Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2017 May 1.
Published in final edited form as: J Magn Reson. 2016 Feb 11;266:73–80. doi: 10.1016/j.jmr.2016.01.018

A Dynamic Look Backward and Forward

Arthur G Palmer III 1
PMCID: PMC4856014  NIHMSID: NIHMS759572  PMID: 26899226

Abstract

The 2015 Gunther Laukien Prize recognized solution NMR studies of protein dynamics and thermodynamics. This Perspective surveys aspects of the development and application of NMR spin relaxation for investigations of protein flexibility and function over multiple time scales in solution. Methods highlighted include analysis of overall rotational diffusion, theoretical descriptions of R1ρ relaxation, and molecular dynamics simulations to interpret NMR spin relaxation. Applications are illustrated for the zinc-finger domain Xfin-31, the calcium-binding proteins calbindin D9k and calmodulin, and the bZip transcription factor of GCN4.

Keywords: chemical exchange, conformational dynamics, kinetics, ligand binding, molecular dynamics simulation, nuclear magnetic resonance spectroscopy, spin relaxation

Introduction

The investigation of the dynamics of proteins, nucleic acids, and other biological macromolecules has a long history in NMR spectroscopy. The first nuclear spin relaxation measurement for a protein was a one-dimensional natural-abundance 13C R1 experiment for ribonuclease A reported in 1971 [1]. Of course, theoretical descriptions of spin relaxation essential for interpreting relaxation measurements already were established by the 1970’s with key papers establishing the Bloch-Wangness-Redfield (BWR) [2] and Bloch-McConnell [3] equations, which form the basis for theoretical descriptions of Carr-Purcell-Meiboom-Gill (CPMG) [4], R1ρ [5], and exchange spectroscopy [6]. Indeed, by 1970, both CPMG [7] and R1ρ [5] experiments had been used to determine the kinetic rate constant for the chair-to-boat isomerization of cyclohexane. However, the congestion of one-dimensional NMR spectra and the low sensitivity of natural abundance 13C NMR spectroscopy severely limited applications of what was widely known, at least to NMR spectroscopists, to be a potentially powerful technique for investigating the dynamic properties of biological molecules. Despite these limitations, critical aspects of protein statistical mechanics were elucidated, summarized for BPTI in an influential review by Wagner in 1983 [8]. Notably, 13C relaxation measurements demonstrated that certain aromatic rings in BPTI rotate rapidly, showing that the interiors of proteins have considerable plasticity. As one measure of the fundamental importance of NMR spin relaxation measurements for understanding protein properties, investigations of relaxation of aromatic groups continue to be pursued to this day [9].

The situation changed dramatically in the two-year period from 1988 to 1989. First, “inverse” 1H-detected techniques for measuring nuclear spin relaxation of 13C and 15N heteronuclei began to be developed. The first methods by Nirmala and Wagner were based on a double DEPT experiment for 13C and were sensitive enough for 13C natural abundance NMR spectroscopy of small soluble proteins [10, 11]. These initial reports were quickly followed by a seminal study of uniformly 15N enriched staphylococcal nuclease using INEPT-based experiments for 15N R1, 15N R2, and the heteronuclear {1H}-15N steady-state NOE [12]. The approach for data acquisition and analysis presented in this paper, although highly refined over the past two decades, remains the basic framework for contemporary work.

The developments in 1988–1989 revolutionized and invigorated studies of biological systems by spin relaxation and subsequent developments in experimental methods have enabled investigation of numerous spin types in proteins and nucleic acids [13]. With these advances, site-specific information on multiple time scales can be obtained on large molecules and molecular machines, ultimately up to ~one-half megadalton in pioneering work by Kay and co-workers [14]. Impressive studies of complex systems demonstrate that important information can be obtained from these measurements. At the same time, these experimental methods have diffused out of the laboratories of specialists and are being employed increasingly outside of academic settings [15].

The main goal of this Perspective is to summarize my own constributions to the development and application of spin relaxation methods, beginning with my introduction to the field as a postdoctoral scientist with Peter E. Wright at The Scripps Research Institute and continuing throughout my career at Columbia University. The 2015 Gunther Laukien Prize Lecture described some aspects of this work and is available at www.enc-conference.org/LaukienRecipients/PastRecipients/. In addition, recent Perspective [16] and Accounts [17] articles have summarized chemical exchange methods and applications of spin relaxation methods to enzymes, respectively. Accordingly, this Perspective focuses on the role of protein flexibility and dynamics in recognition of target molecules by proteins.

First results: Xfin-31 and Calbindin D9k

Investigations of the Zn-finger domain Xfin-31 [1820] and the EF-hand protein calbindin D9k [21, 22] provided early experimental insights and illustrate the research concerns and approaches that continue to be central to my research.

