of the most prominent features of SMALPs is their potential ability to solubilize the integral MPs form lipid vesicles or membranes not using the detergents. By contrast, SMALPs have their obvious drawbacks – the minimal size of the SMALP particle is too large and the amount of lipids inside the SMALP is too low because of the high volume of the SMA chains. For unknown reasons, the use of SMALPs in solution NMR studies was not reported; however, this application of the mimetic needs to be tested in the nearest future.
2.6 Amphipols
Amphipols stay separately in this list of membrane mimetics because they do not contain any lipids or detergents. Amphipols are amphipathic polymers that form the coat around MPs with the thickness of 1.5–2.0 nm [164]. The mimetic is characterized by high propensity to stabilize the MP and to prevent its oligomerization [165]. The prospects of amphipols for structural NMR studies is discussed in the recent review [166]. In brief, five MPs, both α-helical and β-structured, were characterized in amphipols to the date by solution NMR spectroscopy [48], [167], [168], [169], [170], [171], [172], [173], [174], [175], [176]. These studies confirm the ability of some kinds of amphipols to retain the native structure and activity of MPs and simultaneously to provide the high-resolution NMR spectra of their TM domains, with the quality, comparable with observed in detergents and bicelles. Some amphipols are available in deuterated form, which is also advantageous in NMR structural studies of MPs in solution.
Amphipols cannot be a priori deemed to properly mimic the lipid bilayer because of their unnatural composition; therefore, the biological relevance of the spatial structures obtained in amphipols will always be questioned in the absence of protein activity or other data confirming the native folding of the protein under investigation. However, this also refers to the detergent micelles. It is also noteworthy that the amphipol/MP particles are not monodisperse, and a fraction of the sample needs to be selected to obtain the spectrum of high quality, thus decreasing the effective yield of the protein production [173]. In addition, the use of amphipols is not a universal solution for any arbitrary chosen MP. There are negative results in the literature, when certain MPs could not be reconstituted in amphipols or no NMR spectra of reasonable quality were observed for the MP in amphipol environment [173]. Thus, amphipols may be considered as an alternative to LPNs and other lipid-based mimetics when the MPs of interest are prone to aggregation and inactivation.
3 Experimental protocols for the solubilization of MPs in membrane-like environment
Despite the variety of available membrane mimetics, there are only few approaches to solubilize the MPs. In the most fortunate case, the protein precipitates can be dissolved directly by the aqueous solution of membrane mimetic, such as bicelles or detergents [60], [77]. When the α-helical membrane domain is soluble in the mixture of organic solvents, usually TFE/H2O or hexafluoroisopropanol/H2O, the dry powder of lipids or detergents is simply added to such solution. Then the water is added to the mixture until micelles or bicelles are formed, and the solution is lyophilized and redissolved in the aqueous buffer [31], [58]. This option is applicable mainly for the most simple MPs, such as single-helix TM domains of bitopic proteins. MPs can also be extracted by the mild detergents directly from the cell membrane [49] or cotranslationally incorporated into the particles of membrane mimetic during the cell-free reaction [94], [177], [178], [179]. The protein of interest, which is already solubilized by the harsh detergent (SDS, LS) or extracted from the membrane, can be transferred to the desired environment using the affine chromatography. MP is usually immobilized on the wax, such as Ni-Sepharose, and then washed by the solution containing the membrane mimetic of choice [180]. Alternatively, mild detergents with low CMC can be exchanged with the detergents of higher solubility using permeable membrane filtering units [181]. LPNs, SMALPs, and amphipols are prepared from the lipid/detergent solutions of the MP under investigation. After the addition of the MSP, SMA, or amphipol, the detergent is removed either by dialysis or by using a special wax capable of absorbing the small hydrophobic molecules [137], [159], [182].
4 Possible criteria for the selection of membrane mimetic, rational approach
The variety of available membrane mimetics raises one of the most essential problems of MP structural studies – the correct and the rational choice of the membrane-like environment for the protein under investigation. In the first decade of the implementation of solution NMR spectroscopy for protein structure determination, the isotope labeling and recombinant protein production were extremely expensive. Therefore, all studies of MPs were performed in the mixtures of organic solvents or in the only cheap detergent that was available in the deuterated form. This also refers to the first solved structure of the MP – gramicidin A [37]. Scientists of those days did not screen the detergents and did not test the activity of the MPs, they just took the only option they had and performed studies in the only model that was available, regardless the possibly nonnative character of the environment. However, now, when isotope labeling is a routine task, one has to formulate the criteria to select the membrane mimetic out of the vast variety of options. The solution may seem simple – choose the environment that is at most like the bilayer membrane, e.g. LPNs or bicelles. However, there are several obstacles on that path. First, as we will show in the next section, in some cases the use of LPNs or bicelles does not result in the proper folding of the MP. Second, both bicelles and LPNs are characterized by the large size of particles in solution, which does not allow the straightforward structure determination. In large particles, the signals in NMR spectra cannot be assigned using the conventional triple-resonance experiments, and the expensive procedure of the single-point scanning mutagenesis has to be applied for the task [49]. Besides, because the studies of large objects require the deuteration of protein side chains, the spatial structure of the MPs in isotropic bicelles and LPNs cannot be resolved through the conventional NOE approach, the insensitive four-dimensional spectra [57], paramagnetic labeling [77], [183], and use of anisotropic environment are required [184], [185]. Therefore, LPNs, large bicelles, SMALPs, and amphipols are the weapons of the “last chance” and are applied for the structure determination by NMR when other mimetics fail to support the stability, activity, and native folding of the MP under investigation. One would usually like to work in micelles or very small bicelles that do not contain the significant patch of planar bilayer to minimize both time and money expenses. In the next paragraph, we list the possible approaches to select the best detergent for the structural study of the MP under investigation.
The easiest way to screen the membrane mimetics is to monitor the activity of the MP of interest. This can be done for photoactive proteins, such as bacteriorhodopsin, optical rhodopsin, and other similar MPs based on the light absorption spectra [48], [51], [164], [181]. The functionality of the GPCR may be assessed using their ligand/G-protein binding propensity in the model environment [49], [50]. The active state of KcsA potassium channel in SDS was confirmed by the presence of functional tetrameric assembly of the protein in solution [41]. However, these are the exceptional cases; the activity of other MPs cannot be studied that easily. Type I integral MPs, such as receptor tyrosine kinases, require the whole protein to monitor its activity, which is far too large for the NMR studies. The conductance of ion channels cannot be measured in any of the mimetics that are applicable for the solution NMR spectroscopy. Moreover, the absence of the ligand binding or other activity of an MP does not always indicate the improper folding of the protein. There may be subtle differences in the structure that do not affect the overall fold but do prevent the interaction with the ligand. Thus, other criteria apart from the activity of the protein need to be applied for the screening of the membrane mimetics.
The most feasible way to estimate the correctness of the MP folding is to obtain these data directly from the NMR spectra. Such an approach is implemented in the majority of screenings of the membrane mimetics. First, vast detergent screenings relied on the so-called quality of NMR spectra. Krueger-Koplin et al. [76] performed a very wide detergent screening for three different helical MPs in 2004, using dozens of various mimetics. As a merit of the “quality”, they used the number of observed signals in NMR spectra and the lifetime of the sample. The reasonable lifetime is necessary to record the long-term NMR spectra, whereas the number of peaks reports on the internal mobility of the MP and determines whether the protein can be investigated in this mimetic. On the basis of the stated criteria, the authors found the anionic lysolipids, such as LPPG, to be the optimal for all three objects. However, nice spectra with narrow peaks of predicted number and long lifetime of the sample mean that the protein does not precipitate in the mimetic of interest and does not experience the slow motions on the microsecond-to-millisecond timescale but says nothing about the folding. In the later work, Girvin et al. [19], [123] found that Smr binds its ligands only in DMPC/DHPC bicelles, but not in the LPPG, which was, this part refers to LPPG selected for the protein in the detergent screening. For that reason, the dispersion of signals in NMR spectra also needs to be estimated, especially for the helical MPs. If the folding is incorrect, and TM helices are not in the tight contact, the dispersion of signals is low, whereas in case of the tight packing of helices with specific contacts, outlying cross peaks should appear. Such a criterion was used by Zhang et al. [72] in their broad screening of membrane mimetics for the OmpX protein. In addition, the use of the signal dispersion was demonstrated clearly on the example of the voltage sensor of the KvAp potassium channel [80]. Although very good spectra were obtained for the protein in anionic lysolipids, the proper folding of the sensor was observed only in the zwitterionic DPC and DPC/LDAO mixture, where the “quality” of spectrum was lower but the dispersion of signals was much higher.
Nonetheless, even the high dispersion of signals does not ensure the proper folding of the MP. Recently, with the introduction of LPNs, it became possible to measure the NMR spectra in the most native environment, provided by the LPNs, and select the micelles or bicelles based on the similarity of NMR spectra in mimetics with small particles and in bilayer-containing system. This approach was first suggested by Shenkarev et al. in 2010 for the KvAp voltage sensor, and since then, it is actively used in screenings [48], [80], [115], [148]. Similarly, if the MP contains soluble domains, one can record the NMR spectra of separate globular domain and compare them with the spectra of the full-size protein in various membrane mimetics [148], [156].
All these listed techniques use the screening approach, but it would be useful to select rationally the membrane mimetic for the specific protein. To do this, one needs to understand clearly what the variables of the membrane mimetics are and how they influence the quality of NMR spectra and folding of the MP under investigation. The work by Columbus et al. [89] demonstrated that the hydrophobic thickness of membrane mimetics is one of the main determinants of the protein structure and behavior. They took the detergents of various nature and charge to dissolve the model two-helical protein, TM0026. As a result, the appearance of NMR spectra of the protein was shown to be dependent mainly on the effective hydrophobic length of detergent fatty chains; mixtures of two different detergents yielded spectra identical to the other detergents with the hydrophobic length, corresponding to the average length of mixed components. Thus, instead of the wide screening, one could choose first the short-chain detergent and then titrate the sample by the long-chain molecules until the quality of spectra is the best. This principle was utilized in the studies of BniP3 mitochondrial protein: titration of the initial sample in DPC by the long-chain phospholipid DPPC allowed to exclude the unwanted minor conformation of the protein [58]. Mixtures of detergents of various lengths (LMPG/DH7PC, DPC/FOS-16) were used in the screening for the best conditions for the YgaP bacterial protein [90], [91], [148] and to study the effect of the hydrophobic mismatch on the spatial structure of the TLR3 dimeric TM domain [186]. Apart from the length of the detergent chains, charge on the headgroup is also an important factor. For many proteins, the best functionality can be obtained in nonionic detergents, such as DM, DDM, and Facade-EM [48], [49], [181], but in some cases, the addition of charged molecules to the zwitterionic micelles may significantly improve the stability of sample and quality of NMR spectra. In particular, the addition of small amounts of anionic SDS to DPC micelles improved the spectra of the FGFR3 TM domain [65], zz dimer [92], and DAP12 complex [93]. Moreover, the mixture of cationic LDAO with DPC provided the best quality of NMR spectra and the proper folding of the KvAp voltage sensor [80]. It is also necessary to keep in mind the overall concentration of membrane mimetic and lipid-to-protein ratio. At low lipid-to-protein ratio, MPs may start to self-associate, which results in the signal broadening, presence of several sets of signals, and worse overall quality of spectra [187]. By contrast, at a high concentration of membrane mimetic, the rotational diffusion of protein is decelerated, which also leads to the loss of sensitivity and broad signals in NMR spectra [187]. Therefore, the equilibrium between detergent/lipid and protein concentrations needs to be maintained. The size of LPNs and bicelles is also an important parameter for the optimization. Particles of large size decrease the quality of NMR spectra because of the enhanced transverse relaxation, whereas in small particles, the amount of lipids may not be enough to surround the MP, which causes the protein/rim interactions and results in the NMR spectra of poor quality [55]. With all aforesaid, it is clear that the rational selection and “fine tuning” of the membrane mimetic is not a dream and can be performed in the present state of things in solution NMR spectroscopy. All possible approaches and criteria to select the proper environment for an MP of interest are summarized on Figure 3.

