Abstract
Nuclear Magnetic Resonance (NMR) spectroscopy is a powerful tool for investigating the dynamics of biomolecules since it provides a description of motion that is comprehensive, site-specific, and relatively non-invasive. In particular, the study of protein dynamics has benefited from sustained methodological advances in NMR that have expanded the scope and time scales of accessible motion. Yet, many of these advances may not be well known to the more general physical chemistry community. Accordingly, this Perspective provides a glimpse of some of the more powerful methods in liquid state NMR that are helping reshape our understanding of functional motions of proteins.
Keywords: Protein dynamics, Relaxation, RDCs, relaxation dispersion, cross-correlation, order parameters, side chain dynamics, chemical exchange
Since its inception in the late 1940s 1–2 Nuclear Magnetic Resonance (NMR) spectroscopy has cut across all traditional discipline boundaries. It is now a standard tool for synthetic chemistry, biochemistry, and biomedical science. Indeed, the increasingly “black-box” character of commercial spectrometers continues to broaden the accessibility of NMR, by allowing users to focus more on data throughput, and less on its physical origins. The transformation from a curious set of physics experiments to a widespread, increasingly “turn-key” tool is surely the sign of success. And yet, this success may raise questions as to the “bona fide” physical chemistry credentials of NMR. The more general physical chemist may ask:” Is there anything new in NMR? Is there much physical science left in it?”
The answer to these questions is an unequivocal “yes”. This is obvious in solid-state NMR (see e.g. the review by Renault et al. 3) and in research aimed at novel modes for magnetic resonance detection 4–5. But what about liquid-state NMR in chemistry? Here, the methodologies are indeed more mature, with the analytical applications pushing more towards automation. But one area of liquid-state NMR that continues to develop new methods is the study of biomolecular dynamics. The goal of this Perspective is to give the more general physical chemist a sense of the power and scope of liquid-state NMR techniques that are reshaping our understanding of biomolecular dynamics, in particular, those of proteins.
Proteins are Nature’s nano-devices and are examples of complex dynamical systems. Indeed, Gregorio Weber famously described them as “kicking screaming stochastic” molecules 6. A truly predictive understanding of protein function should therefore account for such dynamics. Yet, historically, this assertion was viewed as more of a truism, rather than an imperative. Protein conformational dynamics were often regarded as small perturbations, present to be sure, but usually unnecessary for addressing most aspects of protein behavior. Cases of conformational change were often cast as sharp transitions between distinct static structures, rather than a consequence of thermal fluctuations.
These views no longer dominate. Mounting evidence shows that conformational dynamics can play a vital role in protein stability, recognition, and catalysis. And a powerful force driving these shifting views has been the abundance of protein dynamics information coming from NMR spectroscopy.
NMR is particularly well suited for investigating the potentially complex conformational dynamics of proteins: it can profile dynamics both site-specifically and globally; it is sensitive to motions over a broad time-scale (10−12 – 10−3 s); it is also non-invasive (exogenous tags are optional rather than essential). These assets make NMR an excellent companion to other methods for studying protein dynamics, such as fluorescence depolarization, single-molecule spectroscopy, and small-angle x-ray or quasi-elastic neutron scattering. Moreover, the perception that NMR studies are limited to small proteins (< 20 kDa) is increasingly obsolete. Thanks to sustained advances in spin physics and protein isotope labeling methods, NMR studies are now feasible for molecular leviathans exceeding 100 kDa 7–8.
Sources of Dynamic Information in Liquid-state NMR
Figure 1 shows various types of NMR dynamics experiments and the approximate time-scales of motion they sample. In all of these experiments, the dynamics information ultimately stems from internal Hamiltonians that couple the nuclear spin orientations (spin degrees of freedom) with local classical fields that depend on the nuclear coordinates (spatial degrees of freedom). These Hamiltonians decompose naturally into purely isotropic and anisotropic components that convey dynamic information in different ways 9–12. Below we highlight some concepts concerning these Hamiltonians that are relevant for dynamics studies. Our discussion will center on Hamiltonians relevant to non-proton “heteronuclei”, such as15N, 13C, and 2D(euterium), since the vast majority of NMR experiments use these heteronuclei as local probes of motion in appropriately 15N, 13C, and 2D enriched proteins. These heteronuclei are preferable to protons (1H) since they more readily admit a site-specific description of dynamics; the protons merely serve as the nucleus of final detection. The biosynthetic methods for protein isotope enrichment are described in various reviews 13–14.
Figure 1.
Schematic of popular liquid-state NMR experiments for studying protein dynamics, and the relevant time scales of motion. This perspective focuses on the methods in bold: (heteronuclear) spin relaxation, residual dipolar couplings (RDCs), and spin-locked relaxation dispersion.
We start with the anisotropic Hamiltonians; these depend on the orientation of the molecular bond vectors with respect to the static external magnetic field, B0. Important examples include: (i) heteronuclear magnetic dipole-dipole interactions (HDD) in NH and CH bond vectors between 15N and 13C nuclei (spin 1/2) and their directly bonded protons (1H); (ii) the anisotropic chemical shielding interaction (HCSA) for the same 15N and 13C nuclei; (iii) the nuclear quadrupole interaction (HQ) of a deuterium 2D (spin 1) with its local electric field gradient. The spatial parts of these Hamiltonians are 2nd rank, symmetric, and traceless tensors that describe the local fields associated with the bond vectors. Such fields include the magnetic dipole fields the 15N and 13C spins experience from their covalently bonded protons, and the local electric field gradient sensed by the 2D spins. One can visualize these tensors as little coordinate frames that are rigidly attached to the bond vectors and bear a fixed orientation to them 15. As the bond vector rotates, so do the attached tensors and their associated local fields; the nuclear spin orientations (quantized along B0) do not. Thus, bond vector rotation (reorientation) modulates the local fields sensed by the resident spins. As a result, an ensemble of immobile bond vectors that point in different directions, such as those in a static powder, give rise to extremely broad line shapes. By contrast, mobile bond vectors individually sample multiple orientations, thereby causing rotational averaging of the anisotropic interactions and line narrowing.
In isotropic liquids, the rotational averaging is complete: overall molecular tumbling causes the bond vector orientations to fluctuate among all directions with equal likelihood. Accordingly, the attached HDD, HCSA, and HQ tensors average to zero and have no affect on the NMR resonance positions. Nevertheless, they do cause nuclear spin relaxation. In particular, the tensor fluctuations constitute local fluctuating fields that stimulate spin-state transitions and de-phase coherences, thus enabling relaxation of non-equilibrium bulk spin order (e.g. inverted or transverse magnetization) back to the equilibrium condition (bulk magnetization along B0). The corresponding experimental observables are relaxation rate constants, such as R1 = 1/T1, and R2 = 1/T2, for longitudinal and transverse relaxation, respectively. Thus, heteronuclear spin relaxation rates report on the picosecond (10−12 s) – nanosecond (10−9 s) orientational dynamics of bond vectors (vide infra).
Partial resurrection of the anisotropic Hamiltonians HDD, HCSA, and HQ, is possible by using methods that induce weak protein alignment, while retaining the narrow line widths of conventional liquid-state NMR 16–17. Alignment methods include 18: covalently attaching moieties with large magnetic susceptibility anisotropies that align with B0; co-dissolving the protein with substances that themselves align with B0; and dissolving the protein in porous media with anisotropic cavities that are stretched or compressed along B0. The weak alignment produces small perturbations to the isotropic resonances that reflect the residual affects of the anisotropic Hamiltonians. For example, residual dipolar couplings (RDCs) for NH and CH bonds cause additional 15N-1H and 13C-1H splittings, while residual chemical shift anisotropies (RCSAs) of 15N and 13C cause small resonance shifts. These alignment-induced affects provide information on the average orientations of bond vectors relative to B0, which depends both on overall protein tumbling and internal motion. Thus, similar to spin relaxation, the residual anisotropies report on the orientational dynamics of bond vectors, albeit over much broader time scales, picosecond (10−12 s) – millisecond (10−3 s) (vide infra).
What about the isotropic Hamiltonians? They are responsible for the familiar isotropic chemical shifts and scalar J-coupling constants that are heavily exploited by synthetic and biological chemistry. While these Hamiltonians are indifferent to molecular re-orientation, their isotropic parameters are highly sensitive to local conformation. Dynamics involving the exchange of nuclei among different chemical states can render these parameters time-dependent, leading to various spectral effects. These exchange dynamics often occur among states separated by free energy barriers > kBT, leading to exchange time-constants on the order of microseconds (10−6 s) – milliseconds (10−3 s). These time scales overlap with those of protein binding and enzymatic turnover, and are thus of great interest to biophysics and biophysical chemistry.
Below, we focus on a subset of the experiments in Figure 1 that have proven to be particularly effective within the last fifteen years, including: (i) heteronuclear spin relaxation experiments; (ii) residual dipolar coupling (RDC) measurements; and (iii) chemical exchange experiments. Two caveats deserve mention. First, each of these experiments constitutes a rich and thriving area of NMR research, and has generated critical reviews. This Perspective, which discusses all three, cannot discuss them at a comparable depth or breadth as those reviews. Its more modest goal is to sketch the illuminative power of these methods, in particular, for the more general physical chemist. Along the way, we refer readers seeking more detail to the appropriate critical reviews. Second, the chemical literature often distinguishes “dynamics” from “flexibility”, reserving the former for explicitly time-dependent phenomena, and the latter for equilibrium properties. Yet, both notions invoke atoms in motion. In this Perspective, we use the word “dynamics” in its most elementary sense from classical mechanics: the motions of particles (atoms).