Initial studies of the 25-residue Xfin-31 protein ligated with Zn2+ used two-dimensional DEPT-based experiments to measure backbone and side chain 13C spin relaxation at natural abundance [18]. The first experimental work focused on the backbone 13Cα spins, analyzed with the Lipari-Szabo model-free formalism [23, 24]. Values of the 13C generalized order parameter, S2, are shown in Figure 1. Like the earlier example of staphylococcal nuclease [12], the results showed that bond vectors in secondary structures are highly ordered, with values of the generalized order parameter S2 > 0.85. The termini and loops were more heterogeneously disordered, with the most C-terminal residue having S2 = 0.30. The experimental results subsequently were compared to values calculated from a 100 ps molecular dynamics (MD) simulation in aqueous solution [19]. The MD simulations yielded values of S2 that were more than 0.10 larger than experimental values only for three residues: Tyr 1, Phe 10, and Lys 24. For the remaining residues, the average difference between experimental and simulated values was 0.013 with a sample deviation of 0.06. The two residues near the N- and C-termini have small secondary 13C shifts [25], suggesting that these residues are fairly disordered in solution and over-restrained in the MD simulation, perhaps reflecting the short trajectory length that could be achieved in 1992.

Figure 1.

Figure 1

Conformational flexibity of Xfin-31. (a) Structure of Xfin-31. The protein backbone is depicted in grey. The Zn2+ ion is shown as a red sphere; the Zn2+ ion is ligated by two Cys and two His residues. The Cys side chains are shown as yellow sticks and the His side chains as cyan sticks. (b) (●) Experimental and ( Inline graphic) simulated values of S2 for the 13Cα-1H bond vector. Experimental uncertainties were determined by Monte-Carlo simulations. (c) Correlation between experimental and simulated order parameters. Reprinted with permission from J. Am. Chem. Soc., 1992, 114, 9059–9067. Copyright 1992 American Chemical Society.

Chemical exchange contributions to R2 were observed for six of the backbone 13Cα resonances, five of these are either ligands to the Zn2+ ion or residues or adjacent to such residues. Exchange broadening also was observed for side chain 13Cα resonances of the Zn2+ ligand His 19 and His 23 and the adjacent residue Val 22. The fitted values of the chemical exchange broadening Rex are uncorrelated with the squares of the 13Cα secondary chemical shifts, suggesting that the exchange process is not rapid unfolding of the protein following transient loss of the Zn2+ ion. In these initial studies, further characterization of chemical exchange was not possible, but these phenomena sparked an interest that developed into a major aspect of subsequent work.

Comparison between apo and holo forms of the calcium-binding EF-hand protein calbindin D9k demonstrated the importance of NMR investigations of conformational flexibility for understanding molecular recognition by proteins. In collaboration with Walter Chazin and coworkers, backbone 15N relaxation rate constants were measured for the apo, (Cd2+)1, and (Ca2+)2 states of the protein [21, 22]. Calbindin D9k contains two metal-binding EF-hand motifs and Cd2+ binds to the C-terminal EF-hand, mimicking the unobservable state in which a single Ca2+ ion is bound to that site (because Ca2+, but not Cd2+, binding is cooperative). Subsequently Mäler and coworkers generated an Asn56Ala mutant that blocks binding to the C-terminal EF-hand, and thereby obtaining a (Ca2+)1 species with the metal ion occupying the N-terminal EF-hand [26]. As shown in Figure 2, backbone 15N order parameters show that the N-terminal EF-hand is ordered and the C-terminal EF-hand is highly disordered in the apo state. In contrast, both EF-hands are highly ordered in the holo state. In addition, the C-terminal EF-hand is ordered in both the (Cd2+)1 and (Ca2+)1 states. In the former case, rigidification is the direct result of ion binding in the C-terminal EF-hand. In the latter case, binding to the N-terminal site rigidifies the C-terminal site as a consequence of structural changes propagated through the β-strand interface between the two EF-hand motifs. The key conclusion is that one contribution to the allosteric phenomenon in calbindin D9k results from rigidification of the C-terminal binding site regardless of which binding site is occupied by the first Ca2+ ion.

Figure 2.

Figure 2

Effects of metal-ion binding on flexibility of calbindin D9k. Backbone 15N order parameters for (a) ( Inline graphic) apo and (●) (Ca2+)2 states of calbindin D9k. The inset shows the secondary structure of calbindin D9k, with the α-helices and two EF-hand binding sites labeled. Notably, the values of S2 are reduced in apo compared to C2+-loaded calbindin D9k in the C-terminal EF-hand (site II, residues 55–62). Correlation between the values of S2 for (Ca2+)2 and (b) (Cd2+)1 or (c) (Ca2+)1 states. The values for residues in Site II are shown as open circles and are equally ordered in all three states (in contrast to the apo state). Reprinted with permission from Biochemistry, 1993, 32, 9832–9844. Copyright 1993 American Chemical Society.

Acquisition and Analysis of Relaxation Data

Method developments in my laboratory for fast (ps-ns) time-scale dynamics in proteins have focused on pulse sequence development and optimization [2729], error analysis and model-selection in relaxation data analysis [30], determination of overall rotational diffusion tensors [3133], relationships between relaxation and conformational entropy [34, 35], and temperature dependence of intramolecular motions [3638]. For brevity, only recent results on the analysis of rotational diffusion tensors are discussed herein.