Flowchart, describing the approaches and criteria used to find the proper membrane-like mimetic for an MP of interest.
5 Effect of the mimetic on the structure of integral MPs
To estimate the relevance of data and to interpret the results obtained in solution NMR studies of MPs in membrane mimetics, it is necessary to understand the influence of membrane-like environment on the spatial structure, dynamic behavior, and activity of MPs. On the basis of our experience, every second publication of NMR spatial structure determined in micelles meets the criticism of the reviewers, which is focused on the model character and nonnative properties of the mimetic. By contrast, X-ray structures of MPs are considered as etalons and are thought to correspond to the real state of things in the cell membrane. In addition it is thought that the structure obtained in micelles is always less relevant than the structure of the same protein, which was determined in the environment of bicelles and LPNs. However, both statements are incorrect. We need to point out that all contemporary spatial structures of MPs are determined in the model environment – micelles, bicelles, LPNs, crystals in case of X-ray investigations, and liposomes in case of solid-state NMR studies. All of the listed mimetics are quite far away from the cell membrane, which is especially obvious in case of crystals that contain the negligible amounts of lipids. The exact environment of a protein inside the cell membrane is never known. There is a lot of data regarding the lipid composition of inner and outer leaflets of cell membranes of various kinds [21], but the real membrane has a mosaic structure [188] and contains various microdomains with the special “liquid-ordered” phase of lipid bilayer. Many MPs are active inside such domains; in some cases, the migration between the liquid membrane and the ordered microdomains may be a part of the activation mechanism [189]. Membrane microdomains are thicker than the liquid membrane, have a specific composition, and are usually enriched by sphingolipids and cholesterol [190]. The environment of the membrane microdomain cannot be reproduced by any mimetic with small particles, and even cannot be modeled adequately in liposomes. The real membrane is also rich in surface-associated proteins that can distort the bilayer. Besides, the MP also has its own influence on the surrounding lipids: it can recruit and bind the specific lipid molecules and affect the thickness and packing of lipids in the membrane [191], [192]. Thus, the use of various mimetics may sample the different states of the cell membrane and allow obtaining the different functional states of the MP (i.e. active, inactive, transition state, folding intermediate, etc.). Micelles are believed to adapt their shape to the protein under investigation and, in some cases, may provide even a more physiological environment than bilayer-containing particles because the choice of lipids with the incorrect length of fatty chains in bicelles or LPNs may affect the structure of the MP. The possibility of such errors was illustrated by the recent study of the M2 protein from the influenza virus, which adopted different conformations, depending on the thickness and charge of the bilayer [193]. It is essential to understand what kind of distortion can be introduced by the detergents into the spatial structure of the MP under investigation, to use micelles with necessary care, and to ensure the native state of the protein in such an environment. Several recent studies shed light on the effect of membrane mimetics and crystallization on the spatial structure of various proteins. These studies may help us to understand the influence of the mimetic on the structure of MPs and to judge on the relevance of the obtained structures.
At most, the effect of detergents is pronounced in the case of peripheral MPs, juxtamembrane regions, and water-soluble domains of MPs. The HIV-1 membrane-binding envelope protein was shown to form the curved helix in DHPC micelles and straight helix in DMPC/DHPC bicelles even at q as low as 0.25 [194]. This allowed authors to conclude that the absence of planar bilayer in detergent micelles can distort the structure of protein associated with the membrane surface. On the contrary, the juxtamembrane JMA regions of the human EGFR and HER2 receptors, which are believed to be helical in the context of full-length proteins, formed short amphipathic α-helices in the environment of DPC micelles and were highly mobile and disordered inside the bicelles of various size and composition [131], [195], [196]. Thus, in these particular cases, micelles could provide the more native environment than bicelles because of their ability to adapt the shape of particles to the properties of the incorporated protein. Harsh detergents that are used in both micelles and bicelles often cause the improper folding of the soluble domains of various MPs. The rhodanese domain of the YgaP protein appeared to be misfolded in DPC and DH7PC micelles, whereas the correct structure of the domain was observed in the LMPG/DH7PC mixtures at low excess of the detergents and in DMPC/MSP LPNs [148]. Similarly, the soluble “death domain” of the p75NTR was not folded properly in DPC micelles and DMPC/DHPC bicelles, whereas in various LPNs and in DMPC/CHAPS mixtures, the conformation of the domain was indistinguishable from the structure determined for the isolated “death domain” in solution [156]. Thus, micelles and harsh detergents should be applied very carefully for the MPs with structured juxtamembrane regions and globular cytoplasmic or extracellular domains. LPNs and bicelles containing the mild detergents, such as CHAPS and CHAPSO, should be considered instead.
The slightly different picture is observed for the integral helical and β-barrel proteins without the structured extramembrane domains. The “canonic” integral MP, glycophorin A, a strongly dimeric single-span protein, was investigated in a variety of mimetics, including DPC micelles [56], DMPC/DHPC bicelles [197], bilayers [198], and crystals [199], and no substantial differences between the determined spatial structures was observed (Figure 4A). Conformation of the similar strong dimer of the single-helix mitochondrial Bnip3 protein was determined in bicelles and DPC/DPPC mixture [58], [124] and appeared to be identical within the experimental error. By contrast, the weakly dimerizing TM domain of HER2 receptor adopted completely different conformations in DMPC/DHPC bicelles and DPC micelles [126], [196]. Thus, if the energy of interactions that promote the folding of the helical MP is high enough, the protein structure will not be dependent on the environment, whereas for many proteins, the change in the nature of the membrane mimetic and even the altered length of lipids in the bilayer may lead to the changes in spatial structure. At most instructive are the studies of the voltage sensor of KvAp voltage-gated channel. The protein was misfolded in lysolipids, such as LPPG, whereas the native structure of the domain was observed in LPNs and, surprisingly, in completely unnatural detergents, such as the short-chain lipid DH7PC [83] and cationic detergent LDAO [60]. The spatial structure of the protein in detergents was almost identical to the structure observed for the protein, crystallized from the β-octyl-glycoside in complex with the antibody fragment [200]. This example demonstrates the falseness of the assumption that the proper structure is always adopted by an MP in the environment that we ourselves deem as more native. LPPG and LMPG much more resemble the lipids of the real cell membrane than DH7PC and LDAO but do not support the folding of the voltage sensor. In the past few years, there appeared several other studies comparing the structure of the MP in detergents to the conformation of the protein in bilayer-containing mimetics. All these studies reveal that for many proteins, there is a micellar environment that supports the native folding of the membrane domain. In particular, “unnatural” LDAO micelles were shown to maintain the native state of the BamA insertase as well as DMPC/DH7PC mixtures and LPNs [115]. Bacteriorhodopsin, which is inactive in the majority of detergents, folds properly in DDM micelles as well as in the amphipols and DMPC/MSP LPNs [48]. Finally, KcSA ion channel can be assembled into the tetrameric complex in SDS micelles [41].
![Figure 4: Spatial structures of MPs in various membrane mimetics. (A) Spatial structures of glycophorin A determined in DPC micelles [56], [197], DMPC/DHPC q=0.25 bicelles [197], and in lipidic cubic phase [199]. The structure in lipid bilayers [198] was not deposited to the PDB and is therefore not shown. (B) Spatial structures of OmpX determined in DHPC [85] and DPC [55] micelles, LPNs [153], and crystals with n-octyltetraoxyethylene [201].](/document/doi/10.1515/ntrev-2016-0074/asset/graphic/j_ntrev-2016-0074_fig_004.jpg)
Spatial structures of MPs in various membrane mimetics. (A) Spatial structures of glycophorin A determined in DPC micelles [56], [197], DMPC/DHPC q=0.25 bicelles [197], and in lipidic cubic phase [199]. The structure in lipid bilayers [198] was not deposited to the PDB and is therefore not shown. (B) Spatial structures of OmpX determined in DHPC [85] and DPC [55] micelles, LPNs [153], and crystals with n-octyltetraoxyethylene [201].
Another instructive example is provided by the studies of the β-barrel bacterial outer MP OmpX (Figure 4B). The spatial structure of OmpX was determined in two different detergents: DHPC [85] and DPC [55], small DMPC/MSP LPNs [153], and crystals with n-octyltetraoxyethylene [201]. Assuming that the correct fold of the protein is observed in LPNs, both detergents and crystallization affect the folding of the protein. DHPC and, to a certain extent, DPC disturb the β-sheets, strands become shorter, and interstrand loops become longer than in LPNs. By contrast, crystallization reveals the adverse effect – the OmpX structure is stabilized excessively; all interstrand loops that are mobile in solution become fixed and are now parts of the β-sheet structure. These data demonstrated clearly that the X-ray structures may not be considered as a perfect etalon. Crystallization results in the unnatural protein-protein contacts that affect the mobile and unstructured regions of the protein, making them more rigid and even resulting in the formation of the regular secondary structure. For that reason, it is still unknown whether the NMR-derived structure in DPC micelles [63] or the X-ray structure [202] correspond to the native state of the diacylglycerol kinase (DAGK) – the structures are completely different, but the nature of the observed discrepancy is unknown. Moreover, the solid-state NMR study in real Escherichia coli membranes [203] reveals the secondary structure of DAGK, which is different substantially from both the X-ray and solution NMR data.