Nuclear Spin Relaxation and Re-orientational Motions
Heteronuclear (15N, 13C, 2D) spin relaxation rate measurements provide a powerful means for sequence-specific descriptions of protein dynamics. The relaxation rates of 15N, 13C, and 2D report on the power spectral density functions J(ω) that describe the orientational dynamics of NH, CH, and CD bond vectors. Examples are longitudinal and transverse 2D relaxation rates, R1 (Dz and R2 (D+), which probe CD bond motions,
| (1a) |
| (1b) |
and longitudinal and transverse 15N relaxation rates, which probe NH bond motions,
| (2a) |
| (2b) |
In Equations 1a,b the constant (e2qQ/ħ) reflects the dominance of deuterium relaxation by the nuclear quadrupolar Hamiltonian, HQ (vide supra). eQ is the nuclear quadrupole moment, eq is the electric field gradient, and ħ is Planck’s constant divided by 2π. In Equations 2a,b the NH dipolar coupling constant (expressed in cgs units), and the 15N chemical shift anisotropy (γNB0Δσ) reflect the heteronuclear NH dipolar (HDD) and 15N chemical shift anisotropy (HCSA) interactions, respectively. Analogous HDD and HCSA interactions dominate 13C relaxation in CH bonds. In all cases, the relaxation mechanisms localize to the CD, NH, and CH bond vectors, thus allowing site-specific interpretation of the data.
Generally, the key dynamics quantities in nuclear spin relaxation rates are the spectral density functions, JIS(ω), where we use “IS” to denote an NH, CD, or CH bond vector. JIS(ω) essentially profiles the component frequencies of the orientational dynamics experienced by an IS bond vector. These dynamics include both conformational dynamics and global protein tumbling (Figure 2a). As seen in Equations 1, 2, the intrinsic dynamics information of the relaxation rates is not JIS(ω) directly, but rather, samplings of them at discrete frequencies in the MHz range (0, ωN, ωD, ωH, etc) (Figure 2d). Critically, overall protein tumbling puts an upper limit on the time scales of motion sensed by the relaxation rates.
Figure 2.
Schematic of dynamics information available from heteronuclear spin relaxation. (a) Relaxation probes the orientational dynamics of bond vectors relative to the external static magnetic field, B0. (b) The orientational dynamics are characterized by a time correlation function C(τ). (c) The C(τ) decay depends on the individual bond vector motion, but always decays to zero due to isotropic overall tumbling. (d) The measured relaxation rates sample the spectral density function, which is the Fourier Transform of C(τ). The panel depicts the cases of deuterium R1 and R2 relaxation rates, which sample the JCD(ω) spectral density at 0, ωD, and 2ωD.
JIS(ω) is the Fourier Transform of the time- (auto)correlation function for “IS” bond vector reorientation, C(τ). For example, for a given “IS” bond vector we have
| (3) |
where
| (4) |
In Equation 4, ΩIS(τ) is the angle between where the vector is now, versus a time ‘τ’ earlier. Hence, C(τ) describes the decay in auto-correlations between the orientations of the same “IS” bond-vector at two different instants, separated by a waiting time ‘τ’ (cf. Figure 2b). C(τ) begins at (1/5) for all vectors and decays to zero on time scales set by τc, the overall rotational correlation time(s), which typically lie(s) in the 10–100 ns range. Orientational fluctuations on longer time scales (τ > τc) are invisible to C(τ) (since it has already decayed to zero, cf. Figure 2c), and by extension, to J(ω) and the spin-relaxation rates. Of course, the specific decay of C(τ) depends on the mobility of the specific bond vector. Vectors with greater or lesser orientational mobility will have C(τ) functions that decay faster or slower, respectively. This variability translates into site-specific JIS(ω) and relaxation rates, via Equation 4.
Equations 1,2 show that different kinds of relaxation rates sample J(ω) in different ways; thus, one can increase the breadth of dynamics information by measuring multiple rate constants for each bond vector. There is now a diverse array of two-dimensional (2-d) proton-detected heteronuclear correlation experiments that measure R1, R2, and a large variety of other rate constants for numerous bond vectors, on a residue-by-residue basis 19–21. The basic procedure involves recording a series of 2-d heteronuclear correlation spectra, which contains a sequence of rf-pulses and delays that isolates the particular relaxation process of interest. The relaxation occurs for a variable delay, such that the resulting 2-d cross-peak intensities depend on the extent of relaxation allowed. One then fits the series of cross-peak intensities to exponential decays to extract the site-specific relaxation rates (cf. Figure 3a, b).
Figure 3.
Examples of protein spin relaxation measurements. (a) A typical contour plot of a 2d 13Cmethyl—1Hmethyl heteronuclear correlation spectrum, for methyl side-chain 2D relaxation of CD bonds in CH2D methyl groups. The protein is the mitotic regulator, Pin1. Each cross peak corresponds to a particular side-chain CH2D methyl group. (b) The various cross-peak decays give site-specific 2D R1 and R2. (c) The rates sample the spectral densities that describe the sub-nanosecond orientational dynamics of the CD bond (which is essentially that of the methyl symmetry axis). (d) Changes in Pin1 methyl side-chain motions upon interaction with a phosphopeptide substrate. Each sphere is a side-chain methyl. The changes are in terms of Lipari-Szabo order parameters, S2axis. Red/Blue spheres indicate increases/decreases of S2axis upon substrate interaction.
To turn the raw relaxation rates into dynamics information, several options emerge. First, if one measures a set of relaxation rates for each bond vector that equals or exceeds the number of unique spectral density samplings introduced, then one can directly determine the intrinsic dynamics quantities (the spectral density samplings) without any a priori model assumptions of the dynamics. This method is the “spectral density mapping” approach 22–24. Second, if the relaxation rate is dominated by a single spectral density sampling, then a complete mapping becomes possible via B0 dispersion studies, using field-cycling methods 25–26. However, the data sets are usually too sparse for direct mapping methods. One then pursues the third and most common approach, which is to choose an analytical model form for J(ω), which contains micro-dynamic parameters that one fits to best reproduce the experimental relaxation rates. The best-known version of this approach is the “model-free” formalism of Lipari and Szabo 27–28, which is closely related to the earlier two-step model by Halle and Wennerström 29. First applied by Dellwo and Wand to the CH bond motions of cyclosporine 30, the Lipari-Szabo formalism has become the most popular approach for analyzing protein NMR relaxation data. In its simplest form, the model-free formalism invokes an orientational time-correlation function
| (5) |
and corresponding spectral density function
| (6) |
that describes the internal motion of each bond vector with two parameters: (i) a generalized order parameter S2, which measures the amplitude of internal re-orientational motions. S2 ranges between 0 and 1, with completely unrestricted (isotropic) internal motion yielding 0, and no internal motion (rigid limit) yielding 1; (ii) an effective correlation time, τe, that gauges the rapidity of the internal motion. These site-specific parameters, along with a global correlation time for isotropic overall tumbling (τc), are adjusted to fit the relaxation data. The result is S2 and τe versus sequence, and thus, a profile of internal bond vector motion on time scales up to those of global protein tumbling (picosecond – ns). S2 is “model-free” because it can be related to more specific models of motion, such as the “Wobbling in a Cone” 31 or the 1-d and 3-d Gaussian-Axial-Fluctuation (GAF) models (vide infra) 32–34. Extensions of the original Lipari-Szabo formalism account for chemical exchange contributions (Rex) to transverse relaxation data 35–37, anisotropic overall tumbling 38–40 and additional internal motions 41–43. Of course, these extensions carry the cost of introducing additional fitting parameters.
A key assumption of the Lipari-Szabo formalism is that motions due to conformational dynamics (internal motion) are statistically decoupled from global tumbling. This assumption is eminently reasonable if the conformational dynamics occur on time scales much faster than global tumbling, or do not significantly affect the overall structure. But the decoupling assumption is suspect if the internal motions change the overall shape of the protein, thus changing its overall rotational diffusion properties, on a time scale comparable to global tumbling. The latter scenario is conceivable for flexibly linked multi-domain proteins and intrinsically disordered proteins. There are several tactics to evaluate and address the possibility of coupling between internal and global motion. Prompers and Brüschweiler have developed the isotropic re-orientational eigenmode dynamics (iRED) approach 44. This is a computational approach that evaluates the covariance matrix of bond vector fluctuations over the course of a molecular dynamics (MD) simulations to assess the true extent of (de)coupling. More recently, investigators 45–46 have developed analytical formalisms in which the protein jumps between conformers with different diffusion tensors. Critically, these formalisms allow for correlations between the bond vector orientational dynamics within a given conformation (diffusion tensor), and the trans-tensor jumping. Another analytical approach is the slowly relaxing local structure (SRLS) model of Freed, Meirovitch, and co-workers 47. Their model derives from two-body Smoluchowski theory and yields a spectral density function that explicitly includes fitting parameters for the coupling of internal and global motions, as well as descriptors for the local anisotropy (asymmetry) of internal motion. In principle, these are assets. However, ascertaining whether SRLS provides a significantly better description of motional coupling remains an active area of research 48–49. Stochastic dynamics studies by Wand and co-workers show that the Lipari-Szabo approach, which is simpler to use, adequately characterizes the amplitudes of local anisotropic motion 50, although recent work by Meirovitch and co-workers suggests SLRS may give more accurate internal correlation times τe 51. The recent theoretical study of Halle 52 provides an updated derivation of the model-free and SRLS approaches as a foundation for comparing their physical bases and inherent assumptions. In any event, formulating a general, analytical treatment of coupled internal and global motions that remains “user-friendly” is an outstanding challenge.