In an early study, the quadric method of Brüschweiler and coworkers [39] was extended to joint analysis of 13C and 15N relaxation rate constants to enable more thorough analysis of the overall rotational diffusion tensors of macromolecules, particularly in cases for which the distribution of orientations of 15N bond vectors is not uniform [31]. More recently, the quadric approach was extended to time-dependent diffusion tensors [32, 33]. For the simplest case of interchange between two conformational states described by unequal, but isotropic, diffusion tensors, the spectral density function is given by [40]:

J(ω)=25{So2Sf26D(6D)2+ω2+Sf2(1So2)τ11+ω2τ12+So2(1Sf2)τ21+ω2τ22+(1So2)(1Sf2)τ31+ω2τ32} (1)

in which:

k=[kex2+12(DADB)(k1k1)+36(DADB)2]1/2So2=12[1+(k1k1)(6DA6DB+k1k1)+4k1k1Y20(μAμB)kkex]6D=3(DA+DB)+(kexk)/2τ1=(6D+k)1τ2=(6D+1/τf)1τ3=(6D+k+1/τf)1 (2)

DA and DB are local isotropic diffusion constants for the two states (vide infra); kex = k1 + k−1, k1 and k−1 are the rate constants for transitions from state A to B, and from state B to A, respectively; and Sf2 are τf are the square of the generalized order parameter and effective internal correlation time for internal motions, respectively (assumed to be identical in states A and B and independent of overall rotational diffusion). In addition, μA and μB are unit vectors along the principal axis of the interaction tensor for the relaxation mechanism under consideration (most commonly, the bond vector between two nuclei) and Y20(Ω) is the Legendre polynomial.

The above results are generalized to non-isotropic diffusion tensors by replacing the isotropic diffusion tensors in Eqs. 1 and 2 with the local isotropic diffusion tensors obtained from the quadric analysis [32, 33]. Thus, the local isotropic diffusion constants in each state j = {A, B} are given by:

Dj=ejTQjej (3)

in which ej are the directions cosines defining the orientation of μι in the principal frame of the diffusion tensor, Qj is diagonal with elements Qxx,j = (Dyy,j + Dzz,j)/2, Qyy,j = (Dxx,j + Dzz,j)/2 and Qzz,j = (Dxx,j + Dyy,j)/2, and Dkk,j are the principal values of the diffusion tensor for the jth conformational state. For an axially symmetric diffusion tensor, with Dxx,j = Dyy,j, the errors in the quadric approximation are < 10% for 0.65 • Dzz,j / Dxx,j • 1.75.

If the analysis is limited to molecular sites that are highly ordered in both states, only the first two terms of Eq. 1 are significant and

R2R112{So26D+(1So2)τ1}/{So26D(6D)2+ωX2+(1So2)τ11+ωX2τ12} (4)

is independent of internal motions. In principle, Eq. 4 can be fit to experimentally determined relaxation rate constants to estimate the diffusion tensors and exchange rates. However, even Eq. 4 contains a large number of parameters, so that data analysis is simplified by prior knowledge, particularly of the putative structures of the exchanging conformations. As an illustration, Figure 3 shows experimental data for calmodulin recorded at 316 K by Chang and coworkers [41]. These data have been fit with Eq. 4 assuming (i) the N- and C-terminal EF-hand domains of calmodulin diffuse independently in state A, (ii) the molecule has a rigid conformation with stable central helix in state B, and (iii) interchange between states is in the fast-exchange limit, kex ≫ |DADB|. The diffusion tensors were calculated from the structures of the intact molecule and the N- and C-terminal domains using HydroPro [42]. With these assumptions, the experimental data are fit with a population of the ordered B state of 0.67.

Figure 3.

Figure 3

Relaxation arising from time-dependent diffusion tensors [32, 33]. (a) Schematic of dynamic parameters for protein in equilibrium between two states with different overall rotational diffusion tensors. The two conformations, denoted “A” and “B”, have equilibrium populations pA and pB, respectively. A representative amide bond is depicted in both conformations. The effective internal correlation time is τf and the square of the generalized order parameter is Sf2 for fast intramolecular motions of the amide bond. The amide bond is shown superimposed on the enzyme diffusion axes (dotted), denoted Dxx,j, Dyy,j, and Dzz,j, where j indicates either state A or B. The rate constant for transitions from open to closed conformations is k1, while the reverse rate constant is k−1. (b) Structure of the rigid dumbbell structure from PDB file 1CLL. The N-terminal domain (residues 4−73) is green, and the C-terminal domain (residues 83−148) is reddish-purple. (c) 15N spin relaxation data for calmodulin acquired at 800 MHz and 316 K. (black) Experimental R2/R1 ratios reported by Chang and co-workers [41]; (blue) R2/R1 ratios calculated for the rigid dumbbell structure; (green) R2/R1 ratios calculated for the N-terminal domain; (reddish-purple) R2/R1 ratios calculated for the C-terminal domain; and (orange) fitted model in the fast averaging limit between the rigid conformation and a conformation in which the central helix is disordered (consistent with the loss of data for residues in the central helix) and the N- and C-terminal domains reorient independently. The fitted model yields a population of the rigid dumbbell conformation of 0.67 in the fast-exchange limit.