6 Conclusions
To sum up, here we review the variety of membrane mimetics, applicable for the solution NMR spectroscopy. We formulate the criteria that are used to select the appropriate environment for the MP in mimetic screening and show that the rational approach to the selection and adjustment of the mimetic is possible. On the basis of the recent structural studies of same MPs in different environment, we suggest that none of the structural data should be approached with prejudice, regardless of the nature of membrane-like environment. If due care is taken, the use of detergent micelles can result in the native folding of the MP. By contrast, the protein may be folded improperly even in large bilayer-containing particles. For that reason, all spatial structures of MPs obtained in solution either by NMR or by Cryo-EM need to be validated with functional assays, mutagenesis, or other independent experiments. However, this also refers to the X-ray structures: crystallization may as likely disturb the native folding of the MP as the use of other model membrane-like media.
About the authors

Konstantin S. Mineev is a senior research scientist in the laboratory of biomolecular NMR spectroscopy at the Shemyakin-Ovchnnikov Institute of Bioorganic Chemistry, Moscow. He received his master’s degree at the Moscow Institute of Physics and Technology (MIPT) in 2007 and his PhD in Biophysics at the Lomonosov Moscow State University in 2010. He is working as an adjunct professor in MIPT since 2014. The current research of Dr. Mineev is focused on elucidating the structural basis of the activation mechanisms of type I integral membrane proteins and development and characterization of various membrane mimetics.

Kirill D. Nadezhdin is a research fellow in the laboratory of biomolecular NMR spectroscopy at the Shemyakin-Ovchnnikov Institute of Bioorganic Chemistry, Moscow. He received his master’s degree at the Moscow Institute of Physics and Technology (MIPT) in 2006 and his PhD in Biophysics at the Lomonosov Moscow State University in 2012. He is working as an assistant professor and deputy dean at the Department of Biological and Medical Physics in MIPT since 2015. Scientific interests of Dr. Nadezhdin are focused in the area of structural investigations of amyloid precursor protein and membrane-active peptides.
Acknowledgments
This work is supported by the Russian Science Foundation (grant no. 14-14-00573).
References
[1] Wallin E, von Heijne G. Genome-wide analysis of integral membrane proteins from eubacterial, archaean, and eukaryotic organisms. Protein Sci. 1998, 7, 1029–1038.10.1002/pro.5560070420Search in Google Scholar PubMed PubMed Central
[2] Arinaminpathy Y, Khurana E, Engelman DM, Gerstein MB. Computational analysis of membrane proteins: the largest class of drug targets. Drug Discov. Today 2009, 14, 1130–1135.10.1016/j.drudis.2009.08.006Search in Google Scholar PubMed PubMed Central
[3] Quast RB, Sonnabend A, Stech M, Wüstenhagen DA, Kubick S. High-yield cell-free synthesis of human EGFR by IRES-mediated protein translation in a continuous exchange cell-free reaction format. Sci. Rep. 2016, 6, 30399.10.1038/srep30399Search in Google Scholar PubMed PubMed Central
[4] Meissner A, Sorensen OW. Optimization of three-dimensional TROSY-type HCCH NMR correlation of aromatic (1)H-(13)C groups in proteins. J. Magn. Reson. 1999, 139, 447–450.10.1006/jmre.1999.1796Search in Google Scholar PubMed
[5] Tugarinov V, Hwang PM, Kay LE. Nuclear magnetic resonance spectroscopy of high-molecular-weight proteins. Ann. Rev. Biochem. 2004, 73, 107–146.10.1146/annurev.biochem.73.011303.074004Search in Google Scholar PubMed
[6] Pervushin K, Riek R, Wider G, Wüthrich 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. USA 1997, 94, 12366–12371.10.1142/9789811235795_0006Search in Google Scholar
[7] Tugarinov V, Hwang PM, Ollerenshaw JE, Kay LE. Cross-correlated relaxation enhanced 1H[bond]13C NMR spectroscopy of methyl groups in very high molecular weight proteins and protein complexes. J. Am. Chem. Soc. 2003, 125, 10420–10428.10.1021/ja030153xSearch in Google Scholar PubMed
[8] Kainosho M, Torizawa T, Iwashita Y, Terauchi T, Mei Ono A, Güntert P. Optimal isotope labelling for NMR protein structure determinations. Nature 2006, 440, 52–57.10.1038/nature04525Search in Google Scholar PubMed
[9] Miyanoiri Y, Ishida Y, Takeda M, Terauchi T, Inouye M, Kainosho M. Highly efficient residue-selective labeling with isotope-labeled Ile, Leu, and Val using a new auxotrophic E. coli strain. J. Biomol. NMR. 2016, 65, 109–119.10.1007/s10858-016-0042-0Search in Google Scholar PubMed
[10] Ruschak AM, Kay LE. Methyl groups as probes of supra-molecular structure, dynamics and function. J. Biomol. NMR. 2010, 46, 75–87.10.1007/s10858-009-9376-1Search in Google Scholar PubMed
[11] Velyvis A, Ruschak AM, Kay LE. An economical method for production of 2H,13CH3-threonine for solution NMR studies of large protein complexes: application to the 670 kDa proteasome. PLoS One 2012, 7, e43725.10.1371/journal.pone.0043725Search in Google Scholar
[12] Clark L, Zahm JA, Ali R, Kukula M, Bian L, Patrie SM, Gardner KH, Rosen MK, Rosenbaum DM. Methyl labeling and TROSY NMR spectroscopy of proteins expressed in the eukaryote Pichia pastoris. J. Biomol. NMR. 2015, 62, 239–245.10.1007/s10858-015-9939-2Search in Google Scholar
[13] Rosen MK, Gardner KH, Willis RC, Parris WE, Pawson T, Kay LE. Selective methyl group protonation of perdeuterated proteins. J. Mol. Biol. 1996, 263, 627–636.10.1006/jmbi.1996.0603Search in Google Scholar
[14] Tugarinov V, Kay LE. Ile, Leu, and Val methyl assignments of the 723-residue malate synthase G using a new labeling strategy and novel NMR methods. J. Am. Chem. Soc. 2003, 125, 13868–13878.10.1021/ja030345sSearch in Google Scholar
[15] Religa TL, Sprangers R, Kay LE. Dynamic regulation of archaeal proteasome gate opening as studied by TROSY NMR. Science 2010, 328, 98–102.10.1126/science.1184991Search in Google Scholar
[16] Fernández C, Wüthrich K. NMR solution structure determination of membrane proteins reconstituted in detergent micelles. FEBS Lett. 2003, 555, 144–150.10.1016/S0014-5793(03)01155-4Search in Google Scholar
[17] Dürr UHN, Gildenberg M, Ramamoorthy A. The magic of bicelles lights up membrane protein structure. Chem. Rev. 2012, 112, 6054–6074.10.1021/cr300061wSearch in Google Scholar PubMed PubMed Central
[18] Dürr UHN, Soong R, Ramamoorthy A. When detergent meets bilayer: birth and coming of age of lipid bicelles. Prog. Nucl. Magn. Reson. Spectrosc. 2013, 69, 1–22.10.1016/j.pnmrs.2013.01.001Search in Google Scholar PubMed PubMed Central
[19] Poget SF, Girvin ME. Solution NMR of membrane proteins in bilayer mimics: small is beautiful, but sometimes bigger is better. Biochim. Biophys. Acta 2007, 1768, 3098–3106.10.1016/j.bbamem.2007.09.006Search in Google Scholar PubMed PubMed Central
[20] Malhotra K, Alder NN. Advances in the use of nanoscale bilayers to study membrane protein structure and function. Biotechnol. Genet. Eng. Rev. 2014, 30, 79–93.10.1080/02648725.2014.921502Search in Google Scholar PubMed
[21] Warschawski DE, Arnold AA, Beaugrand M, Gravel A, Chartrand É, Marcotte I. Choosing membrane mimetics for NMR structural studies of transmembrane proteins. Biochim. Biophys. Acta 2011, 1808, 1957–1974.10.1016/j.bbamem.2011.03.016Search in Google Scholar PubMed
[22] Mäler L. Solution NMR studies of peptide-lipid interactions in model membranes. Mol. Membr. Biol. 2012, 29, 155–176.10.3109/09687688.2012.683456Search in Google Scholar PubMed
[23] Raschle T, Hiller S, Etzkorn M, Wagner G. Nonmicellar systems for solution NMR spectroscopy of membrane proteins. Curr. Opin. Struct. Biol. 2010, 20, 471–479.10.1016/j.sbi.2010.05.006Search in Google Scholar PubMed PubMed Central
[24] Sanders CR, Sönnichsen F. Solution NMR of membrane proteins: practice and challenges. Magn. Reson. Chem. 2006, 44, S24–S40.10.1002/mrc.1816Search in Google Scholar PubMed
[25] Gayen S, Li Q, Kang C. Solution NMR study of the transmembrane domain of single-span membrane proteins: opportunities and strategies. Curr. Protein Pept. Sci. 2012, 13, 585–600.10.2174/138920312803582979Search in Google Scholar PubMed
[26] Liang B, Tamm LK. NMR as a tool to investigate the structure, dynamics and function of membrane proteins. Nat. Struct. Mol. Biol. 2016, 23, 468–474.10.1038/nsmb.3226Search in Google Scholar PubMed PubMed Central
[27] Barsukov IL, Nolde DE, Lomize AL, Arseniev AS. Three-dimensional structure of proteolytic fragment 163-231 of bacterioopsin determined from nuclear magnetic resonance data in solution. Eur. J. Biochem. 1992, 206, 665–672.10.1111/j.1432-1033.1992.tb16972.xSearch in Google Scholar PubMed
[28] Pervushin KV, Orekhov VYu null, Popov AI, Musina LYu null, Arseniev AS. Three-dimensional structure of (1-71)bacterioopsin solubilized in methanol/chloroform and SDS micelles determined by 15N-1H heteronuclear NMR spectroscopy. Eur. J. Biochem. 1994, 219, 571–583.10.1111/j.1432-1033.1994.tb19973.xSearch in Google Scholar PubMed