Applications of heteronuclear spin relaxation experiments have decisively demonstrated the importance of protein dynamics in ligand binding, catalysis, and evolution. The initial studies focused on protein backbone dynamics – the motions of methine CαH bonds, as probed by 13C relaxation rates 36,53, and the motions of NH bonds as probed by 15N spin relaxation 35. Today, NH bond motion studies predominate, due to the relative ease of producing 15N-enriched protein, and the freedom from erroneous relaxation measurements arising from homo-nuclear scalar coupling.
The 15N spin relaxation experiments give a residue-by-residue description of the backbone NH bond motions relative to B0. These motions reflect the orientational dynamics of the host peptide plane, caused by local φ, Ψ torsional fluctuations and long-range (e.g. segmental) dynamics. The NH studies have become a regular assay in dynamics-function investigations involving mutation and/or ligand binding. Examples are those by Palmer and co-workers 54, and Peng and co-workers 55. Using a Lipari-Szabo analysis of NH relaxation rates, the investigators compared the NH S2 versus sequence for various loop mutants to investigate sequence-dynamics-binding correlations, and thus, the extent to which sequence encodes recognition-specific fluctuations. NH backbone dynamics studies have also illuminated the recognition mechanisms of intrinsically disordered proteins – a large class of proteins that defy the traditional structure-function dogma, yet, are essential for cell cycle signaling. In particular, NH dynamics show that such proteins recognize their targets, through coupled folding and binding 56–58. NH studies have further investigated the interplay between the picosecond-nanosecond fluctuations of NH bond vectors and the slower dynamics associated with enzymatic turnover, as in the comprehensive studies by Kern and co-workers on mesophilic versus thermophilic adenylate kinase 59–60. Finally, the growing number of backbone dynamics studies underscores the diversity of dynamic responses of proteins to binding. A common stereotype is that ligand binding invariably reduces the intrinsic flexibility of protein. While this is true in many cases, it is by no means a universal response. For example, backbone dynamics studies by Zidek et al 61 indicated that the backbone flexibility of mouse major urinary protein actually increased upon binding its pheromone ligand. This phenomenon is in other proteins, and frequently in side chain dynamics studies (vide infra). Further examples of backbone dynamics studies by spin relaxation are in the comprehensive review by Jarymowycz and Stone 62.
What kinds of specific atomic motions underlie the model-free S2 values? To address this question, molecular dynamics simulations (MD) are illuminating. Investigating the interplay between MD-simulations and NMR dynamics experiments is a thriving field, and sustained improvements in MD force-fields have led, in some cases, to near quantitative agreement between NH S2 calculated from MD trajectories and their experimental counterparts 63. For greater depth, we refer the reader to recent work by Brüschweiler and co-workers 63 and Palmer and co-workers 64. Notably, MD simulations show that many S2 values reflect bond vectors undergoing restricted torsion angle fluctuations. This motivated the development of the 1d and 3d Gaussian Axial Fluctuation (GAF) models 32–34. In particular, in the 3D GAF model, the NH S2 reflect the torsional fluctuations of the peptide plane about three orthogonal axes defined by the equilibrium peptide plane orientation. Typically, the dominant fluctuations occur about an axis defined by Cαi-1—Cαi atoms. The appropriateness of the GAF models indicates that, in some cases, a quadratic potential function is a good approximation of the potential of mean force guiding the NH bond motions.
For further insight into the shape of the potentials of mean force, investigators have varied the sample pressure (see e.g. the work of Akasaka and co-workers 65) or temperature 66–67. In particular, measurements of NH S2 versus temperature by Palmer, Rance and co-workers 67 revealed that quadratic potentials (axially symmetric or 1d GAF) must be enhanced with temperature-dependent force constants that decrease with increasing temperature; moreover, the force constants for NHs separate into two classes corresponding to bond vectors in α helices versus β-sheets.
For studies of side chain dynamics, deuterium (2D) relaxation measurements have proven to be particularly powerful 23,68–69. The 2D rates probe the motions of CD bonds of CH2D and/or CHD2 methyl isotopomers relative to the external field. These motions include torsional fluctuations within a rotamer state, transitions between rotamer states, or both. Typically, one uses Lipari-Szabo model-free formalism to get S2 values for the CD bonds, which factor as S2 = 0.111S2axis. The 0.111 reflects the common rapid rotation of the CD bond about the methyl symmetry axis, whereas S2axis reflects the site-specific motion of the methyl symmetry axis itself. S2axis is thus the central object interest. The abundance of methyls in protein hydrophobic cores ensures a broad sampling of internal motions.
Side chain dynamics studies are less prevalent than those of the backbone, yet they are equally powerful probes of protein functional dynamics. Their importance is well illustrated by the comprehensive work of Wand and co-workers on the side chain dynamics of calmodulin (CaM) 19. Their work shows that side chain studies can expose functional dynamics, such as those sensitive to ligand binding, that would otherwise be missed by a purely NH backbone investigation 19,70–71. Moreover, their variable temperature studies show that the S2axis values follow a “trimodal” distribution, reflecting three classes of motion: “J”, “α”, and “ω”, which correspond to decreasing amplitudes of side chain motion owing to decreased rotamer-state hopping. Remarkably, the S2axis values lacked significant correlation with many intuitive structural parameters, such as burial depth, local packing density, or solvent accessible surface area 19. Thus, defining the structural determinants of S2axis requires further investigation.
Methyl side chain motions have also highlighted the importance of protein conformational entropy in binding thermodynamics. These results derive from formalisms introduced by Akke et al 72 for free energy, and subsequently Li and Wand 73 and Yang and Kay 74 that use changes in Lipari-Szabo order parameters to relate binding free energy to site-specific changes in conformational dynamics. In particular, S2 is an equilibrium average over the ensemble of orientations adopted by a bond vector. Changes in S2 upon ligand binding thus reflect changes in the ensemble breadth and thus, conformational entropy. While these methods are applicable for both the backbone (e.g. NH) and side chain bond vectors, the side chain studies have been most effective for mapping binding-induced changes in conformational entropy. In particular, Wand and co-workers have analyzed the changes CaM S2axis in parallel with thermodynamic studies for a series of peptide ligands. Their results show that changes in protein conformational entropy, sensed by changes in S2axis, play a highly significant role in modulating the binding affinity of CaM toward its various targets 75. This result is highly significant for the design and modification of protein-ligand interactions (e.g. drug design), since such entropy cannot be calculated from static structures. This work well complements recent computational efforts to correlate conformational entropy and NMR order parameters 76.
Methyl side chain dynamics studies have also provided tantalizing evidence for coupled changes in sub-nanosecond dynamics as mechanisms for site-to-site communication within proteins 77–78. In particular, Figure 3c,d illustrate the changes in the methyl side-chain S2 for the mitotic regulator Pin1, upon binding saturating amounts of a known phosphopeptide substrate. The red and blue spheres indicate increased or decreased S2axis, and thus, losses or gains of internal motion, upon substrate interaction. Notably, the sites that lose motion are highly conserved hydrophobic residues that connect the known functional sites of Pin1. These results support the notion of dynamically driven allostery, as proposed by Cooper and Dryden 79.
Methyl side chain dynamics are also appealing for studying the internal motions of supramolecular systems. More specifically, the methyl Transverse Relaxation Optimized SpectroscopY (TROSY) methods developed by Kay and co-workers, which exploit cross-correlated relaxation to select for the slowly relaxing coherences of the methyl spin system, now enable extraction of methyl S2axis in systems exceeding 80 kDa 8,69,80–81. Another attractive feature of the methyl-TROSY experiments is that it permits sensitive measurement of both 13C and 2D relaxation rates for the same methyl groups, thus enhancing accuracy and broadening the scope of the resulting dynamics information. More recently, these methods have been applied to 345kDa KaiC 82, and the 670kDa 20S proteasome 83 and its substrates 84. These systems are truly Brobdingnagian by NMR standards, yet, common enough in biology. Thus, methyl-TROSY opens the possibility of understanding, for the first time, the internal moving parts of these biomolecular machines.
Residual Dipolar Couplings
The most commonly measured residual anisotropies stem from heteronuclear dipolar couplings (HDD), or residual dipolar couplings (RDCs) 16–17. For an IS bond vector with vibrationally averaged interspin distance rIS, the RDC (in hertz) is
| (7) |
where
| (8) |
As depicted in Figure 4a, the angle ΩIS specifies the orientation of the IS bond vector relative to the external field B0. Thus, the RDC is the “full-blown” dipolar-coupling constant (cgs units), scaled down by the orientational average, 〈P2 (cosΩIS)〉. The average is over the ensemble of orientations sampled by the bond vector. This average is zero under conventional isotropic tumbling conditions, but becomes non-zero under conditions of weak alignment, and is tuned by judicious adjustments of alignment media parameters 18.
Figure 4.
Conformational dynamics from residual dipolar couplings (RDCs). (a) The RDC reports on the average orientation of bond vectors (typically NH, CH) with respect to the external static field B0. (b) The overall orientation is decomposed into the orientation with respect to a molecular-fixed frame (the order tensor), and the orientation of that tensor with respect to B0. (c) RDCs are typically measured by comparing apparent heteronuclear splittings under isotropic versus aligned conditions. (d) Examples of the manifold of conformations sampled by ubiquitin (red region, upper left), derived from extensive RDCs followed by order parameter analysis. The manifold is reprinted from Lange, O. F.; Lakomek, N. A.; Fares, C.; Schroder, G. F.; Walter, K. F.; Becker, S.; Meiler, J.; Grubmuller, H.; Griesinger, C.; de Groot, B. L. Science 2008, 320, 1471-5, with permission. Copyright 2008 AAAS.. (e) Example of the manifold of conformations sampled by the TAR-RNA loops, derived from RDC data. The figure is reprinted from Zhang, Q.; Stelzer, A. C.; Fisher, C. K.; Al-Hashimi, H. M. Nature 2007, 450, 1263-7, with permission. Copyright 2007 Nature Publishing Group. In both cases (d) and (e), these orientational motions occur on time scales (τ > τc) that are invisible to spin relaxation. The manifolds encompass the conformations observed in structures of complexes, indicating that the ligands select from a pre-existing dynamic ensemble of conformers.