Method developments in my laboratory for slow (μs-ms) time-scale dynamics have focused on pulse sequence development and optimization [43], theoretical analyses of spin-locking experiments outside the fast exchange limit [44], and interpretation of chemical exchange parameters [45, 46]. Contributions in these areas have been reviewed recently in this journal [16]. For brevity, only theoretical analysis of the R1ρ relaxation rate constant is discussed herein.

A series of results, beginning in 2002 have yielded a number of expressions of increasing accuracy for the R1ρ relaxation rate constant arising from two-site [44, 47, 48] and n-site [49] chemical exchange between conformational or chemical states. For two-site exchange, assuming that the differences in the intrinsic relaxation rate constants between the two states are small, the most accurate expression is given by the Laguerre approximant [48]:

R1ρ=R10cos2θ+R20sin2θ+sin2θpApBΔω2kexωA2ωB2/ωe2+kex2sin2θpApBΔω2{1+2kex2(pAωA2+pBωB2)ωA2ωB2+ωe2kex2} (5)

in which R10 and R20 are the population-average relaxation rate constants in the absence of chemical exchange (dominated by dipole-dipole, chemical shift anisotropy, and quadrupolar relaxation mechanisms in biological macromolecules); ω1 is the amplitude of the spin-lock radiofrequency field; tanθ=ω1/Ω¯; Ω¯=pAΩA+pBΩB is the population-average resonance offset frequency; pk and Ωk are the equilibrium population and the resonance offset frequency, respectively, for the kth state; kex = k1 + k−1; k1 and k−1 are the forward and reverse kinetic rate constants, respectively; Δω = ΩB − ΩA; ωe2=Ω¯2+ω12; and ωk2=Ωk2+ω12. Other results have been reported for cases in which the differences in transverse relaxation rate constants for different states are not negligible (relative to kex) [16, 50]. One particularly simple such expression is given by:

R1ρ=R1,A0cos2θ+R2,A0sin2θ+sin2θk1{Δω2+(ΔR20)2}k1+ΔR20(ωA2+k12)k1{ωB2+(k1+ΔR20)2}+ΔR20ω12 (6)

in which ΔR20=ΔR2,B0ΔR2,A0, R2k0 is the intrinsic transverse relaxation rate constant for the kth state, and differences between R1 relaxation rate constants have been neglected (and are usually small). Note that Eq. 6 corrects a typographical error in Eq. 11 of the original publication [16].

A simple derivation is provided below to illustrate the approaches used to obtain approximate expressions for R1ρ, because most existing derivations rely on direct expansion of the characteristic polynomial for the evolution matrix [44, 48, 50]. As shown in Eq. 5, R1ρ has the form:

R1ρ=R10cos2θ+R20sin2θλ (7)

in which λ is the largest (least negative) eigenvalue of the evolution matrix:

L+K=[LA00LB]+[k1k1k1k1]I (8)

in which

Lk=[0Ωk0Ωk0ω10ω10] (9)

and 0 and I are the zero and identity matrices, respectively, with dimensions set by context (3 × 3 in Eq. 8). The key step in the derivation is based on the observation that the matrix K can be factored as:

K=[k1k1k1k1]I=[II][k1Ik1I]=UVT (10)

The eigenvalues of L + K are determined by equating the characteristic polynomial to 0:

0=|L+KλI|=|LλI+UVT|=|I+VT(LλI)1U||LλI| (11)

in which |A| is the determinant of the matrix A and the last line is obtained using the matrix determinant lemma. A series of formal manipulations that utilize the block structure of L yields:

0=|I+VT(LλI)1U||LλI|=|I+VT(LλI)1U||LAλI||LBλI|=|I+kex[pB(LAλI)1+pA(LBλI)1]||LAλI||LBλI|=|(LAλI)(LBλI)+kex[pA(LAλI)+pB(LBλI)]|=|(LAλI)(LBλI)+kex(L¯λI)|=|(LALBkexL¯)λ(LA+LBkexI)+λ2| (12)

in which L¯=pALA+pBLB. Notably, the right-hand side of the last line now is the determinant of a matrix with dimension 3 × 3, rather than 6 × 6 as in the starting expression. When λ is small in magnitude, the determinant can be expanded to first order:

0=|(LALBkexL¯)λ(LA+LBkexI)+λ2|=|(LALBkexL¯)||Iλ(LALBkexL¯)1(LA+LBkexI)+(LALBkexL¯)1λ2|=|Iλ(LALBke xL¯)1(LA+LBkexI)+(LALBkexL¯)1λ2|=1λtr{(LALBkexL¯)1(LA+LBkexI)} (13)

in which tr{A} is the trace of the matrix A and terms proportional to λn for n • 2 are not considered. The final equation is solved for λ to give:

λ=tr{(LALBkexL¯)1(LA+LBkexI)} (14)

as a first-order approximation. Straightforward algebra then leads to:

λ=Rex=sin2θpApBΔω2kexωA2ωB2/ωe2+kex2 (15)

which, when substituted into Eq. 7, yields the result of Trott and Palmer [44]. Equation 14 remains valid if the expressions for LA and LB are generalized to include intrinsic relaxation rate constants that differ for the two states (except that Eq. 14 then yields the full expression for R1ρ, rather than Rex).