[29] Grabchuk IA, Orekhov VY, Musina LY, Arseniev AS. 1H-15N NMR signal assignment and the secondary structure of bacteriorhodopsin(1-231) in solution. Bioorg. Khim. 1997, 23, 616–629.Search in Google Scholar
[30] 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 1998, 37, 8817–8824.10.1021/bi980511mSearch in Google Scholar PubMed
[31] Mineev KS, Lyukmanova EN, Krabben L, Serebryakova MV, Shulepko MA, Arseniev AS, Kordyukova LV, Veit M. Structural investigation of influenza virus hemagglutinin membrane-anchoring peptide. Protein Eng. Des. Sel. 2013, 26, 547–552.10.1093/protein/gzt034Search in Google Scholar PubMed
[32] Cohen LS, Arshava B, Neumoin A, Becker JM, Güntert P, Zerbe O, Naider F. Comparative NMR analysis of an 80-residue G protein-coupled receptor fragment in two membrane mimetic environments. Biochim. Biophys. Acta 2011, 1808, 2674–2684.10.1016/j.bbamem.2011.07.011Search in Google Scholar
[33] Mortishire-Smith RJ, Pitzenberger SM, Burke CJ, Middaugh CR, Garsky VM, Johnson RG. Solution structure of the cytoplasmic domain of phopholamban: phosphorylation leads to a local perturbation in secondary structure. Biochemistry 1995, 34, 7603–7613.10.1021/bi00023a006Search in Google Scholar
[34] Jung H, Windhaber R, Palm D, Schnackerz KD. NMR and circular dichroism studies of synthetic peptides derived from the third intracellular loop of the beta-adrenoceptor. FEBS Lett. 1995, 358, 133–136.10.1016/0014-5793(94)01409-TSearch in Google Scholar
[35] Shenkarev ZO, Finkina EI, Nurmukhamedova EK, Balandin SV, Mineev KS, Nadezhdin KD, Yakimenko ZA, Tagaev AA, Temirov YV, Arseniev AS, Ovchinnikova TV. Isolation, structure elucidation, and synergistic antibacterial activity of a novel two-component lantibiotic lichenicidin from Bacillus licheniformis VK21. Biochemistry 2010, 49, 6462–6472.10.1021/bi100871bSearch in Google Scholar
[36] Tulumello DV, Deber CM. Efficiency of detergents at maintaining membrane protein structures in their biologically relevant forms. Biochim. Biophys. Acta 2012, 1818, 1351–1358.10.1016/j.bbamem.2012.01.013Search in Google Scholar
[37] Arseniev AS, Barsukov IL, Bystrov VF, Lomize AL, Ovchinnikov YA. 1H-NMR study of gramicidin A transmembrane ion channel: head-to-head right-handed, single-stranded helices. FEBS Lett. 1985, 186, 168–174.10.1016/0014-5793(85)80702-XSearch in Google Scholar
[38] Lomize AL, Pervushin KV, Arseniev AS. Spatial structure of (34–65)bacterioopsin polypeptide in SDS micelles determined from nuclear magnetic resonance data. J. Biomol. NMR. 1992, 2, 361–372.10.1007/BF01874814Search in Google Scholar PubMed
[39] Pervushin K, Sobol A, Musina L, Abdulaeva G, Arsenev A. Spatial structure of (1-36)bacterioopsin in methanol-chloroform and SDS micelles. Mol. Biol. 1992, 26, 920–933.Search in Google Scholar
[40] Chill JH, Louis JM, Baber JL, Bax A. Measurement of 15N relaxation in the detergent-solubilized tetrameric KcsA potassium channel. J. Biomol. NMR. 2006, 36, 123–136.10.1007/s10858-006-9071-4Search in Google Scholar PubMed
[41] Chill JH, Louis JM, Miller C, Bax A. NMR study of the tetrameric KcsA potassium channel in detergent micelles. Protein Sci. 2006, 15, 684–698.10.1110/ps.051954706Search in Google Scholar PubMed PubMed Central
[42] Chill JH, Louis JM, Delaglio F, Bax A. Local and global structure of the monomeric subunit of the potassium channel KcsA probed by NMR. Biochim. Biophys. Acta 2007, 1768, 3260–3270.10.1016/j.bbamem.2007.08.006Search in Google Scholar PubMed PubMed Central
[43] Krishnamani V, Hegde BG, Langen R, Lanyi JK. Secondary and Tertiary Structure of Bacteriorhodopsin in the SDS Denatured State. Biochemistry 2012, 51, 1051–1060.10.1021/bi201769zSearch in Google Scholar
[44] Gong X-M, Ding Y, Yu J, Yao Y, Marassi FM. Structure of the Na,K-ATPase regulatory protein FXYD2b in micelles: implications for membrane-water interfacial arginines. Biochim. Biophys. Acta 2015, 1848, 299–306.10.1016/j.bbamem.2014.04.021Search in Google Scholar
[45] Li Y, Surya W, Claudine S, Torres J. Structure of a conserved Golgi complex-targeting signal in coronavirus envelope proteins. J. Biol. Chem. 2014, 289, 12535–12549.10.1074/jbc.M114.560094Search in Google Scholar
[46] Miyamoto K, Togiya K. Solution structure of LC4 transmembrane segment of CCR5. PLoS One 2011, 6, e20452.10.1371/journal.pone.0020452Search in Google Scholar
[47] Schubert M, Kolbe M, Kessler B, Oesterhelt D, Schmieder P. Heteronuclear multidimensional nmr spectroscopy of solubilized membrane proteins: resonance assignment of native bacteriorhodopsin. ChemBioChem 2002, 3, 1019–1023.10.1002/1439-7633(20021004)3:10<1019::AID-CBIC1019>3.0.CO;2-CSearch in Google Scholar
[48] Etzkorn M, Raschle T, Hagn F, Gelev V, Rice AJ, Walz T, Wagner G. Cell-free expressed bacteriorhodopsin in different soluble membrane mimetics: biophysical properties and NMR accessibility. Structure 2013, 21, 394–401.10.1016/j.str.2013.01.005Search in Google Scholar
[49] Isogai S, Deupi X, Opitz C, Heydenreich FM, Tsai C-J, Brueckner F, Schertler GFX, Veprintsev DB, Grzesiek S. Backbone NMR reveals allosteric signal transduction networks in the β1-adrenergic receptor. Nature 2016, 530, 237–241.10.1038/nature16577Search in Google Scholar
[50] Nygaard R, Zou Y, Dror RO, Mildorf TJ, Arlow DH, Manglik A, Pan AC, Liu CW, Fung JJ, Bokoch MP, Thian FS, Kobilka TS, Shaw DE, Mueller L, Prosser RS, Kobilka BK. The dynamic process of β2-adrenergic receptor activation. Cell 2013, 152, 532–542.10.1016/j.cell.2013.01.008Search in Google Scholar
[51] Stehle J, Silvers R, Werner K, Chatterjee D, Gande S, Scholz F, Dutta A, Wachtveitl J, Klein-Seetharaman J, Schwalbe H. Characterization of the simultaneous decay kinetics of metarhodopsin states II and III in rhodopsin by solution-state NMR spectroscopy. Angew. Chem. Int. Ed. Engl. 2014, 53, 2078–2084.10.1002/anie.201309581Search in Google Scholar
[52] Ozawa S, Kimura T, Nozaki T, Harada H, Shimada I, Osawa M. Structural basis for the inhibition of voltage-dependent K+ channel by gating modifier toxin. Sci. Rep. 2015, 5, 14226.10.1038/srep14226Search in Google Scholar
[53] Horst R, Liu JJ, Stevens RC, Wüthrich K. β2-Adrenergic receptor activation by agonists studied with 19F NMR spectroscopy. Angew. Chem. Int. Ed. Engl. 2013, 52, 10762–10765.10.1002/anie.201305286Search in Google Scholar PubMed PubMed Central
[54] Thompson AA, Liu JJ, Chun E, Wacker D, Wu H, Cherezov V, Stevens RC. GPCR stabilization using the bicelle-like architecture of mixed sterol-detergent micelles. Methods 2011, 55, 310–317.10.1016/j.ymeth.2011.10.011Search in Google Scholar PubMed PubMed Central
[55] Hagn F, Etzkorn M, Raschle T, Wagner G. Optimized phospholipid bilayer nanodiscs facilitate high-resolution structure determination of membrane proteins. J. Am. Chem. Soc. 2013, 135, 1919–1925.10.1021/ja310901fSearch in Google Scholar PubMed PubMed Central
[56] MacKenzie KR, Prestegard JH, Engelman DM. A transmembrane helix dimer: structure and implications. Science 1997, 276, 131–133.10.1126/science.276.5309.131Search in Google Scholar PubMed
[57] 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 2008, 321, 1206–1210.10.1126/science.1161302Search in Google Scholar PubMed PubMed Central
[58] Sulistijo ES, Mackenzie KR. Structural basis for dimerization of the BNIP3 transmembrane domain. Biochemistry 2009, 48, 5106–5120.10.1021/bi802245uSearch in Google Scholar PubMed
[59] Bocharov EV, Lesovoy DM, Pavlov KV, Pustovalova YE, Bocharova OV, Arseniev AS. Alternative packing of EGFR transmembrane domain suggests that protein-lipid interactions underlie signal conduction across membrane. Biochim. Biophys. Acta 2016, 1858, 1254–1261.10.1016/j.bbamem.2016.02.023Search in Google Scholar PubMed
[60] Shenkarev Z, Paramonov A, Lyukmanova E, Shingarova L, Yakimov S, Dubinnyi M, Chupin V, Kirpichnikov M, Blommers M, Arseniev A. NMR structural and dynamical investigation of the isolated voltage-sensing domain of the potassium channel KvAP: implications for voltage gating. J. Am. Chem. Soc. 2010, 132, 5630–5637.10.1021/ja909752rSearch in Google Scholar PubMed
[61] Mineev KS, Khabibullina NF, Lyukmanova EN, Dolgikh DA, Kirpichnikov MP, Arseniev AS. Spatial structure and dimer – monomer equilibrium of the ErbB3 transmembrane domain in DPC micelles. Biochim. Biophys. Acta 2011, 1808, 2081–2088.10.1016/j.bbamem.2011.04.017Search in Google Scholar PubMed
[62] 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. 2008, 105, 15370–15375.10.1073/pnas.0808115105Search in Google Scholar PubMed PubMed Central
[63] Van Horn WD, Kim H-J, Ellis CD, Hadziselimovic A, Sulistijo ES, Karra MD, Tian C, Sonnichsen FD, Sanders CR. Solution nuclear magnetic resonance structure of membrane-integral diacylglycerol kinase. Science 2009, 324, 1726–1729.10.1126/science.1171716Search in Google Scholar PubMed PubMed Central
[64] Arora A, Abildgaard F, Bushweller JH, Tamm LK. Structure of outer membrane protein A transmembrane domain by NMR spectroscopy. Nat. Struct. Biol. 2001, 8, 334–338.10.1038/86214Search in Google Scholar PubMed