The most common RDCs are those of the backbone NH and CH bond vectors, and CH bond vectors of methyl side-chains. The measurements typically involve recording proton-detected 2-d heteronuclear correlation spectra that maintain the I-S spin-spin coupling during the indirect dimension. Spectra are recorded under isotropic conditions, and under conditions of partial alignment. Differences in the apparent JIS splitting between the two conditions give parameters 18,85–86 (Figure 4c).
The dynamic information from the RDCs lies in the ensemble average of Equations 7,8. That average encompasses all orientations sampled by the IS bond vector over the time-course of the NMR signal, and is therefore sensitive to motions ranging from picosecond-millisecond time scales. Such motions include overall rotational diffusion and internal conformational dynamics. To disentangle these effects, one expresses the RDC in terms of the coordinates of the IS bond with respect to coordinate frame fixed in the molecule that reflects the overall molecular ordering, as schematized in Figure 4b. Specifically, the overall protein alignment is expressed by a second-rank, symmetric, and traceless order tensor Sij 87–88. Sij has five independent elements in an arbitrary molecular-fixed frame, such as that implicit in a PDB file. Determining these elements requires measuring at least five distinct RDCs per rigid molecule, per alignment medium. The symmetry of Sij means it can be diagonalized to produce three principle values (eigenvalues) and three principles axes (eigenvectors). These axes constitute the principle axes system (PAS) of the ordering tensor, in which a given RDC simplifies to
| (9) |
SZZ, SYY, and SXX are the three principle values of the Sij tensor; they measure the average alignment of the corresponding principle axes via
| (10) |
where Θii is the angle that the ith principle axis makes with respect to B0. The lower-case x,y, and z are the coordinates of the unit vector representing the IS bond (its direction cosines), with respect to the PAS of Sij. The convention |SZZ| ≥ |SYY| ≥ |SXX| means that the PAS Z-axis is the direction within the protein that has greatest alignment along B0. In this context, Equation 9 essentially says that the variation in RDCs among the bond vectors reflects their projection along this Z-axis. The PAS thus provides a common molecular reference frame whereby the average orientations of widely separated bond vectors can be compared. To determine the principle values and PAS from the raw RDC data, one can start from a 3-d structural model, and then apply algorithms that effectively rotate the molecular coordinates to diagonalize Sij while reproducing the experimental RDCs 89–90.
Equation 9 suggests two main strategies for extracting dynamic information from the RDCs 18. The first subdivides the protein into approximately rigid fragments; the order tensor is derived for each using Equation 9. For a totally rigid protein, a common order tensor should fit all fragments; thus RDCs from the separate fragments should each produce the same order tensor (same principle values and axes). Failure to comply with this expected agreement is evidence of inter-fragment mobility. This strategy works well for inter-domain dynamics, in which the domains are approximated as rigid structures 91.
The second approach relates the RDCs to Lipari-Szabo order parameters (vide supra) 92–93. Here, one interprets the x, y, and z expressions in Equation 9 as ensemble averages over the bond vector orientations sampled within the PAS system, and thus;
| (11) |
The key assumption of Equation 11 is that the internal motions are statistically decoupled from those determining the overall molecular alignment. In effect, it says that sub-millisecond fluctuations in bond vector orientation within the PAS will tend to reduce from its static value. One can extract the ensemble averages of the x,y, and z coordinates by measuring RDCs from at least five independent alignment media 18,92. From these averages, and with the aid of a 3-d structural model, one can construct , the Lipari-Szabo generalized order parameter. More specific versions of this approach, such as the 3-d GAF, give analytical expressions for the fluctuations in terms of axial fluctuations of the peptide plane axis 94. Since RDCs sense motions over a broader range of time scales than spin relaxation, we expect ; hence, the former can serve as upper limits to scale the latter. In this context, an important breakthrough is the structure-free Gaussian-Axial-Fluctuation approach (SF GAF) from Blackledge and co-workers 95. This approach enables independent fits of without reference to a structural model or . Critically, their results are consistent with those determined by previously established methods, and reproduce motions with correlation times ~ 400 ns (gleaned from MD simulations). Of course, such motions cannot be detected from spin-relaxation, which are inherently blind to re-orientational dynamics beyond the overall rotational correlation time, τc.
A practical challenge of the approach is finding at least five independent media that are congenial to the protein’s well being. Nevertheless, the information gained is impressive, and has begun reshaping our appreciation of the importance of conformational transitions in molecular recognition. The recent work on ubiquitin provides an example 96. Analysis of the RDCs showed that the recognition loop samples a range of conformers that coincide with those collectively covered by known ubiquitin-ligand complexes (Figure 4d). Thus, different binding partners select from a manifold of inter-converting structures inherently sampled by the free protein. Similarly, Al-Hashimi and co-workers used RDCs to demonstrate an equally impressive range of motion for tar-RNA 97. The resulting manifold of loop conformations was also consistent with conformational selection by its various binding targets (Figure 4e). In the case of RNA, a novel extension method was used to isolate the internal motion affects on the RDCs 98. Both examples of RDC-derived dynamics highlight the complementary power of RDCs of detect “supra-τc” dynamics 99 that are invisible to spin-relaxation, and central to biomolecular recognition.
Chemical Exchange and Spin-locked Relaxation Dispersion
Chemical exchange studies based on the isotropic chemical shift are “vintage” NMR methods for studying equilibrium dynamics. Nevertheless, their impact on protein dynamics has never been more powerful. Thanks to parallel advances in spin-physics, isotope-labeling, and computational chemistry, chemical exchange studies are revealing the importance of protein dynamics on microsecond-millisecond time-scale affecting receptor binding, folding-unfolding transitions, and enzyme catalysis.
The most common exchange experiments focus on the isotropic chemical shift. In the simplest case, a nucleus hops between two states with chemical shifts, δa and δb, equilibrium populations pa, and pb, and forward and backward rate constants kab = kexpb and kba = kexpa.
| (12) |
The exchange makes the chemical shift a fluctuating quantity, with consequences to the line shape and relaxation parameters that depend on how kex= kab + kba compares to the frequency difference in chemical shifts, Δω = γB0(δa−δb) 100–101. One speaks of fast, intermediate, or slow exchange on the chemical shift time scale for kex that is >, ~, or < Δω, respectively.
In fortuitous cases, the exchange rates are sufficiently slow, and the populations comparable, such that one observes separate, state-specific resonances. One can then apply 2-d exchange spectroscopy (EXSY) 102 to determine both the lifetimes and relative populations. 2-d heteronuclear exchange experiments on isotope-enriched samples are particularly effective for mapping slow exchange in proteins 103, and recent improvements in sensitivity and resolution have expanded their applicability to larger proteins 104–105.
More often, however, the exchange occurs between lopsided populations. This can cause excessive line broadening and decreased intensity of the minor state resonances, such that they, and the conformations they represent, are “invisible” 100,106–108. In these cases, heteronuclear (15N, 13C) transverse relaxation dispersion experiments, optimized for the relaxation properties of proteins 20,109–110 are extremely powerful. These experiments measure the dependence of 15N or 13C transverse relaxation on the strength of an applied radio-frequency spin-lock, quoted in frequency units. Nuclei with exchange rates comparable to the spin-lock frequencies show a “dispersion” curve – transverse relaxation rates that decrease with increasing spin-lock frequency (this dispersion is not to be confused with the B0 dispersion studies carried out with field-cycling). These dispersion curves are essentially the spectral density functions that describe the fluctuations in isotropic chemical shifts that result from the exchange dynamics. Fits of these curves to those predicted by exchange models yields the lifetimes of the exchange-coupled states, and, in optimal situations, the fractional populations and chemical shift differences. Examples of dispersion curves are in Figure 5a. Thus, the intrinsic chemical shift of the “invisible minor” state can be deduced from the dispersion of the observed major state. By far, the simplest models of two-state exchange 111–112 dominate the fitting of dispersion curves. Studies that go beyond the confines of the two-state model 113–114 are fewer, and in a younger stage of development, owing to the increased of number of fitting parameters required relative to the typical sensitivity and abundance of dispersion data.
Figure 5.
Example of spin-locked relaxation dispersion for characterizing “invisible” protein states. (a) 15N relaxation dispersion curves, R2,eff versus νCPMG, for residues of the FF domain, using Carr-Purcell-Meiboom-Gill (CPMG) spin-locking. The numbers in the upper refer to specific residues. Critically, only one set of NMR resonances are observed, that of the major “N” state (97%). (b) The curves are fit to functions based on two-state chemical exchange, to extract populations, rate constants, and conformational parameters of the invisible state, including chemical shifts, RDCs, and RCSAs. Together they enable the determination of the “invisible” conformation, the folding intermediate “I”. Figure 5a reprinted from Korzhnev, D. M.; Religa, T. L.; Banachewicz, W.; Fersht, A. R.; Kay, L. E. Science 2010, 329, 1312-6 with permission, and Figure 5b reprinted from Al-Hashimi in Science 2010, 329, 129. Copyright 2010 AAAS.