The key insight from Eqs. 7 and 15 is obtained by noting that as pb → 0, ωA2ωe2. Therefore, ωA2ωB2/ωe2ωB2; that is, the R1ρ experiment depends on the effective field in the rotating frame of reference for the B, usually sparsely populated and unobservable, state. As a result, the R1ρ experiment is sensitive to the signed value of the minor state chemical shifts, rather than the absolute value reported by the related Carr-Purcell-Meiboom-Gill experiment. This property facilitates structure determination of sparsely populated states of biological macromolecules, particularly in the weak rf field limit utilized in CEST [51] and DEST [52] experiments.

The accuracies of different approximate expressions for R1ρ calculated for a minor state population pB = 0.03 are compared in Figure 4. When the differences between intrinsic values of the relaxation rate constants are small, the Laguerre formula (Eq. 5) is very accurate (even for much large values of pB), while the Kay and Baldwin formula (Eq. 5 in [50]) and the simpler result of Eq. 6 have similar accuracies. When the difference in intrinsic relaxation rate constants is large, Eq. 5 becomes highly inaccurate, while the other two expressions maintain similar levels of accuracy.

Figure 4.

Figure 4

Theoretical calculations of R relaxation rate constants using (black) exact numerical calculation from the Bloch-McConnell equations, (orange) the Laguerre approximation of Eq. 5, (reddish-purple), the expression in Eq. 6, and green) the expression derived by Baldwin and Kay (Eq. 5 in [50]). (a) The calculations were performed by varying ω1/2π from 25 to 10,000 Hz with pb = 0.03, ΩA = 0 s−1, ΩB = 1000 s−1, kex = 1000 s−1, R1,A0=R1,B0=2.5s1, R2,A0=6s1, and R2,B0=12s1. (b) The calculations were performed as in (a) with the value of R2,B0=200s1. (c) The calculations were performed as in (a) except that ω1/2π = 100 Hz and kex was varied from 25 to 10,000 s−1. (d) The calculations were performed as in (c) with the value of R2,B0=200s1.

Application to Ligand-binding

Dynamics of the bZip domain of the yeast transcription factor GCN4 has been explored by both experimental NMR spin relaxation [53, 54] and MD simulations [55]. The most recent experimental data set, published in 2015 [54], and simulated trajectories, published in 2013, [55] illustrate the technical advances in these fields since the work described above on Xfin-31. Relaxation data were acquired for 15N spins in the bZip domain at four static magnetic fields, 600, 700, 800, and 900 MHz, compared to the single 500 MHz field used for the Xfin-31 investigations. In addition, the MD simulations consisted of four 100 ns trajectories, compared to the single 100 ps trajectory obtained for Xfin-31.

The initial experimental study of the bZip domain used 15N spin relaxation recorded at 500 MHz 1H Larmor frequency for three temperatures T = 290, 300, and 310 K [53]. These data only allowed S2 to be determined in a site-specific manner. Data were analyzed assuming a single average rotational correlation time for residues in the basic region given by the mean correlation time determined for residues in the ordered coiled-coil domain. More recently, relaxation data have been recorded at four static magnetic fields (600, 700, 800, and 900 MHz) at T = 300 K [54]. Importantly, spectral density mapping of the data recorded at four static magnetic fields suggested that the motional models required to describe the data would include at most five parameters, because the sets of spectral densities J(ωX) measured at the 15N and 1H Larmor frequencies were linear functions of ωX2, in which ωX is ωN or 0.87ωH, respectively (with the fifth parameter obtained from J(0)) Accordingly, relaxation data were fit with a series of model-free spectral density functions of which the most complex was the two-time-scale model-free function proposed by Clore and coworkers [56]. The values of S2 determined from the newer data set are shown in Figure 5a and the correlation between the two data sets is shown in Fig. 5b. As proposed previously [36, 57], the temperature dependence of the order parameters are characterized by the parameter Λ = d{ln(1 – S)/dlnT and values of Λ are shown in Fig. 5c.

Figure 5.