[65] Bocharov EV, Lesovoy DM, Goncharuk SA, Goncharuk MV, Hristova K, Arseniev AS. Structure of FGFR3 transmembrane domain dimer: implications for signaling and human pathologies. Structure 2013, 21, 2087–2093.10.1016/j.str.2013.08.026Search in Google Scholar PubMed PubMed Central
[66] Nadezhdin KD, García-Carpio I, Goncharuk SA, Mineev KS, Arseniev AS, Vilar M. Structural basis of p75 transmembrane domain dimerization. J. Biol. Chem. 2016, 291, 12346–12357.10.1074/jbc.M116.723585Search in Google Scholar PubMed PubMed Central
[67] Bondarenko V, Mowrey D, Tillman T, Cui T, Liu LT, Xu Y, Tang P. NMR structures of the transmembrane domains of the α4β2 nAChR. Biochim. Biophys. Acta 2012, 1818, 1261–1268.10.1016/j.bbamem.2012.02.008Search in Google Scholar PubMed PubMed Central
[68] Mowrey DD, Liu Q, Bondarenko V, Chen Q, Seyoum E, Xu Y, Wu J, Tang P. Insights into distinct modulation of α7 and α7β2 nicotinic acetylcholine receptors by the volatile anesthetic isoflurane. J. Biol. Chem. 2013, 288, 35793–35800.10.1074/jbc.M113.508333Search in Google Scholar PubMed PubMed Central
[69] 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 2005, 44, 15834–15841.10.1021/bi051656dSearch in Google Scholar PubMed
[70] Zhou Y, Cierpicki T, Jimenez RHF, Lukasik SM, Ellena JF, Cafiso DS, Kadokura H, Beckwith J, Bushweller JH. NMR solution structure of the integral membrane enzyme DsbB: functional insights into DsbB-catalyzed disulfide bond formation. Mol. Cell 2008, 31, 896–908.10.1016/j.molcel.2008.08.028Search in Google Scholar PubMed PubMed Central
[71] Liang B, Tamm LK. Structure of outer membrane protein G by solution NMR spectroscopy. Proc. Natl. Acad. Sci. USA 2007, 104, 16140–16145.10.1073/pnas.0705466104Search in Google Scholar PubMed PubMed Central
[72] Zhang Q, Horst R, Geralt M, Ma X, Hong W-X, Finn MG, Stevens RC. Microscale NMR screening of new detergents for membrane protein structural biology. J. Am. Chem. Soc. 2008, 130, 7357–7363.10.1021/ja077863dSearch in Google Scholar PubMed PubMed Central
[73] Stanczak P, Zhang Q, Horst R, Serrano P, Wüthrich K. Micro-coil NMR to monitor optimization of the reconstitution conditions for the integral membrane protein OmpW in detergent micelles. J. Biomol. NMR. 2012, 54, 129–133.10.1007/s10858-012-9658-xSearch in Google Scholar PubMed PubMed Central
[74] Horst R, Stanczak P, Wüthrich K. NMR polypeptide backbone conformation of the E. coli outer membrane protein W. Structure 2014, 22, 1204–1209.10.1016/j.str.2014.05.016Search in Google Scholar PubMed PubMed Central
[75] Marassi FM, Ding Y, Schwieters CD, Tian Y, Yao Y. Backbone structure of Yersinia pestis Ail determined in micelles by NMR-restrained simulated annealing with implicit membrane solvation. J. Biomol. NMR. 2015, 63, 59–65.10.1007/s10858-015-9963-2Search in Google Scholar PubMed PubMed Central
[76] 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. 2004, 28, 43–57.10.1023/B:JNMR.0000012875.80898.8fSearch in Google Scholar
[77] Maslennikov I, Klammt C, Hwang E, Kefala G, Okamura M, Esquivies L, Mörs 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. USA 2010, 107, 10902–10907.10.1073/pnas.1001656107Search in Google Scholar PubMed PubMed Central
[78] Zhuang T, Jap BK, Sanders CR. Solution NMR approaches for establishing specificity of weak heterodimerization of membrane proteins. J. Am. Chem. Soc. 2011, 133, 20571–20580.10.1021/ja208972hSearch in Google Scholar PubMed PubMed Central
[79] Barrett PJ, Song Y, Van Horn WD, Hustedt EJ, Schafer JM, Hadziselimovic A, Beel AJ, Sanders CR. The amyloid precursor protein has a flexible transmembrane domain and binds cholesterol. Science 2012, 336, 1168–1171.10.1126/science.1219988Search in Google Scholar PubMed PubMed Central
[80] Shenkarev ZO, Lyukmanova EN, Paramonov AS, Shingarova LN, Chupin VV, Kirpichnikov MP, Blommers MJJ, Arseniev AS. Lipid-protein nanodiscs as reference medium in detergent screening for high-resolution NMR studies of integral membrane proteins. J. Am. Chem. Soc 2010, 132, 5628–5629.10.1021/ja9097498Search in Google Scholar PubMed
[81] Gautier A, Mott HR, Bostock MJ, Kirkpatrick JP, Nietlispach D. Structure determination of the seven-helix transmembrane receptor sensory rhodopsin II by solution NMR spectroscopy. Nat. Struct. Mol. Biol. 2010, 17, 768–774.10.1038/nsmb.1807Search in Google Scholar PubMed PubMed Central
[82] Reckel S, Gottstein D, Stehle J, Löhr F, Verhoefen M-K, Takeda M, Silvers R, Kainosho M, Glaubitz C, Wachtveitl J, Bernhard F, Schwalbe H, Güntert P, Dötsch V. Solution NMR Structure of Proteorhodopsin. Angew. Chem. Int. Ed. 2011, 50, 11942–11946.10.1002/anie.201105648Search in Google Scholar PubMed PubMed Central
[83] Butterwick JA, MacKinnon R. Solution structure and phospholipid interactions of the isolated voltage-sensor domain from KvAP. J. Mol. Biol. 2010, 403, 591–606.10.1016/j.jmb.2010.09.012Search in Google Scholar PubMed PubMed Central
[84] Renault M, Saurel O, Czaplicki J, Demange P, Gervais V, Löhr F, Réat 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. 2009, 385, 117–130.10.1016/j.jmb.2008.10.021Search in Google Scholar PubMed
[85] Fernández C, Hilty C, Wider G, Güntert P, Wüthrich K. NMR structure of the integral membrane protein OmpX. J. Mol. Biol. 2004, 336, 1211–1221.10.1016/j.jmb.2003.09.014Search in Google Scholar PubMed
[86] Edrington TC, Kintz E, Goldberg JB, Tamm LK. Structural basis for the interaction of lipopolysaccharide with outer membrane protein H (OprH) from Pseudomonas aeruginosa. J. Biol. Chem. 2011, 286, 39211–39223.10.1074/jbc.M111.280933Search in Google Scholar PubMed PubMed Central
[87] Kucharska I, Seelheim P, Edrington T, Liang B, Tamm LK. OprG harnesses the dynamics of its extracellular loops to transport small amino acids across the outer membrane of Pseudomonas aeruginosa. Structure 2015, 23, 2234–2245.10.1016/j.str.2015.10.009Search in Google Scholar PubMed PubMed Central
[88] Kucharska I, Liang B, Ursini N, Tamm LK. Molecular Interactions of lipopolysaccharide with an outer membrane protein from Pseudomonas aeruginosa probed by solution NMR. Biochemistry 2016, 55, 5061–5072.10.1021/acs.biochem.6b00630Search in Google Scholar PubMed PubMed Central
[89] Columbus L, Lipfert J, Jambunathan K, Fox DA, Sim AYL, Doniach S, Lesley SA. Mixing and matching detergents for membrane protein NMR structure determination. J. Am. Chem. Soc. 2009, 131, 7320–7326.10.1021/ja808776jSearch in Google Scholar PubMed PubMed Central
[90] Eichmann C, Tzitzilonis C, Bordignon E, Maslennikov I, Choe S, Riek R. Solution NMR structure and functional analysis of the integral membrane protein YgaP from Escherichia coli. J. Biol. Chem. 2014, 289, 23482–23503.10.1074/jbc.M114.571935Search in Google Scholar PubMed PubMed Central
[91] Eichmann C, Tzitzilonis C, Nakamura T, Kwiatkowski W, Maslennikov I, Choe S, Lipton SA, Riek R. S-nitrosylation induces structural and dynamical changes in a Rhodanese family protein. J. Mol. Biol. 2016, 428, 3737–3751.10.1016/j.jmb.2016.07.010Search in Google Scholar PubMed PubMed Central
[92] Call ME, Schnell JR, Xu C, Lutz RA, Chou JJ, Wucherpfennig KW. The structure of the ζζ transmembrane dimer reveals features essential for its assembly with the T cell receptor. Cell 2006, 127, 355–368.10.1016/j.cell.2006.08.044Search in Google Scholar PubMed PubMed Central
[93] Call ME, Wucherpfennig KW, Chou JJ. The structural basis for intramembrane assembly of an activating immunoreceptor complex. Nature Immunol. 2010, 11, 1023–1029.10.1038/ni.1943Search in Google Scholar
[94] Lyukmanova EN, Shenkarev ZO, Khabibullina NF, Kopeina GS, Shulepko MA, Paramonov AS, Mineev KS, Tikhonov RV, Shingarova LN, Petrovskaya LE, Dolgikh DA, Arseniev AS, Kirpichnikov MP. Lipid-protein nanodiscs for cell-free production of integral membrane proteins in a soluble and folded state: comparison with detergent micelles, bicelles and liposomes. Biochim. Biophys. Acta 2012, 1818, 349–358.10.1016/j.bbamem.2011.10.020Search in Google Scholar
[95] Glover KJ, Whiles JA, Wu G, Yu N, Deems R, Struppe JO, Stark RE, Komives EA, Vold RR. Structural evaluation of phospholipid bicelles for solution-state studies of membrane-associated biomolecules. Biophys. J. 2001, 81, 2163–2171.10.1016/S0006-3495(01)75864-XSearch in Google Scholar
[96] Lee D, Walter KFA, Brückner A-K, Hilty C, Becker S, Griesinger C. Bilayer in small bicelles revealed by lipid-protein interactions using NMR spectroscopy. J. Am. Chem. Soc. 2008, 130, 13822–13823.10.1021/ja803686pSearch in Google Scholar
[97] Chou JJ, Baber JL, Bax A. Characterization of phospholipid mixed micelles by translational diffusion. J. Biomol. NMR. 2004, 29, 299–308.10.1023/B:JNMR.0000032560.43738.6aSearch in Google Scholar
[98] Lind J, Nordin J, Mäler L. Lipid dynamics in fast-tumbling bicelles with varying bilayer thickness: effect of model transmembrane peptides. Biochim. Biophys. Acta 2008, 1778, 2526–2534.10.1016/j.bbamem.2008.07.010Search in Google Scholar
[99] Mineev KS, Nadezhdin KD, Goncharuk SA, Arseniev AS. Characterization of small isotropic bicelles with various compositions. Langmuir 2016, 32, 6624–6637.10.1021/acs.langmuir.6b00867Search in Google Scholar