Traditionally, identifying the inter-converting conformations that give rise to the relaxation dispersion has been difficult. But this has changed. Retrospective analyses of protein chemical shift databases, paralleled by advances in quantum chemistry calculations (e.g. density functional theory) have dramatically increased our ability to predict conformation from chemical shifts and vice versa 115–118. As a consequence, the modest chemical shift has gone beyond a mere resonance index, to become a pre-eminent probe of structural dynamics. Notably, because the dispersion profiles from major state resonances can give the chemical shifts of minor “invisible” states, one can begin to model the conformations of the latter. This has opened the door to characterizing rare conformational excursions on the microsecond – millisecond time scale relevant for catalysis and folding (vide infra).
A growing number of relaxation dispersion studies implicate the role of millisecond conformational fluctuations with enzyme catalysis. Notable examples include Ribonuclease A (RNase A) 119, E. coli DHFR 120, and adenylate kinase 59,121. For RNase A, conformational exchange rates kex from dispersion agreed well with kcat and the rate constant for product release. Solvent kinetic isotope studies gave similar reductions of kex and kcat. Structural interpretations of the dispersion data suggested RNase A samples the “invisible” (< 5%) conformation of the following state. For DHFR, extensive dispersion studies showed similar behavior: within a given intermediate state of its catalytic cycle, the enzyme samples “invisible” conformations characteristic of the ground states of the preceding and following states. Furthermore, DHFR mutations that perturb millisecond exchange dynamics also perturb hydride transfer, but not substrate binding 122. Collectively, these findings suggest that the millisecond conformational fluctuations sensed by NMR influence enzyme catalysis. Yet, it is also well accepted that the atomic motions for the chemical step (crossing of the transition state free energy barrier) are orders of magnitude faster than the millisecond processes. Indeed, recent reviews 123–124 emphasize that millisecond enzyme motions cannot and do not influence the chemical step. The inconsistency of these two views is likely only apparent 125. The millisecond dynamics sensed by relaxation dispersion correspond to the motions of multiple moieties, whose individual motions might be quite fast (10−12 – 10−9 sec), but whose allied motion – an alliance essential for progressing along the reaction coordinate – is far less frequent. The relative infrequency of these cooperative motions manifests as longer millisecond dynamics sensed by NMR dispersion. Such motions are not the actual chemical step. Nevertheless, they reflect essential conformational changes (e.g. those associated with substrate binding and product release) that can be rate-limiting. As such, their characterization by NMR will continue to be important for understanding the overall catalytic process.
An exciting new method consists of relaxation dispersion studies carried out under conditions of weak protein alignment. These experiments can reveal invisible state properties beyond the chemical shift, including RDCs and RCSAs 126–127. In turn, this has enabled the determination of invisible state structures. In particular, Kay and co-workers demonstrated the proof-of-concept by determining the structure of an invisible state of SH3 domain 128; that corresponded to its ligand-bound form. They then applied a similar strategy to a de novo problem, in which they solved the structure of a transient folding intermediate of the FF domain 129, which was present at only 3% (cf. Figure 5b). Most recently, they have determined a transient state of a mutant T4 lysozyme 130.
Finally, the methyl-TROSY methods described above 80 have granted access to the millisecond chemical exchange dynamics of supra-molecular machines. Kay and co-workers have used relaxation dispersion 83 along with Paramagnetic Relaxation Enhancement (PRE) experiments 84 to characterize the functional motions of the 20S core-particle proteasome (MW ~ 670 kDA), responsible for timely protein degradation in the cell. More recently, they have characterized the folding-unfolding dynamics of small protein substrates within the proteasome 131. Many key biological processes in the cell involve supra-molecular machines, such as the proteasome. The methyl-TROSY dispersion methods now enable a detailed investigation of their inner moving parts.
Future Challenges
The above passages tried to give a sense of the means and power by which liquid-state NMR can illuminate functional protein dynamics. Of course, the field itself remains dynamic, with ample room for methods development and novel applications. What are some future “challenge areas” for dynamic NMR? Of course, the answers will vary violently from one person to the next; nevertheless, here are a few possibilities.
Characterizing correlated or collective motions in proteins is of great interest, since they are likely critical for the long-range site-to-site communication that support allosteric communication and signal transduction. Yet, characterizing these motions by NMR has been historically challenging. In particular, the spin-relaxation experiments above report on auto-correlation functions, which measure the orientational fluctuations of the same bond vector. But to detect correlated motions, we need cross-correlation functions that compare the fluctuations of pairs of distinct bond vectors. Most NMR relaxation experiments that are sensitive to cross-correlation functions exploit pairs of vectors that share a common origin, and thus naturally undergo correlated re-orientations. Such experiments include the aforementioned TROSY methods that have dramatically increased the sensitivity of NMR for larger biomolecules 80,132, and provided more accurate means for detecting chemical exchange 133–134. But from a correlated motion perspective, these experiments are not sufficient; greater interest lies in correlated motions between more distal vector pairs. Toward this end, multiple quantum relaxation measurements have proven most effective 135–139. Transverse relaxation limits the spatial range of these methods to NH/CH bond vectors separated by one or two bonds. Apparent correlated changes in methyl side-chain S2 suggest collective picosecond-nanosecond motions related to allostery 140–141. Other dynamics experiments probing slower time scale processes, including RDCs and hydrogen-bond scalar couplings 142 or relaxation dispersion 143–144, also point to correlated motion and site-to-site communication. Nevertheless, improved methods for the direct detection of longer range correlated motions, while retaining comprehensive site-by-site description of more standard NMR dynamics experiments, is an area waiting for further advances.
The challenge of detecting correlated motion draws attention to another area that will certainly intensify: the synergy between computation and NMR dynamics. Computational studies can explore how correlated or collective motions might leave distinct signatures on experimental observables. In this context a recent breakthrough is the work by Griesinger, Salvatella, and co-workers, 145 that generates an ensemble of structures consistent with conventional auto/cross-correlated spin-relaxation and RDC data as a means to identify long-range correlated motions that are not (yet) directly available from experiment. Additionally, the increase in the trajectory lengths of all-atom simulations 146, and newer “coarse-grained” methods for sampling conformational space 147–148 opens the door for understanding the slow time scale motions detected by NMR in far richer detail and with far less model bias. Longer trajectories, coupled with experimental data 149 should also help clarify the extent the functional motions can be attributed to a few “essential degrees of freedom”.
A third challenge is that of molecular context: how does the protein dynamics information gained from in vitro NMR translate to the cellular environment? Of course, if in vitro dynamics-function relationships correlate with cellular assays, then this lends biological credence to the former. Nevertheless, it is desirable to systematically investigate how phenomena such as cellular crowding, local variations in viscosity, and diffusion, might assist or interfere with protein functional motions. The burgeoning “in cell” NMR studies 150–151 raise the possibility of carrying out those investigations. At the moment, such studies must cope with technical difficulties associated with inhomogeneous broadening and comparatively low sensitivity. Conceivably, liquid-state in-cell methods, in alliance with the recent development of cellular solid-state NMR experiments 152, may provide substantial progress.
A fourth challenge area is to continue investigations of the interplay between protein evolution and conformational dynamics. Thanks to the advances in bioinformatics, there is now an abundance of tools to identify and analyze conserved protein sequences, and correlated changes therein. To understand the evolution of these sequences, it is reasonable to consider any and all properties pertaining to protein function and stability, including conformational dynamics. Specifically, it makes sense to investigate whether and how protein sequences might reflect the selection of functional dynamics, or promote better adaptability under conditions of stress. Such studies could be particularly important for understanding protein evolution related to drug-resistance, and thus better anticipate sites of drug-resistant mutations.
Finally, of interest to both fundamental and applied research is defining how protein dynamics information might benefit molecular design. The examples above clearly show the involvement of protein dynamics with binding and catalysis. So it might seem that protein dynamics investigations would be a mainstay in the design of novel therapeutics. But, historically, this has been a rarity. One bottleneck is that a detailed dynamics analysis often requires time investments beyond those accepted by high-throughput programs. Moreover, the information gained does not easily translate into metrics that can immediately guide synthetic chemistry. To become truly integrated in the design process, the dynamics information must go beyond an inventory of protein motions, to further describe concretely how it can enhance traditional methods for defining and predicting structure-activity relationships and pharmacokinetic properties. Parallel dynamics studies of the ligand are also valuable, as the ligands are the central objects of structure-activity studies. Efforts along these lines are growing, 153 but more systematic efforts are needed.
The study of protein dynamics is an effort to understand the behavior of complex dynamical systems. NMR has been an illuminating force in this effort, due to research efforts that maintain a deep interest in the physical basis of NMR. The techniques discussed in this Perspective are only a subspace within the larger space of dynamics experiments available from NMR. Nevertheless, it is hoped that this Perspective conveys a sense of the excitement and opportunities dynamic NMR holds for physical chemists. Indeed, in the study of biomolecular motion, liquid-state NMR is “alive and kicking”, like the dynamic molecules they study.
Acknowledgments
This work was supported in part by National Institutes of Health Grant RO1-GM083081. JWP is grateful to Kimberly A. Wilson, Thomas E. Frederick, and Michael W. Staude for discussions, critical reading, and assistance with figures. JWP thanks Prof. Patricia L. Clark, Prof J. Patrick Loria, Jill J. Bouchard, John S. Zintsmaster, and Cheryl L. Schairer for useful discussions. JWP also thanks the anonymous reviewers for their suggestions, and apologizes to investigators not cited because of space limitations.