Figure 5

Backbone flexibility of the apo bZip domain of GCN4 [33, 54]. (a) 15N S2 determined from relaxation measurements at four static magnetic fields from 600 to 900 MHz (1H Larmor frequency) and T = 300. (b) Correlation between order parameters measured with a single magnetic field of 500 MHz and with four static magnetic fields at T = 300. (c) Temperature dependence of S2 characterized as Λ = d{ln(1 − S)}/dlnT, using data obtained at a single magnetic field of 500 MHz and T = 290, 300, and 310 K. Values of Λ could not be reliably determined for the highly ordered coil-coil residues. Regions of the protein are colored as follows: (reddish-purple) region (i) of basic region (residues 3–12), (green) region (ii) of basic region (residues 13–25); (black) coiled-coil (residues 26–55): black, and (orange) disordered C-term (residues 56–58). The molecular model shown in (a) is width coded according to the reciprocals of S2. The dashed lines in (c) correspond to the mean ± one standard deviation for the values of Λ in region (ii).

The protein displays four regions with distinct dynamic properties: (i) the most N-terminal residues 3–12 and (ii) residues 13–25 in the basic region; (iii) residues 26–55 in the coiled-coil domain, and the (iv) C-terminal residues 56–58. The N-terminal region (i) shows a strong N → C increase in both the values of S2 and Λ, while region (ii) shows a strong N → C increase in S2, but Λ appears relatively constant. The order parameters are large and uniform in the coiled-coil region, but the temperature variation was too small to be characterized. The C-terminal residues are disordered with a uniform value of Λ.

The role of the change in backbone flexibility in DNA binding was explored by estimating the conformational entropic penalty to binding using a theoretical approach originally developed by Akke and coworkers [58] and extended by Kay and coworkers [59] and by Wand and coworkers [60, 61]. Assuming that the order parameters for the basic region in the protein-DNA complex would be similar to those of the coiled-coil domain, the backbone conformational entropic penalty to binding was estimated to be −0.6 kJ mol−1 K−1, which is consistent with estimates derived from isothermal titration calorimetry [53].

The MD simulations of the GCN4 bZip domain were validated by comparison of chemical shifts calculated over the trajectories to the experimental values. Previous work on ribonuclease H had established that even 100 ns simulations were capable of reproducing the averaging of chemical shifts between different conformations [62]. The second of the four acquired trajectories best recapitulated the chemical shifts, and by inference the conformational preferences, of the basic region of the bZip domain, including for example a transiently formed helix-capping interaction between residues Ala 5 to Ala 9, Leu 6 to Arg 10, and Lys 7 to Asn 11. Using Sparta+ [63], the root-mean-square-deviations between calculated and experimental shifts were 0.63, 0.54, 0.13, and 1.38 ppm for 13Cα, 13C′, 1HN, and 15N, respectively.

The population of helical conformations was calculated as an average over the trajectory. The residue-specific populations are shown as a graph in Figure 6a. The two distinct segments of the basic region identified from the patterns of S2 and Λ contain two transiently populated helices: H1, corresponding to residues 4–8, and H2, corresponding to residues 13–19. Mapping of the helical populations onto the structure of the bZip domain/DNA complex in Figure 6b shows that transient helix H2 is part of the basic region that makes contacts in the major groove of the DNA ligand in the complex: residues Asn11, Ala14, Ala15, Ser18, and Arg19 are highly conserved among bZip domains and form base-specific DNA interactions in the complex. The good agreement between the calculated and experimental chemical shifts is accompanied by good agreement between calculated and experimental values of S2, shown in Figure 6c, further supporting the conclusion that the conformational dynamics observed in the MD simulation accurately reflect dynamics in the solution state of the GCN4 bZip domain.

Figure 6.

Figure 6

Molecular dynamics simulations of the apo bZip domain of GCN4. (a) Fractional helicity for each residue of GCN4 bZip domain observed in a 100 ns MD trajectory. Helical conformations were assigned for each MD snapshot using the program STRIDE [73]. Helix populations predicted from the experimental chemical shifts using δ2D [74] are shown as a dashed line. (b) Fractional helicity of each residue, as shown in (a) is mapped onto the crystal structure of the GCN4 bZip domain bound to DNA (PDB code 1YSA) according to the color bar shown. Side chains are displayed for residues that make contacts with the DNA in the x-ray crystal structure. (c) Comparison of (black) experimental and (orange) simulated order parameters (S2). Reprinted with permission from Chem. Theory Comput., 2013, 9, 5190–5200. Copyright 2013 American Chemical Society.

Investigations by NMR spin relaxation techniques, beginning with calbindin D9k, have addressed both thermodynamic and kinetic aspects of ligand binding by proteins. The observation of a populated helix H2 in the center of the DNA-recognition element of the bZip domain of GCN4 suggests a two-step mechanism for binding: (i) initial formation of an encounter complex through interaction with the pre-formed helix H2 and (ii) subsequent folding of the rest of the basic region to form the high affinity helical structure. This model can be represented as a “selected-fit” initial step followed by an “induced-fit” second step or as “coupled-folding-and-binding” [64, 65]. In general, ligand recognition by proteins can occur through parallel mechanisms with varying aspects of selected- or induced-fit models [66]. In the present case, the thermodynamic entropic penalty to binding is reduced by the presence of helix-capping sequences and transient helical conformations in the apo state of the bZip domain that increase the measured order parameters compared to a random coil peptide (e.g. the most C-terminal residues).