[100] Struppe J, Whiles JA, Vold RR. Acidic phospholipid bicelles: a versatile model membrane system. Biophys. J. 2000, 78, 281–289.10.1016/S0006-3495(00)76591-XSearch in Google Scholar
[101] Marcotte I, Dufourc EJ, Ouellet M, Auger M. Interaction of the neuropeptide met-enkephalin with zwitterionic and negatively charged bicelles as viewed by 31P and 2H solid-state NMR. Biophys. J. 2003, 85, 328–339.10.1016/S0006-3495(03)74477-4Search in Google Scholar
[102] Barbosa-Barros L, de la Maza A, López-Iglesias C, López O. Ceramide effects in the bicelle structure. Colloids Surf. A Physicochem. Eng. Asp. 2008, 317, 576–584.10.1016/j.colsurfa.2007.11.044Search in Google Scholar
[103] Ye W, Liebau J, Mäler L. New membrane mimetics with galactolipids: lipid properties in fast-tumbling bicelles. J. Phys. Chem. B. 2013, 117, 1044–1050.10.1021/jp311093pSearch in Google Scholar PubMed
[104] Liebau J, Pettersson P, Zuber P, Ariöz C, Mäler L. Fast-tumbling bicelles constructed from native Escherichia coli lipids. Biochim. Biophys. Acta 2016, 1858, 2097–2105.10.1016/j.bbamem.2016.06.008Search in Google Scholar PubMed
[105] Sasaki H, Fukuzawa S, Kikuchi J, Yokoyama S, Hirota H, Tachibana K. Cholesterol doping induced enhanced stability of bicelles. Langmuir 2003, 19, 9841–9844.10.1021/la0345183Search in Google Scholar
[106] Gayen A, Mukhopadhyay C. Evidence for effect of GM1 on opioid peptide conformation: NMR study on leucine enkephalin in ganglioside-containing isotropic phospholipid bicelles. Langmuir 2008, 24, 5422–5432.10.1021/la704056dSearch in Google Scholar PubMed
[107] Gayen A, Goswami SK, Mukhopadhyay C. NMR evidence of GM1-induced conformational change of Substance P using isotropic bicelles. Biochim. Biophys. Acta 2011, 1808, 127–139.10.1016/j.bbamem.2010.09.023Search in Google Scholar PubMed
[108] Yamaguchi T, Uno T, Uekusa Y, Yagi-Utsumi M, Kato K. Ganglioside-embedding small bicelles for probing membrane-landing processes of intrinsically disordered proteins. Chem. Commun. (Camb.) 2013, 49, 1235–1237.10.1039/c2cc38016aSearch in Google Scholar PubMed
[109] Cavagnero S, Dyson HJ, Wright PE. Improved low pH bicelle system for orienting macromolecules over a wide temperature range. J. Biomol. NMR. 1999, 13, 387–391.10.1023/A:1008360022444Search in Google Scholar
[110] Sanders CR, Schwonek JP. Characterization of magnetically orientable bilayers in mixtures of dihexanoylphosphatidylcholine and dimyristoylphosphatidylcholine by solid-state NMR. Biochemistry 1992, 31, 8898–8905.10.1021/bi00152a029Search in Google Scholar PubMed
[111] Gabriel NE, Roberts MF. Interaction of short-chain lecithin with long-chain phospholipids: characterization of vesicles that form spontaneously. Biochemistry 1986, 25, 2812–2821.10.1021/bi00358a012Search in Google Scholar PubMed
[112] Gabriel NE, Roberts MF. Spontaneous formation of stable unilamellar vesicles. Biochemistry 1984, 23, 4011–4015.10.1021/bi00313a001Search in Google Scholar PubMed
[113] Sanders CR, Prestegard JH. Magnetically orientable phospholipid bilayers containing small amounts of a bile salt analogue, CHAPSO. Biophys. J. 1990, 58, 447–460.10.1016/S0006-3495(90)82390-0Search in Google Scholar
[114] Lee SC, Bennett BC, Hong W-X, Fu Y, Baker KA, Marcoux J, Robinson CV, Ward AB, Halpert JR, Stevens RC, Stout CD, Yeager MJ, Zhang Q. Steroid-based facial amphiphiles for stabilization and crystallization of membrane proteins. Proc. Natl. Acad. Sci. USA 2013, 110, e1203–1211.10.1073/pnas.1221442110Search in Google Scholar PubMed PubMed Central
[115] Morgado L, Zeth K, Burmann BM, Maier T, Hiller S. Characterization of the insertase BamA in three different membrane mimetics by solution NMR spectroscopy. J. Biomol. NMR. 2015, 61, 333–345.10.1007/s10858-015-9906-ySearch in Google Scholar PubMed
[116] Matsui R, Ohtani M, Yamada K, Hikima T, Takata M, Nakamura T, Koshino H, Ishida Y, Aida T. Chemically locked bicelles with high thermal and kinetic stability. Angew. Chem. Int. Ed. 2015, 54, 13284–13288.10.1002/anie.201506781Search in Google Scholar PubMed
[117] Triba MN, Warschawski DE, Devaux PF. Reinvestigation by phosphorus NMR of lipid distribution in bicelles. Biophys. J. 2005, 88, 1887–1901.10.1529/biophysj.104.055061Search in Google Scholar PubMed PubMed Central
[118] Beaugrand M, Arnold AA, Hénin J, Warschawski DE, Williamson PTF, Marcotte I. Lipid concentration and molar ratio boundaries for the use of isotropic bicelles. Langmuir 2014, 30, 6162–6170.10.1021/la5004353Search in Google Scholar PubMed PubMed Central
[119] Ye W, Lind J, Eriksson J, Mäler L. Characterization of the morphology of fast-tumbling bicelles with varying composition. Langmuir 2014, 30, 5488–5496.10.1021/la500231zSearch in Google Scholar PubMed
[120] Li M, Morales HH, Katsaras J, Kučerka N, Yang Y, Macdonald PM, Nieh M-P. Morphological characterization of DMPC/CHAPSO bicellar mixtures: a combined SANS and NMR study. Langmuir 2013, 29, 15943–15957.10.1021/la402799bSearch in Google Scholar PubMed
[121] Park SH, Prytulla S, De Angelis AA, Brown JM, Kiefer H, Opella SJ. High-resolution NMR spectroscopy of a GPCR in aligned bicelles. J. Am. Chem. Soc. 2006, 128, 7402–7403.10.1021/ja0606632Search in Google Scholar PubMed PubMed Central
[122] Prosser RS, Evanics F, Kitevski JL, Al-Abdul-Wahid MS. Current applications of bicelles in NMR studies of membrane-associated amphiphiles and proteins. Biochemistry 2006, 45, 8453–8465.10.1021/bi060615uSearch in Google Scholar PubMed
[123] Poget SF, Cahill SM, Girvin ME. Isotropic bicelles stabilize the functional form of a small multidrug-resistance pump for NMR structural studies. J. Am. Chem. Soc. 2007, 129, 2432–2433.10.1021/ja0679836Search in Google Scholar PubMed PubMed Central
[124] 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. 2007, 282, 16256–16266.10.1074/jbc.M701745200Search in Google Scholar PubMed
[125] 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. 2010, 400, 231–243.10.1016/j.jmb.2010.05.016Search in Google Scholar PubMed
[126] 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. 2008, 283, 6950–6956.10.1074/jbc.M709202200Search in Google Scholar PubMed
[127] Bocharov EV, Mineev KS, Goncharuk MV, Arseniev AS. Structural and thermodynamic insight into the process of “weak” dimerization of the ErbB4 transmembrane domain by solution NMR. Biochim. Biophys. Acta 2012, 1818, 2158–2170.10.1016/j.bbamem.2012.05.001Search in Google Scholar PubMed
[128] 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 2010, 98, 881–889.10.1016/j.bpj.2009.11.008Search in Google Scholar PubMed PubMed Central
[129] 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. 2008, 283, 29385–29395.10.1074/jbc.M803089200Search in Google Scholar PubMed PubMed Central
[130] Lau T-L, Kim C, Ginsberg MH, Ulmer TS. The structure of the integrin alphaIIbbeta3 transmembrane complex explains integrin transmembrane signalling. EMBO J. 2009, 28, 1351–1361.10.1038/emboj.2009.63Search in Google Scholar PubMed PubMed Central
[131] Endres NF, Das R, Smith AW, Arkhipov A, Kovacs E, Huang Y, Pelton JG, Shan Y, Shaw DE, Wemmer DE, Groves JT, Kuriyan J. Conformational coupling across the plasma membrane in activation of the EGF receptor. Cell 2013, 152, 543–556.10.1016/j.cell.2012.12.032Search in Google Scholar PubMed PubMed Central
[132] Schmidt T, Ye F, Situ AJ, An W, Ginsberg MH, Ulmer TS. A Conserved ectodomain-transmembrane domain linker motif tunes the allosteric regulation of cell surface receptors. J. Biol. Chem. 2016, 291, 17536–17546.10.1074/jbc.M116.733683Search in Google Scholar PubMed PubMed Central
[133] Liu Y, Kahn RA, Prestegard JH. Dynamic structure of membrane-anchored Arf*GTP. Nat. Struct. Mol. Biol. 2010, 17, 876–881.10.1038/nsmb.1853Search in Google Scholar PubMed PubMed Central
[134] Zhang M, Huang R, Im S-C, Waskell L, Ramamoorthy A. Effects of membrane mimetics on cytochrome P450-cytochrome b5 interactions characterized by NMR spectroscopy. J. Biol. Chem. 2015, 290, 12705–12718.10.1074/jbc.M114.597096Search in Google Scholar PubMed PubMed Central
[135] Poget SF, Harris R, Cahill SM, Girvin ME. 1H, 13C, 15N backbone NMR assignments of the Staphylococcus aureus small multidrug-resistance pump (Smr) in a functionally active conformation. Biomol. NMR Assign. 2010, 4, 139–142.10.1007/s12104-010-9228-7Search in Google Scholar PubMed PubMed Central
[136] Dev J, Park D, Fu Q, Chen J, Ha HJ, Ghantous F, Herrmann T, Chang W, Liu Z, Frey G, Seaman MS, Chen B, Chou JJ. Structural basis for membrane anchoring of HIV-1 envelope spike. Science 2016, 353, 172–175.10.1126/science.aaf7066Search in Google Scholar PubMed PubMed Central
[137] Denisov IG, Grinkova YV, Lazarides AA, Sligar SG. Directed self-assembly of monodisperse phospholipid bilayer Nanodiscs with controlled size. J. Am. Chem. Soc. 2004, 126, 3477–3487.10.1021/ja0393574Search in Google Scholar PubMed
[138] Midtgaard SR, Pedersen MC, Kirkensgaard JJK, Sørensen KK, Mortensen K, Jensen KJ, Arleth L. Self-assembling peptides form nanodiscs that stabilize membrane proteins. Soft Matter 2014, 10, 738–752.10.1039/C3SM51727FSearch in Google Scholar