Biography
Biography. Jeffrey (Jeff) W. Peng received his PhD in Biophysics from the University of Michigan in 1993, developing methods for studies of protein dynamics under Prof. Gerhard Wagner. Afterwards, he trained with Prof. Dr. Richard Ernst at ETH-Zürich, as a Damon Runyon Walter Winchell postdoctoral fellow. He then joined the NMR/Biophysics group at Vertex Pharmaceuticals Inc. Cambridge MA. In 2003, he joined the faculty at the University of Notre Dame, where he is now Associate Professor of Chemistry and Biochemistry, and a Concurrent Associate Professor in Physics. http://chemistry.nd.edu/faculty/detail/jpeng/
References
- 1.Bloch F, Hansen WW, Packard M. Phys Rev. 1946;69:127. [Google Scholar]
- 2.Purcell EM, Torrey HC, Pound RV. Phys Rev. 1946;69:37–38. [Google Scholar]
- 3.Renault M, Cukkemane A, Baldus M. Angew Chem Int Ed Engl. 2010;49:8346–57. doi: 10.1002/anie.201002823. [DOI] [PubMed] [Google Scholar]
- 4.Rugar D, Budakian R, Mamin HJ, Chui BW. Nature. 2004;430:329–32. doi: 10.1038/nature02658. [DOI] [PubMed] [Google Scholar]
- 5.Schroder L, Meldrum T, Smith M, Lowery TJ, Wemmer DE, Pines A. Phys Rev Lett. 2008;100:257603:1–4. doi: 10.1103/PhysRevLett.100.257603. [DOI] [PubMed] [Google Scholar]
- 6.Weber G. Adv Prot Chem. 1975;29:1–83. doi: 10.1016/s0065-3233(08)60410-6. [DOI] [PubMed] [Google Scholar]
- 7.Fiaux J, Bertelsen EB, Horwich AL, Wuthrich K. Nature. 2002;418:207–11. doi: 10.1038/nature00860. [DOI] [PubMed] [Google Scholar]
- 8.Kay LE. J Magn Reson. 2005;173:193–207. doi: 10.1016/j.jmr.2004.11.021. [DOI] [PubMed] [Google Scholar]
- 9.Redfield AG. IBM Journal. 1957 Jan;:19–31. [Google Scholar]
- 10.Wangsness RK, Bloch F. Phys Rev. 1953;89:728–39. [Google Scholar]
- 11.Abragam A. Principles of Nuclear Magnetism. Clarendon Press; 1961. [Google Scholar]
- 12.Ernst RR, Bodenhausen G, Wokaun A. Oxford Science Publications. 1987. [Google Scholar]
- 13.Lohr F, Reckel S, Karbyshev M, Connolly PJ, Abdul-Manan N, Bernhard F, Moore JM, Dotsch V. J Biomol NMR. 2012 doi: 10.1007/s10858-012-9601-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Tugarinov V, Kanelis V, Kay LE. Nat Protoc. 2006;1:749–54. doi: 10.1038/nprot.2006.101. [DOI] [PubMed] [Google Scholar]
- 15.Goldman M. J Magn Reson. 1984;60:437–52. [Google Scholar]
- 16.Tolman JR, Flanagan JM, Kennedy MA, Prestegard JH. Proc Natl Acad Sci U S A. 1995;92:9279–83. doi: 10.1073/pnas.92.20.9279. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Tjandra N, Bax A. Science. 1997;278:1111–4. doi: 10.1126/science.278.5340.1111. [DOI] [PubMed] [Google Scholar]
- 18.Tolman JR, Ruan K. Chem Rev. 2006;106:1720–36. doi: 10.1021/cr040429z. [DOI] [PubMed] [Google Scholar]
- 19.Igumenova TI, Frederick KK, Wand AJ. Chem Rev. 2006;106:1672–99. doi: 10.1021/cr040422h. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Palmer AG, 3rd, Massi F. Chem Rev. 2006;106:1700–19. doi: 10.1021/cr0404287. [DOI] [PubMed] [Google Scholar]
- 21.Mittermaier A, Kay LE. Science. 2006;312:224–8. doi: 10.1126/science.1124964. [DOI] [PubMed] [Google Scholar]
- 22.Peng JW, Wagner G. J Magn Reson. 1992a;98:308–22. [Google Scholar]
- 23.Millet O, Muhandiram DR, Skrynnikov NR, Kay LE. J Am Chem Soc. 2002;124:6439–48. doi: 10.1021/ja012497y. [DOI] [PubMed] [Google Scholar]
- 24.Skrynnikov NR, Millet O, Kay LE. J Am Chem Soc. 2002;124:6449–60. doi: 10.1021/ja012498q. [DOI] [PubMed] [Google Scholar]
- 25.Redfield AG. J Biomol NMR. 2012;52:159–77. doi: 10.1007/s10858-011-9594-1. [DOI] [PubMed] [Google Scholar]
- 26.Korb JP, Goddard Y, Pajski J, Diakova G, Bryant RG. J Phys Chem B. 2011;115:12845–58. doi: 10.1021/jp2053426. [DOI] [PubMed] [Google Scholar]
- 27.Lipari G, Szabo A. J Am Chem Soc. 1982a;104:4546–59. [Google Scholar]
- 28.Lipari G, Szabo A. J Am Chem Soc. 1982b;104:4546–59. [Google Scholar]
- 29.Halle B, Wennerstrom H. Journal of Chemical Physics. 1981;75:1928–43. [Google Scholar]
- 30.Dellwo MJ, Wand AJ. Journal of the American Chemical Society. 1989;111:4571–78. [Google Scholar]
- 31.Kinosita K, Jr, Kawato S, Ikegami A. Biophys J. 1977;20:289–305. doi: 10.1016/S0006-3495(77)85550-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Bruschweiler R, Wright PE. J Am Chem Soc. 1994;116:8426–27. [Google Scholar]
- 33.Bremi T, Brüschweiler R. J Am Chem Soc. 1997;119:6672–73. [Google Scholar]
- 34.Lienin SF, Bremi T, Brutscher B, Bruschweiler R, Ernst RR. J Am Chem Soc. 1998;120:9870–79. [Google Scholar]
- 35.Kay LE, Torchia DA, Bax A. Biochemistry. 1989;28:8972–79. doi: 10.1021/bi00449a003. [DOI] [PubMed] [Google Scholar]
- 36.Palmer AG, 3rd, Rance M, Wright PE. J Am Chem Soc. 1991;113:4371–80. [Google Scholar]
- 37.Mandel AM, Akke M, Palmer AG., 3rd J Mol Biol. 1995;246:144–63. doi: 10.1006/jmbi.1994.0073. [DOI] [PubMed] [Google Scholar]
- 38.Bruschweiler R, Liao X, Wright PE. Science. 1995;268:886–9. doi: 10.1126/science.7754375. [DOI] [PubMed] [Google Scholar]
- 39.Lee LK, Rance M, Chazin WJ, Palmer AG., 3rd J Biol NMR. 1997;9:287–98. doi: 10.1023/a:1018631009583. [DOI] [PubMed] [Google Scholar]
- 40.Tjandra N, Garrett DS, Gronenborn AM, Bax A, Clore GM. Nat Struct Biol. 1997;4:443–9. doi: 10.1038/nsb0697-443. [DOI] [PubMed] [Google Scholar]
- 41.Clore GM, Szabo A, Bax A, Kay LE, Driscoll PC, Gronenborn AM. J Am Chem Soc. 1990;112:4989–91. [Google Scholar]
- 42.Baber JL, Szabo A, Tjandra N. J Am Chem Soc. 2001;123:3953–59. doi: 10.1021/ja0041876. [DOI] [PubMed] [Google Scholar]
- 43.Ryabov YE, Fushman D. J Am Chem Soc. 2007;129:3315–27. doi: 10.1021/ja067667r. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Prompers JJ, Bruschweiler R. J Am Chem Soc. 2002;124:4522–34. doi: 10.1021/ja012750u. [DOI] [PubMed] [Google Scholar]
- 45.Ryabov Y, Clore GM, Schwieters CD. J Chem Phys. 2012;136:034108. doi: 10.1063/1.3675602. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Wong V, Case DA, Szabo A. Proc Natl Acad Sci U S A. 2009;106:11016–21. doi: 10.1073/pnas.0809994106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Meirovitch E, Shapiro YE, Polimeno A, Freed JH. Prog Nucl Magn Reson Spectrosc. 2010;56:360–405. doi: 10.1016/j.pnmrs.2010.03.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Meirovitch E, Polimeno A, Freed JH. J Chem Phys. 2010;132:207101. doi: 10.1063/1.3429599. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Meirovitch E, Shapiro YE, Zerbetto M, Polimeno A. J Phys Chem B. 2012;116:886–94. doi: 10.1021/jp208767s. [DOI] [PubMed] [Google Scholar]
- 50.Frederick KK, Sharp KA, Warischalk N, Wand AJ. J Phys Chem B. 2008;112:12095–103. doi: 10.1021/jp8038576. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Shapiro YE, Polimeno A, Freed JH, Meirovitch E. J Phys Chem B. 2011;115:354–65. doi: 10.1021/jp107130m. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Halle B. J Chem Phys. 2009;131:224507:1–22. doi: 10.1063/1.3269991. [DOI] [PubMed] [Google Scholar]
- 53.Nirmala N, Wagner G. J Am Chem Soc. 1988;110:7557–58. [Google Scholar]
- 54.Siggers K, Soto C, Palmer AG., 3rd Biophys J. 2007;93:2447–56. doi: 10.1529/biophysj.106.100578. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Peng T, Zintsmaster JS, Namanja AT, Peng JW. Nat Struct Mol Biol. 2007;14:325–31. doi: 10.1038/nsmb1207. [DOI] [PubMed] [Google Scholar]