The experimental relaxation data for the basic region residues were fit with the extended model-free formalism, which characterizes internal dynamics on two time scales. For the bZip domain of GCN4, the faster dynamics are characterized by values of Sf20.6 for region (i) (residues 2–12) that increase across region (ii) (residues 13–25) to a maximum value of ~0.8. The average effective internal correlation time for the basic region is τf ≈ 50 ps. In contrast, the slower time scale dynamics are characterized by values of Ss2 that increase dramatically from 0.09 for residue 3 to 0.80 for residue 25 with average internal correlation times of τf ≈ 1.5 ns. These results suggest that the dominant conformational dynamic process required to reach the fully ordered state involves restriction of the conformational states sampled on the 1.5 ns time scale. This time scale is within the range expected for propagation of an α-helix and faster than the expected dissociation time constant for an encounter complex.

Discussion and Conclusion

The nearly three decades of modern applications of NMR spectroscopy to spin-relaxation measurements in proteins have seen dramatic developments in experimental and theoretical methods and equally dramatic advances in the complexity of biological systems under investigation [13]. Work in my own laboratory has focused on spin-relaxation studies of proteins in solution; however, the methods and experimental approaches pioneered in solution NMR spectroscopy have been adapted to nucleic acids in solution [6769] and to solid-state NMR studies of biological macromolecules [7072]. These advances suggest that the renaissance in NMR spin relaxation of biological macromolecules that began in 1988 is still vibrant and is likely to generate additional Gunther Laukien Prize awards in the future.

Highlights.