[139] Kondo H, Ikeda K, Nakano M. Formation of size-controlled, denaturation-resistant lipid nanodiscs by an amphiphilic self-polymerizing peptide. Colloids Surf. B. Biointerfaces 2016, 146, 423–430.10.1016/j.colsurfb.2016.06.040Search in Google Scholar PubMed
[140] Denisov IG, McLean MA, Shaw AW, Grinkova YV, Sligar SG. Thermotropic phase transition in soluble nanoscale lipid bilayers. J. Phys. Chem. B. 2005, 109, 15580–15588.10.1021/jp051385gSearch in Google Scholar PubMed PubMed Central
[141] Nakano M, Fukuda M, Kudo T, Miyazaki M, Wada Y, Matsuzaki N, Endo H, Handa T. Static and dynamic properties of phospholipid bilayer nanodiscs. J. Am. Chem. Soc. 2009, 131, 8308–8312.10.1021/ja9017013Search in Google Scholar PubMed
[142] Stepien P, Polit A, Wisniewska-Becker A. Comparative EPR studies on lipid bilayer properties in nanodiscs and liposomes. Biochim. Biophys. Acta 2015, 1848, 60–66.10.1016/j.bbamem.2014.10.004Search in Google Scholar PubMed
[143] Mörs K, Roos C, Scholz F, Wachtveitl J, Dötsch V, Bernhard F, Glaubitz C. Modified lipid and protein dynamics in nanodiscs. Biochim. Biophys. Acta 2013, 1828, 1222–1229.10.1016/j.bbamem.2012.12.011Search in Google Scholar PubMed
[144] Dörr JM, Koorengevel MC, Schäfer M, Prokofyev AV, Scheidelaar S, van der Cruijsen EAW, Dafforn TR, Baldus M, Killian JA. Detergent-free isolation, characterization, and functional reconstitution of a tetrameric K+ channel: the power of native nanodiscs. Proc. Natl. Acad. Sci. USA 2014, 111, 18607–18612.10.1073/pnas.1416205112Search in Google Scholar PubMed PubMed Central
[145] Lyukmanova EN, Shenkarev ZO, Paramonov AS, Sobol AG, Ovchinnikova TV, Chupin VV, Kirpichnikov MP, Blommers MJJ, Arseniev AS. Lipid-protein nanoscale bilayers: a versatile medium for NMR investigations of membrane proteins and membrane-active peptides. J. Am. Chem. Soc. 2008, 130, 2140–2141.10.1021/ja0777988Search in Google Scholar PubMed
[146] Shenkarev ZO, Paramonov AS, Lyukmanova EN, Gizatullina AK, Zhuravleva AV, Tagaev AA, Yakimenko ZA, Telezhinskaya IN, Kirpichnikov MP, Ovchinnikova TV, Arseniev AS. Peptaibol antiamoebin I: spatial structure, backbone dynamics, interaction with bicelles and lipid-protein nanodiscs, and pore formation in context of barrel-stave model. Chem. Biodivers. 2013, 10, 838–863.10.1002/cbdv.201200421Search in Google Scholar PubMed
[147] Shenkarev ZO, Lyukmanova EN, Paramonov AS, Panteleev PV, Balandin SV, Shulepko MA, Mineev KS, Ovchinnikova TV, Kirpichnikov MP, Arseniev AS. Lipid-protein nanodiscs offer new perspectives for structural and functional studies of water-soluble membrane-active peptides. Acta Naturae. 2014, 6, 84–94.10.32607/20758251-2014-6-2-84-94Search in Google Scholar
[148] Tzitzilonis C, Eichmann C, Maslennikov I, Choe S, Riek R. Detergent/nanodisc screening for high-resolution NMR studies of an integral membrane protein containing a cytoplasmic domain. PLoS One 2013, 8, e54378.10.1371/journal.pone.0054378Search in Google Scholar PubMed PubMed Central
[149] Kucharska I, Edrington TC, Liang B, Tamm LK. Optimizing nanodiscs and bicelles for solution NMR studies of two β-barrel membrane proteins. J. Biomol. NMR. 2015, 61, 261–274.10.1007/s10858-015-9905-zSearch in Google Scholar PubMed PubMed Central
[150] Yu T-Y, Raschle T, Hiller S, Wagner G. Solution NMR spectroscopic characterization of human VDAC-2 in detergent micelles and lipid bilayer nanodiscs. Biochim. Biophys. Acta 2012, 1818, 1562–1569.10.1016/j.bbamem.2011.11.012Search in Google Scholar PubMed PubMed Central
[151] Raschle T, Hiller S, Yu T-Y, Rice AJ, Walz T, Wagner G. Structural and functional characterization of the integral membrane protein VDAC-1 in lipid bilayer nanodiscs. J. Am. Chem. Soc. 2009, 131, 17777–17779.10.1021/ja907918rSearch in Google Scholar PubMed PubMed Central
[152] Glück JM, Wittlich M, Feuerstein S, Hoffmann S, Willbold D, Koenig BW. Integral membrane proteins in nanodiscs can be studied by solution NMR spectroscopy. J. Am. Chem. Soc. 2009, 131, 12060–12061.10.1021/ja904897pSearch in Google Scholar PubMed
[153] Hagn F, Wagner G. Structure refinement and membrane positioning of selectively labeled OmpX in phospholipid nanodiscs. J. Biomol. NMR. 2015, 61, 249–260.10.1007/s10858-014-9883-6Search in Google Scholar PubMed PubMed Central
[154] Fox DA, Larsson P, Lo RH, Kroncke BM, Kasson PM, Columbus L. Structure of the Neisserial outer membrane protein Opa60: loop flexibility essential to receptor recognition and bacterial engulfment. J. Am. Chem. Soc. 2014, 136, 9938–9946.10.1021/ja503093ySearch in Google Scholar PubMed PubMed Central
[155] Bibow S, Carneiro MG, Sabo TM, Schwiegk C, Becker S, Riek R, Lee D. Measuring membrane protein bond orientations in nanodiscs via residual dipolar couplings. Protein Sci. 2014, 23, 851–856.10.1002/pro.2482Search in Google Scholar PubMed PubMed Central
[156] Mineev KS, Goncharuk SA, Kuzmichev PK, Vilar M, Arseniev AS. NMR Dynamics of transmembrane and intracellular domains of p75NTR in lipid-protein nanodiscs. Biophys. J. 2015, 109, 772–782.10.1016/j.bpj.2015.07.009Search in Google Scholar PubMed PubMed Central
[157] Zhang M, Huang R, Ackermann R, Im S-C, Waskell L, Schwendeman A, Ramamoorthy A. Reconstitution of the Cytb5-CytP450 complex in nanodiscs for structural studies using NMR spectroscopy. Angew. Chem. Int. Ed. Engl. 2016, 55, 4497–4499.10.1002/anie.201600073Search in Google Scholar PubMed
[158] Denisov IG, Sligar SG. Nanodiscs for structural and functional studies of membrane proteins. Nat. Struct. Mol. Biol. 2016, 23, 481–486.10.1038/nsmb.3195Search in Google Scholar PubMed PubMed Central
[159] Knowles TJ, Finka R, Smith C, Lin Y-P, Dafforn T, Overduin M. Membrane proteins solubilized intact in lipid containing nanoparticles bounded by styrene maleic acid copolymer. J. Am. Chem. Soc. 2009, 131, 7484–7485.10.1021/ja810046qSearch in Google Scholar PubMed
[160] Orwick-Rydmark M, Lovett JE, Graziadei A, Lindholm L, Hicks MR, Watts A. Detergent-free incorporation of a seven-transmembrane receptor protein into nanosized bilayer Lipodisq particles for functional and biophysical studies. Nano Lett. 2012, 12, 4687–4692.10.1021/nl3020395Search in Google Scholar PubMed
[161] Sahu ID, McCarrick RM, Troxel KR, Zhang R, Smith HJ, Dunagan MM, Swartz MS, Rajan PV, Kroncke BM, Sanders CR, Lorigan GA. DEER EPR measurements for membrane protein structures via bifunctional spin labels and lipodisq nanoparticles. Biochemistry 2013, 52, 6627–6632.10.1021/bi4009984Search in Google Scholar PubMed PubMed Central
[162] Zhang R, Sahu ID, Liu L, Osatuke A, Comer RG, Dabney-Smith C, Lorigan GA. Characterizing the structure of lipodisq nanoparticles for membrane protein spectroscopic studies. Biochim. Biophys. Acta 2015, 1848, 329–333.10.1016/j.bbamem.2014.05.008Search in Google Scholar PubMed PubMed Central
[163] Craig AF, Clark EE, Sahu ID, Zhang R, Frantz ND, Al-Abdul-Wahid MS, Dabney-Smith C, Konkolewicz D, Lorigan GA. Tuning the size of styrene-maleic acid copolymer-lipid nanoparticles (SMALPs) using RAFT polymerization for biophysical studies. Biochim. Biophys. Acta 2016, 1858, 2931–2939.10.1016/j.bbamem.2016.08.004Search in Google Scholar PubMed
[164] Gohon Y, Dahmane T, Ruigrok RWH, Schuck P, Charvolin D, Rappaport F, Timmins P, Engelman DM, Tribet C, Popot J-L, Ebel C. Bacteriorhodopsin/amphipol complexes: structural and functional properties. Biophys. J. 2008, 94, 3523–3537.10.1529/biophysj.107.121848Search in Google Scholar PubMed PubMed Central
[165] Popot J-L, Althoff T, Bagnard D, Banères J-L, Bazzacco P, Billon-Denis E, Catoire LJ, Champeil P, Charvolin D, Cocco MJ, Crémel G, Dahmane T, de la Maza LM, Ebel C, Gabel F, Giusti F, Gohon Y, Goormaghtigh E, Guittet E, Kleinschmidt JH, Kühlbrandt W, Le Bon C, Martinez KL, Picard M, Pucci B, Sachs JN, Tribet C, van Heijenoort C, Wien F, Zito F, Zoonens M. Amphipols from A to Z. Annu. Rev. Biophys. 2011, 40, 379–408.10.1146/annurev-biophys-042910-155219Search in Google Scholar PubMed
[166] Planchard N, Point É, Dahmane T, Giusti F, Renault M, Le Bon C, Durand G, Milon A, Guittet É, Zoonens M, Popot J-L, Catoire LJ. The use of amphipols for solution NMR studies of membrane proteins: advantages and constraints as compared to other solubilizing media. J. Membr. Biol. 2014, 247, 827–842.10.1007/s00232-014-9654-zSearch in Google Scholar PubMed
[167] Zoonens M, Catoire LJ, Giusti F, Popot J-L. NMR study of a membrane protein in detergent-free aqueous solution. Proc. Natl. Acad. Sci. USA 2005, 102, 8893–8898.10.1073/pnas.0503750102Search in Google Scholar PubMed PubMed Central
[168] Catoire LJ, Zoonens M, van Heijenoort C, Giusti F, Popot J-L, Guittet E. Inter- and intramolecular contacts in a membrane protein/surfactant complex observed by heteronuclear dipole-to-dipole cross-relaxation. J. Magn. Reson. 2009, 197, 91–95.10.1016/j.jmr.2008.11.017Search in Google Scholar PubMed
[169] Catoire LJ, Damian M, Giusti F, Martin A, van Heijenoort C, Popot J-L, Guittet E, Banères J-L. Structure of a GPCR ligand in its receptor-bound state: leukotriene B4 adopts a highly constrained conformation when associated to human BLT2. J. Am. Chem. Soc. 2010, 132, 9049–9057.10.1021/ja101868cSearch in Google Scholar PubMed