- 56.Lacy ER, Filippov I, Lewis WS, Otieno S, Xiao L, Weiss S, Hengst L, Kriwacki RW. Nat Struct Mol Biol. 2004;11:358–64. doi: 10.1038/nsmb746. [DOI] [PubMed] [Google Scholar]
- 57.Sugase K, Dyson HJ, Wright PE. Nature. 2007;447:1021–5. doi: 10.1038/nature05858. [DOI] [PubMed] [Google Scholar]
- 58.Wang Y, Fisher JC, Mathew R, Ou L, Otieno S, Sublet J, Xiao L, Chen J, Roussel MF, Kriwacki RW. Nat Chem Biol. 2011;7:214–21. doi: 10.1038/nchembio.536. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Wolf-Watz M, Thai V, Henzler-Wildman K, Hadjipavlou G, Eisenmesser EZ, Kern D. Nat Struct Mol Biol. 2004;11:945–9. doi: 10.1038/nsmb821. [DOI] [PubMed] [Google Scholar]
- 60.Henzler-Wildman KA, Lei M, Thai V, Kerns SJ, Karplus M, Kern D. Nature. 2007;450:913–6. doi: 10.1038/nature06407. [DOI] [PubMed] [Google Scholar]
- 61.Zidek L, Novotny MV, Stone MJ. Nat Struct Biol. 1999;6:1118–21. doi: 10.1038/70057. [DOI] [PubMed] [Google Scholar]
- 62.Jarymowycz VA, Stone MJ. Chem Rev. 2006;106:1624–71. doi: 10.1021/cr040421p. [DOI] [PubMed] [Google Scholar]
- 63.Showalter SA, Johnson E, Rance M, Bruschweiler R. J Am Chem Soc. 2007;129:14146–7. doi: 10.1021/ja075976r. [DOI] [PubMed] [Google Scholar]
- 64.Trbovic N, Kim B, Friesner RA, Palmer AG., 3rd Proteins. 2008;71:684–94. doi: 10.1002/prot.21750. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Sareth S, Li H, Yamada H, Woodward CK, Akasaka K. FEBS Lett. 2000;470:11–4. doi: 10.1016/s0014-5793(00)01283-7. [DOI] [PubMed] [Google Scholar]
- 66.Massi F, Palmer AG., 3rd J Am Chem Soc. 2003;125:11158–9. doi: 10.1021/ja035605k. [DOI] [PubMed] [Google Scholar]
- 67.Johnson E, Palmer AG, 3rd , Rance M. Proteins. 2007;66:796–803. doi: 10.1002/prot.21274. [DOI] [PubMed] [Google Scholar]
- 68.Muhandiram DR, Yamazaki T, Sykes BD, Kay LE. J Am Chem Soc. 1995;117:11536–44. [Google Scholar]
- 69.Tugarinov V, Ollerenshaw JE, Kay LE. J Am Chem Soc. 2005;127:8214–25. doi: 10.1021/ja0508830. [DOI] [PubMed] [Google Scholar]
- 70.Lee AS, Kinnear SA, Wand AJ. Nat Struct Biol. 2000;7:72–77. doi: 10.1038/71280. [DOI] [PubMed] [Google Scholar]
- 71.Frederick KK, Marlow MS, Valentine KG, Wand AJ. Nature. 2007;448:325–9. doi: 10.1038/nature05959. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Akke M, Brüschweiler R, Palmer AG., III J Am Chem Soc. 1993;115:9832–33. [Google Scholar]
- 73.Li Z, Raychaudhuri S, Wand AJ. Protein Sci. 1996;5:2647–50. doi: 10.1002/pro.5560051228. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Yang D, Kay LE. J Mol Biol. 1996;263:369–82. doi: 10.1006/jmbi.1996.0581. [DOI] [PubMed] [Google Scholar]
- 75.Marlow MS, Dogan J, Frederick KK, Valentine KG, Wand AJ. Nat Chem Biol. 2010;6:352–8. doi: 10.1038/nchembio.347. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Li DW, Bruschweiler R. J Am Chem Soc. 2009;131:7226–7. doi: 10.1021/ja902477s. [DOI] [PubMed] [Google Scholar]
- 77.Fuentes EJ, Gilmore SA, Mauldin RV, Lee AL. J Mol Biol. 2006;364:337–51. doi: 10.1016/j.jmb.2006.08.076. [DOI] [PubMed] [Google Scholar]
- 78.Namanja AT, Peng T, Zintsmaster JS, Elson AC, Shakour MG, Peng JW. Structure. 2007;15:313–27. doi: 10.1016/j.str.2007.01.014. [DOI] [PubMed] [Google Scholar]
- 79.Cooper A, Dryden DT. Eur Biophys J. 1984;11:103–9. doi: 10.1007/BF00276625. [DOI] [PubMed] [Google Scholar]
- 80.Tugarinov V, Hwang PM, Ollerenshaw JE, Kay LE. J Am Chem Soc. 2003;125:10420–8. doi: 10.1021/ja030153x. [DOI] [PubMed] [Google Scholar]
- 81.Tugarinov V, Kay LE. Biochemistry. 2005;44:15970–7. doi: 10.1021/bi0519809. [DOI] [PubMed] [Google Scholar]
- 82.Chang YG, Kuo NW, Tseng R, LiWang A. Proc Natl Acad Sci U S A. 2011;108:14431–6. doi: 10.1073/pnas.1104221108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Sprangers R, Kay LE. Nature. 2007;445:618–22. doi: 10.1038/nature05512. [DOI] [PubMed] [Google Scholar]
- 84.Religa TL, Sprangers R, Kay LE. Science. 2010;328:98–102. doi: 10.1126/science.1184991. [DOI] [PubMed] [Google Scholar]
- 85.Bax A. Protein Sci. 2003;12:1–16. doi: 10.1110/ps.0233303. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Lipsitz RS, Tjandra N. Annu Rev Biophys Biomol Struct. 2004;33:387–413. doi: 10.1146/annurev.biophys.33.110502.140306. [DOI] [PubMed] [Google Scholar]
- 87.Saupe A. Angew Chem Internat Edit. 1968;7:97–112. [Google Scholar]
- 88.Losonczi JA, Andrec M, Fischer MW, Prestegard JH. J Magn Reson. 1999;138:334–42. doi: 10.1006/jmre.1999.1754. [DOI] [PubMed] [Google Scholar]
- 89.Dosset P, Hus JC, Marion D, Blackledge M. J Biomol NMR. 2001;20:223–31. doi: 10.1023/a:1011206132740. [DOI] [PubMed] [Google Scholar]
- 90.Zweckstetter M. Nat Protoc. 2008;3:679–90. doi: 10.1038/nprot.2008.36. [DOI] [PubMed] [Google Scholar]
- 91.Maciejewski M, Tjandra N, Barlow PN. Biochemistry. 2011;50:8138–49. doi: 10.1021/bi200575b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Meiler J, Prompers JJ, Peti W, Griesinger C, Bruschweiler R. J Am Chem Soc. 2001;123:6098–107. doi: 10.1021/ja010002z. [DOI] [PubMed] [Google Scholar]
- 93.Tolman JR. J Am Chem Soc. 2002;124:12020–30. doi: 10.1021/ja0261123. [DOI] [PubMed] [Google Scholar]
- 94.Bouvignies G, Markwick PR, Blackledge M. Proteins. 2008;71:353–63. doi: 10.1002/prot.21703. [DOI] [PubMed] [Google Scholar]
- 95.Salmon L, Bouvignies G, Markwick P, Lakomek N, Showalter S, Li DW, Walter K, Griesinger C, Bruschweiler R, Blackledge M. Angew Chem Int Ed Engl. 2009;48:4154–7. doi: 10.1002/anie.200900476. [DOI] [PubMed] [Google Scholar]
- 96.Lange OF, Lakomek NA, Fares C, Schroder GF, Walter KF, Becker S, Meiler J, Grubmuller H, Griesinger C, de Groot BL. Science. 2008;320:1471–5. doi: 10.1126/science.1157092. [DOI] [PubMed] [Google Scholar]
- 97.Zhang Q, Stelzer AC, Fisher CK, Al-Hashimi HM. Nature. 2007;450:1263–7. doi: 10.1038/nature06389. [DOI] [PubMed] [Google Scholar]
- 98.Zhang Q, Sun X, Watt ED, Al-Hashimi HM. Science. 2006;311:653–6. doi: 10.1126/science.1119488. [DOI] [PubMed] [Google Scholar]
- 99.Lakomek NA, Lange OF, Walter KF, Fares C, Egger D, Lunkenheimer P, Meiler J, Grubmuller H, Becker S, de Groot BL, et al. Biochem Soc Trans. 2008;36:1433–7. doi: 10.1042/BST0361433. [DOI] [PubMed] [Google Scholar]
- 100.Ishima R, Torchia DA. J Biomol NMR. 1999;14:369–72. doi: 10.1023/a:1008324025406. [DOI] [PubMed] [Google Scholar]
- 101.Palmer AG, 3rd , Kroenke CD, Loria JP. Methods Enzymol. 2001;339:204–38. doi: 10.1016/s0076-6879(01)39315-1. [DOI] [PubMed] [Google Scholar]
- 102.Jeener J, Meier BH, Bachmann P, Ernst RR. J Chem Phys. 1979;71:4546–53. [Google Scholar]
- 103.Montelione GT, Wagner G. J Am Chem Soc. 1989;111:3096–98. [Google Scholar]
- 104.Li Y, Palmer AG., 3rd J Biomol NMR. 2009;45:357–60. doi: 10.1007/s10858-009-9385-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Robson SA, Peterson R, Bouchard LS, Villareal VA, Clubb RT. J Am Chem Soc. 2010;132:9522–3. doi: 10.1021/ja1017865. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Blackledge MJ, Bruschweiler R, Griesinger C, Schmidt JM, Xu P, Ernst RR. Biochemistry. 1993;32:10960–74. doi: 10.1021/bi00092a005. [DOI] [PubMed] [Google Scholar]
- 107.Millet O, Loria JP, Kroenke CD, Pons M, Palmer AG., III J Am Chem Soc. 2000;122:2867–77. [Google Scholar]
- 108.Mulder FA, Mittermaier A, Hon B, Dahlquist FW, Kay LE. Nat Struct Biol. 2001;8:932–5. doi: 10.1038/nsb1101-932. [DOI] [PubMed] [Google Scholar]
- 109.Loria JP, Rance M, Palmer AG., III J Biolmol NMR. 1999;15:151–55. doi: 10.1023/a:1008355631073. [DOI] [PubMed] [Google Scholar]
- 110.Mulder FAA, de Graaf RA, Kaptein R, Boelens R. J Magn Reson. 1998;131:351–57. doi: 10.1006/jmre.1998.1380. [DOI] [PubMed] [Google Scholar]
- 111.Carver JP, Richards RE. J Magn Reson. 1972;6:89–105. [Google Scholar]
- 112.Trott O, Palmer AG., III J Magn Reson. 2002;154:157–60. doi: 10.1006/jmre.2001.2466. [DOI] [PubMed] [Google Scholar]
- 113.Grey MJ, Wang C, Palmer AG., 3rd J Am Chem Soc. 2003;125:14324–35. doi: 10.1021/ja0367389. [DOI] [PubMed] [Google Scholar]
- 114.Korzhnev DM, Bezsonova I, Lee S, Chalikian TV, Kay LE. J Mol Biol. 2009;386:391–405. doi: 10.1016/j.jmb.2008.11.055. [DOI] [PubMed] [Google Scholar]
- 115.Xu XP, Case DA. Biopolymers. 2002;65:408–23. doi: 10.1002/bip.10276. [DOI] [PubMed] [Google Scholar]
- 116.Cavalli A, Salvatella X, Dobson CM, Vendruscolo M. Proc Natl Acad Sci U S A. 2007;104:9615–20. doi: 10.1073/pnas.0610313104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Shen Y, Lange O, Delaglio F, Rossi P, Aramini JM, Liu G, Eletsky A, Wu Y, Singarapu KK, Lemak A, et al. Proc Natl Acad Sci U S A. 2008;105:4685–90. doi: 10.1073/pnas.0800256105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Mulder FA, Filatov M. Chem Soc Rev. 2010;39:578–90. doi: 10.1039/b811366c. [DOI] [PubMed] [Google Scholar]
- 119.Kovrigin EL, Loria JP. Biochemistry. 2006;45:2636–47. doi: 10.1021/bi0525066. [DOI] [PubMed] [Google Scholar]
- 120.Boehr DD, McElheny D, Dyson HJ, Wright PE. Science. 2006;313:1638–42. doi: 10.1126/science.1130258. [DOI] [PubMed] [Google Scholar]
- 121.Henzler-Wildman KA, Thai V, Lei M, Ott M, Wolf-Watz M, Fenn T, Pozharski E, Wilson MA, Petsko GA, Karplus M, et al. Nature. 2007;450:838–44. doi: 10.1038/nature06410. [DOI] [PubMed] [Google Scholar]
- 122.Bhabha G, Lee J, Ekiert DC, Gam J, Wilson IA, Dyson HJ, Benkovic SJ, Wright PE. Science. 2011;332:234–8. doi: 10.1126/science.1198542. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Pisliakov AV, Cao J, Kamerlin SC, Warshel A. Proc Natl Acad Sci U S A. 2009;106:17359–64. doi: 10.1073/pnas.0909150106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Adamczyk AJ, Cao J, Kamerlin SC, Warshel A. Proc Natl Acad Sci U S A. 2011;108:14115–20. doi: 10.1073/pnas.1111252108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Hammes GG, Benkovic SJ, Hammes-Schiffer S. Biochemistry. 2011;50:10422–30. doi: 10.1021/bi201486f. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Igumenova TI, Brath U, Akke M, Palmer AG., 3rd J Am Chem Soc. 2007;129:13396–7. doi: 10.1021/ja0761636. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Vallurupalli P, Hansen DF, Stollar E, Meirovitch E, Kay LE. Proc Natl Acad Sci U S A. 2007;104:18473–7. doi: 10.1073/pnas.0708296104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Vallurupalli P, Hansen DF, Kay LE. Proc Natl Acad Sci U S A. 2008;105:11766–71. doi: 10.1073/pnas.0804221105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Korzhnev DM, Religa TL, Banachewicz W, Fersht AR, Kay LE. Science. 2010;329:1312–6. doi: 10.1126/science.1191723. [DOI] [PubMed] [Google Scholar]
- 130.Bouvignies G, Vallurupalli P, Hansen DF, Correia BE, Lange O, Bah A, Vernon RM, Dahlquist FW, Baker D, Kay LE. Nature. 2011;477:111–4. doi: 10.1038/nature10349. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Ruschak AM, Religa TL, Breuer S, Witt S, Kay LE. Nature. 2010;467:868–71. doi: 10.1038/nature09444. [DOI] [PubMed] [Google Scholar]
- 132.Pervushin K, Riek R, Wider G, Wuthrich K. Proc Natl Acad Sci U S A. 1997;94:12366–71. doi: 10.1073/pnas.94.23.12366. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Tjandra N, Wingfield P, Stahl S, Bax A. J Biomol NMR. 1996;8:273–84. doi: 10.1007/BF00410326. [DOI] [PubMed] [Google Scholar]
- 134.Kroenke CD, Loria JP, Lee LK, Rance M, Palmer AG., III J Am Chem Soc. 1998;120:7905–15. [Google Scholar]
- 135.Reif B, Hennig M, Griesinger C. Science. 1997;276:1230–3. doi: 10.1126/science.276.5316.1230. [DOI] [PubMed] [Google Scholar]
- 136.Brutscher B, Skrynnikov NR, Bremi T, Bruschweiler R, Ernst RR. J Magn Reson. 1998;130:346–51. doi: 10.1006/jmre.1997.1312. [DOI] [PubMed] [Google Scholar]
- 137.Carlomagno T, Felli IC, Czech M, Fischer R, Sprinzl M, Griesinger C. J Am Chem Soc. 1999;121 [Google Scholar]
- 138.Kloiber K, Konrat R. J Biomol NMR. 2000;17:265–8. doi: 10.1023/a:1008393903160. [DOI] [PubMed] [Google Scholar]
- 139.Lundstrom P, Mulder FA, Akke M. Proc Natl Acad Sci U S A. 2005;102:16984–9. doi: 10.1073/pnas.0504361102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Petit CM, Zhang J, Sapienza PJ, Fuentes EJ, Lee AL. Proc Natl Acad Sci U S A. 2009;106:18249–54. doi: 10.1073/pnas.0904492106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141.Namanja AT, Wang XJ, Xu B, Mercedes-Camacho AY, Wilson KA, Etzkorn FA, Peng JW. Proc Natl Acad Sci U S A. 2011;108:12289–94. doi: 10.1073/pnas.1019382108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Bouvignies G, Bernado P, Meier S, Cho K, Grzesiek S, Bruschweiler R, Blackledge M. Proc Natl Acad Sci U S A. 2005;102:13885–90. doi: 10.1073/pnas.0505129102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.Tzeng SR, Kalodimos CG. Nature. 2009;462:368–72. doi: 10.1038/nature08560. [DOI] [PubMed] [Google Scholar]
- 144.Lipchock JM, Loria JP. Structure. 2010;18:1596–607. doi: 10.1016/j.str.2010.09.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Fenwick RB, Esteban-Martin S, Richter B, Lee D, Walter KF, Milovanovic D, Becker S, Lakomek NA, Griesinger C, Salvatella X. J Am Chem Soc. 2011;133:10336–9. doi: 10.1021/ja200461n. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.Lindorff-Larsen K, Piana S, Dror RO, Shaw DE. Science. 2011;334:517–20. doi: 10.1126/science.1208351. [DOI] [PubMed] [Google Scholar]
- 147.Bacallado S, Chodera JD, Pande V. J Chem Phys. 2009;131:045106:1–10. doi: 10.1063/1.3192309. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Morcos F, Chatterjee S, McClendon CL, Brenner PR, Lopez-Rendon R, Zintsmaster J, Ercsey-Ravasz M, Sweet CR, Jacobson MP, Peng JW, et al. PLoS Comput Biol. 2010;6:e1001015. doi: 10.1371/journal.pcbi.1001015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.Henzler-Wildman K, Kern D. Nature. 2007;450:964–72. doi: 10.1038/nature06522. [DOI] [PubMed] [Google Scholar]
- 150.Serber Z, Selenko P, Hansel R, Reckel S, Lohr F, Ferrell JE, Jr, Wagner G, Dotsch V. Nat Protoc. 2006;1:2701–9. doi: 10.1038/nprot.2006.181. [DOI] [PubMed] [Google Scholar]
- 151.Wang Q, Zhuravleva A, Gierasch LM. Biochemistry. 2011;50:9225–36. doi: 10.1021/bi201287e. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Renault M, Tommassen-van Boxtel R, Bos MP, Post JA, Tommassen J, Baldus M. Proc Natl Acad Sci U S A. 2012;109:4863–8. doi: 10.1073/pnas.1116478109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Namanja AT, Wang XJ, Xu B, Mercedes-Camacho AY, Wilson BD, Wilson KA, Etzkorn FA, Peng JW. J Am Chem Soc. 2010;132:5607–9. doi: 10.1021/ja9096779. [DOI] [PMC free article] [PubMed] [Google Scholar]