[170] Catoire LJ, Zoonens M, van Heijenoort C, Giusti F, Guittet E, Popot J-L. Solution NMR mapping of water-accessible residues in the transmembrane beta-barrel of OmpX. Eur. Biophys. J. 2010, 39, 623–630.10.1007/s00249-009-0513-2Search in Google Scholar PubMed
[171] Catoire LJ, Damian M, Baaden M, Guittet E, Banères J-L. Electrostatically-driven fast association and perdeuteration allow detection of transferred cross-relaxation for G protein-coupled receptor ligands with equilibrium dissociation constants in the high-to-low nanomolar range. J. Biomol. NMR 2011, 50, 191–195.10.1007/s10858-011-9523-3Search in Google Scholar PubMed
[172] Etzkorn M, Zoonens M, Catoire LJ, Popot J-L, Hiller S. How amphipols embed membrane proteins: global solvent accessibility and interaction with a flexible protein terminus. J. Membr. Biol. 2014, 247, 965–970.10.1007/s00232-014-9657-9Search in Google Scholar PubMed
[173] Elter S, Raschle T, Arens S, Viegas A, Gelev V, Etzkorn M, Wagner G. The use of amphipols for NMR structural characterization of 7-TM proteins. J. Membr. Biol. 2014, 247, 957–964.10.1007/s00232-014-9669-5Search in Google Scholar PubMed PubMed Central
[174] Dahmane T, Giusti F, Catoire LJ, Popot J-L. Sulfonated amphipols: synthesis, properties, and applications. Biopolymers 2011, 95, 811–823.10.1002/bip.21683Search in Google Scholar PubMed
[175] Bazzacco P, Billon-Denis E, Sharma KS, Catoire LJ, Mary S, Le Bon C, Point E, Banères J-L, Durand G, Zito F, Pucci B, Popot J-L. Nonionic homopolymeric amphipols: application to membrane protein folding, cell-free synthesis, and solution nuclear magnetic resonance. Biochemistry 2012, 51, 1416–1430.10.1021/bi201862vSearch in Google Scholar PubMed
[176] Feinstein HE, Tifrea D, Sun G, Popot J-L, de la Maza LM, Cocco MJ. Long-term stability of a vaccine formulated with the amphipol-trapped major outer membrane protein from Chlamydia trachomatis. J. Membr. Biol. 2014, 247, 1053–1065.10.1007/s00232-014-9693-5Search in Google Scholar PubMed PubMed Central
[177] Focke PJ, Hein C, Hoffmann B, Matulef K, Bernhard F, Dötsch V, Valiyaveetil FI. Combining in Vitro Folding with Cell Free Protein Synthesis for Membrane Protein Expression. Biochemistry 2016, 55, 4212–4219.10.1021/acs.biochem.6b00488Search in Google Scholar PubMed PubMed Central
[178] Roos C, Kai L, Haberstock S, Proverbio D, Ghoshdastider U, Ma Y, Filipek S, Wang X, Dötsch V, Bernhard F. High-level cell-free production of membrane proteins with nanodiscs. Methods Mol. Biol. 2014, 1118, 109–130.10.1007/978-1-62703-782-2_7Search in Google Scholar PubMed
[179] Shenkarev ZO, Lyukmanova EN, Butenko IO, Petrovskaya LE, Paramonov AS, Shulepko MA, Nekrasova OV, Kirpichnikov MP, Arseniev AS. Lipid-protein nanodiscs promote in vitro folding of transmembrane domains of multi-helical and multimeric membrane proteins. Biochim. Biophys. Acta 2013, 1828, 776–784.10.1016/j.bbamem.2012.11.005Search in Google Scholar PubMed
[180] Serebryany E, Zhu GA, Yan ECY. Artificial membrane-like environments for in vitro studies of purified G-protein coupled receptors. Biochim. Biophys. Acta 2012, 1818, 225–233.10.1016/j.bbamem.2011.07.047Search in Google Scholar PubMed
[181] Zhang Q, Ma X, Ward A, Hong W-X, Jaakola V-P, Stevens RC, Finn MG, Chang G. Designing facial amphiphiles for the stabilization of integral membrane proteins. Angew. Chem. Int. Ed. Engl. 2007, 46, 7023–7025.10.1002/anie.200701556Search in Google Scholar PubMed
[182] Tribet C, Audebert R, Popot JL. Amphipols: polymers that keep membrane proteins soluble in aqueous solutions. Proc. Natl. Acad. Sci. USA 1996, 93, 15047–15050.10.1073/pnas.93.26.15047Search in Google Scholar PubMed PubMed Central
[183] Kroncke BM, Columbus L. Identification and removal of nitroxide spin label contaminant: impact on PRE studies of α-helical membrane proteins in detergent. Protein Sci. 2012, 21, 589–595.10.1002/pro.2038Search in Google Scholar PubMed PubMed Central
[184] Thiagarajan-Rosenkranz P, Draney AW, Smrt ST, Lorieau JL. A Positively Charged Liquid Crystalline Medium for Measuring Residual Dipolar Couplings in Membrane Proteins by NMR. J. Am. Chem. Soc. 2015, 137, 11932–11934.10.1021/jacs.5b07515Search in Google Scholar PubMed
[185] Warner LR, Varga K, Lange OF, Baker SL, Baker D, Sousa MC, Pardi A. Structure of the BamC two-domain protein obtained by Rosetta with a limited NMR data set. J. Mol. Biol. 2011, 411, 83–95.10.1016/j.jmb.2011.05.022Search in Google Scholar PubMed PubMed Central
[186] Mineev KS, Goncharuk SA, Arseniev AS. Toll-like receptor 3 transmembrane domain is able to perform various homotypic interactions: an NMR structural study. FEBS Lett. 2014, 588, 3802–3807.10.1016/j.febslet.2014.08.031Search in Google Scholar PubMed
[187] Stanczak P, Horst R, Serrano P, Wüthrich K. NMR characterization of membrane protein-detergent micelle solutions by use of microcoil equipment. J. Am. Chem. Soc. 2009, 131, 18450–18456.10.1021/ja907842uSearch in Google Scholar PubMed PubMed Central
[188] Singer SJ, Nicolson GL. The fluid mosaic model of the structure of cell membranes. Science 1972, 175, 720–731.10.1126/science.175.4023.720Search in Google Scholar PubMed
[189] Ruysschaert J-M, Lonez C. Role of lipid microdomains in TLR-mediated signalling. Biochim. Biophys. Acta 2015, 1848, 1860–1867.10.1016/j.bbamem.2015.03.014Search in Google Scholar PubMed
[190] Risselada HJ, Marrink SJ. The molecular face of lipid rafts in model membranes. Proc. Natl. Acad. Sci. USA 2008, 105, 17367–17372.10.1073/pnas.0807527105Search in Google Scholar PubMed PubMed Central
[191] Hedger G, Sansom MSP. Lipid interaction sites on channels, transporters and receptors: recent insights from molecular dynamics simulations. Biochim. Biophys. Acta 2016, 1858, 2390–2400.10.1016/j.bbamem.2016.02.037Search in Google Scholar PubMed PubMed Central
[192] Hedger G, Sansom MSP, Koldsø H. The juxtamembrane regions of human receptor tyrosine kinases exhibit conserved interaction sites with anionic lipids. Sci. Rep. 2015, 5, 9198.10.1038/srep09198Search in Google Scholar PubMed PubMed Central
[193] Saotome K, Duong-Ly KC, Howard KP. Influenza A M2 protein conformation depends on choice of model membrane: conformation of M2 Protein in Lipid Bilayers. Biopolymers 2015, 104, 405–411.10.1002/bip.22617Search in Google Scholar
[194] Chou JJ, Kaufman JD, Stahl SJ, Wingfield PT, Bax A. Micelle-induced curvature in a water-insoluble HIV-1 Env peptide revealed by NMR dipolar coupling measurement in stretched polyacrylamide gel. J. Am. Chem. Soc. 2002, 124, 2450–2451.10.1021/ja017875dSearch in Google Scholar
[195] Mineev KS, Panova SV, Bocharova OV, Bocharov EV, Arseniev AS. The membrane mimetic affects the spatial structure and mobility of EGFR transmembrane and juxtamembrane domains. Biochemistry 2015, 54, 6295–6298.10.1021/acs.biochem.5b00851Search in Google Scholar
[196] Bragin PE, Mineev KS, Bocharova OV, Volynsky PE, Bocharov EV, Arseniev AS. HER2 Transmembrane domain dimerization coupled with self-association of membrane-embedded cytoplasmic juxtamembrane regions. J. Mol. Biol. 2016, 428, 52–61.10.1016/j.jmb.2015.11.007Search in Google Scholar
[197] Mineev KS, Bocharov EV, Volynsky PE, Goncharuk MV, Tkach EN, Ermolyuk YS, Schulga AA, Chupin VV, Maslennikov IV, Efremov RG, Arseniev AS. Dimeric structure of the transmembrane domain of glycophorin a in lipidic and detergent environments. Acta Naturae 2011, 3, 90–98.10.32607/20758251-2011-3-2-90-98Search in Google Scholar
[198] Smith SO, Song D, Shekar S, Groesbeek M, Ziliox M, Aimoto S. Structure of the transmembrane dimer interface of glycophorin A in membrane bilayers. Biochemistry 2001, 40, 6553–6558.10.1021/bi010357vSearch in Google Scholar
[199] Trenker R, Call ME, Call MJ. Crystal structure of the glycophorin A transmembrane dimer in lipidic cubic phase. J. Am. Chem. Soc. 2015, 137, 15676–15679.10.1021/jacs.5b11354Search in Google Scholar
[200] Jiang Y, Lee A, Chen J, Ruta V, Cadene M, Chait BT, MacKinnon R. X-ray structure of a voltage-dependent K+ channel. Nature 2003, 423, 33–41.10.1038/nature01580Search in Google Scholar
[201] Vogt J, Schulz GE. The structure of the outer membrane protein OmpX from Escherichia coli reveals possible mechanisms of virulence. Structure 1999, 7, 1301–1309.10.1016/S0969-2126(00)80063-5Search in Google Scholar
[202] Li D, Lyons JA, Pye VE, Vogeley L, Aragão D, Kenyon CP, Shah STA, Doherty C, Aherne M, Caffrey M. Crystal structure of the integral membrane diacylglycerol kinase. Nature 2013, 497, 521–524.10.1038/nature12179Search in Google Scholar PubMed PubMed Central
[203] Chen Y, Zhang Z, Tang X, Li J, Glaubitz C, Yang J. Conformation and topology of diacylglycerol kinase in E. coli membranes revealed by solid-state NMR spectroscopy. Angew. Chem. Int. Ed. Engl. 2014, 53, 5624–5628.10.1002/anie.201311203Search in Google Scholar PubMed
©2017 Walter de Gruyter GmbH, Berlin/Boston
This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.