  • Historical review of NMR spin relaxation in proteins

  • Treatment of time-dependent rotational diffusion

  • Derivation of R1ρ relaxation rate constant

  • Contributions of flexibility to DNA binding by the GCN4 transcription factor

Acknowledgments

I am grateful to the current and past members of my research group for their invaluable contributions to my career at Columbia University. I thank Mark Rance (Univ. Cincinnati) for a longstanding collaboration that has been crucial to many aspects of my work and Peter E. Wright (The Scripps Research Institute) for his mentorship since I joined his research group as a National Science Foundation Postdoctoral Fellow in 1989. My undergraduate and graduate advisors, Colin F. Mackay (Haverford College) and Nancy L. Thompson (Univ. North Carolina), respectively, catalyzed my interest in dynamic fluctuations in biology and chemistry, the dominant theme throughout my research career.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Allerhand A, Doddrell D, Glushko V, Cochran DW, Wenkert E, Lawson PJ, Gurd FRN. Conformation and segmental motion of native and denatured ribonuclease A in solution. Application of natural-abundance carbon-13 partially relaxed Fourier transform nuclear magnetic resonance. J Amer Chem Soc. 1971;93:544–546. doi: 10.1021/ja00731a053. [DOI] [PubMed] [Google Scholar]
  • 2.Redfield AG. The theory of relaxation processes. Adv Magn Reson. 1965;1:1–32. [Google Scholar]
  • 3.McConnell HM. Reaction rates by nuclear magnetic resonance. J Chem Phys. 1958;28:430–431. [Google Scholar]
  • 4.Carver JP, Richards RE. A general two-site solution for the chemical exchange produced dependence of T2 upon the Carr-Purcell pulse separation. J Magn Reson. 1972;6:89–105. [Google Scholar]
  • 5.Deverell C, Morgan RE, Strange JH. Studies of chemical exchange by nuclear magnetization relaxation in the rotating frame. Mol Phys. 1970;18:553–559. [Google Scholar]
  • 6.Forsén S, Hoffman RA. Study of moderately rapid chemical exchange reactions by means of nuclear magnetic double resonance. J Chem Phys. 1963;39:2892. [Google Scholar]
  • 7.Allerhand A, Chen FM, Gutowsky HS. Spin-echo NMR studies of chemical exchange. 3. Conformational isomerization of cyclohexane and D11-cyclohexane. J Chem Phys. 1965;42:3040. doi: 10.1063/1.1696165. [DOI] [PubMed] [Google Scholar]
  • 8.Wagner G. Characterization of the distribution of internal motions in the basic pancreatic trypsin inhibitor using a large number of internal NMR probes. Q Rev Biophys. 1983;16:1–57. doi: 10.1017/s0033583500004911. [DOI] [PubMed] [Google Scholar]
  • 9.Kasinath V, Fu Y, Sharp KA, Wand AJ. A sharp thermal transition of fast aromatic-ring dynamics in ubiquitin, Angew. Chem Int Ed Engl. 2015;54:102–107. doi: 10.1002/anie.201408220. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Nirmala NR, Wagner G. Measurement of 13C relaxation times in proteins by two-dimensional heteronuclear 1H-13C correlation spectroscopy. J Am Chem Soc. 1988;110:7557–7558. [Google Scholar]
  • 11.Nirmala NR, Wagner G. Measurement of 13C spin-spin relaxation times by two-dimensional heteronuclear 1H-13C correlation spectroscopy. J Magn Reson. 1989;82:659–661. [Google Scholar]
  • 12.Kay LE, Torchia DA, Bax A. Backbone dynamics of proteins as studied by 15N inverse detected heteronuclear NMR spectroscopy: Applications to Staphylococcal nuclease. Biochemistry. 1989;28:8972–8979. doi: 10.1021/bi00449a003. [DOI] [PubMed] [Google Scholar]
  • 13.Palmer AG. NMR spectroscopy: NMR relaxation methods. In: Dyson HJ, editor. Biophysical Techniques for Structural Characterization of Macromolecules. Academic Press; Oxford: 2012. pp. 216–244. [Google Scholar]
  • 14.Religa TL, Sprangers R, Kay LE. Dynamic regulation of archaeal proteasome gate opening as studied by TROSY NMR. Science. 2010;328:98–102. doi: 10.1126/science.1184991. [DOI] [PubMed] [Google Scholar]
  • 15.Phillips AH, Schoeffler AJ, Matsui T, Weiss TM, Blankenship JW, Zobel K, Giannetti AM, Dueber EC, Fairbrother WJ. Internal motions prime cIAP1 for rapid activation. Nat Struct Mol Biol. 2014;21:1068–1074. doi: 10.1038/nsmb.2916. [DOI] [PubMed] [Google Scholar]
  • 16.Palmer AG. Chemical exchange in biomacromolecules: past, present, and future. J Magn Reson. 2014;241:3–17. doi: 10.1016/j.jmr.2014.01.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Palmer AG. Enzyme dynamics from NMR spectroscopy. Acc Chem Res. 2015;48:457–465. doi: 10.1021/ar500340a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Palmer AG, Rance M, Wright PE. Intramolecular motions of a zinc finger DNA-binding domain from xfin characterized by proton-detected natural abundance 13C heteronuclear NMR spectroscopy. J Am Chem Soc. 1991;113:4371–4380. [Google Scholar]
  • 19.Palmer AG, Case DA. Molecular dynamics analysis of NMR relaxation in a zinc-finger peptide. J Am Chem Soc. 1992;114:9059–9067. [Google Scholar]
  • 20.Palmer AG, Hochstrasser RA, Millar DP, Rance M, Wright PE. Characterization of amino acid side chain dynamics in a zinc-finger peptide using 13C NMR spectroscopy and time-resolved fluorescence spectroscopy. J Am Chem Soc. 1993;115:6333–6345. [Google Scholar]
  • 21.Kördel J, Skelton NJ, Akke M, Palmer AG, Chazin WJ. Backbone dynamics of calcium-loaded calbindin D9k studied by two-dimensional proton-detected NMR spectroscopy. Biochemistry. 1992;31:4856–4866. doi: 10.1021/bi00135a017. [DOI] [PubMed] [Google Scholar]
  • 22.Akke M, Kördel J, Skelton NJ, Palmer AG, Chazin WJ. Effects of ion binding on the backbone dynamics of calbindin D9k determined by 15N relaxation. Biochemistry. 1993;32:9832–9844. doi: 10.1021/bi00088a039. [DOI] [PubMed] [Google Scholar]
  • 23.Lipari G, Szabo A. Model-free approach to the interpretation of nuclear magnetic resonance relaxation in macromolecules. I. Theory and range of validity. J Am Chem Soc. 1982;104:4546–4559. [Google Scholar]
  • 24.Lipari G, Szabo A. Model-free approach to the interpretation of nuclear magnetic resonance relaxation in macromolecules. 2. Analysis of experimental results. J Am Chem Soc. 1982;104:4559–4570. [Google Scholar]
  • 25.Lee MS, Palmer AG, Wright PE. Relationship between 1H and 13C NMR chemical shifts and the secondary and tertiary structure of a zinc finger peptide. J Biomol NMR. 1992;2:307–322. doi: 10.1007/BF01874810. [DOI] [PubMed] [Google Scholar]
  • 26.Mäler L, Blankenship J, Rance M, Chazin WJ. Site-site communication in the EF-hand Ca2+-binding protein calbindin D9k. Nat Struct Biol. 2000;7:245–250. doi: 10.1038/73369. [DOI] [PubMed] [Google Scholar]
  • 27.Mandel AM, Palmer AG. Measurement of relaxation rate constants using constant-time accordion heteronuclear NMR spectroscopy. J Magn Reson, Ser A. 1994;110:62–72. [Google Scholar]
  • 28.Carr PA, Fearing DA, Palmer AG. 3D accordion spectroscopy for measuring 15N and 13CO relaxation rates in poorly-resolved NMR spectra. J Magn Reson. 1998;132:25–33. doi: 10.1006/jmre.1998.1374. [DOI] [PubMed] [Google Scholar]
  • 29.Kroenke CD, Loria JP, Lee LK, Rance M, Palmer AG. Longitudinal and transverse 1H-15N dipolar/15N chemical shift anisotropy relaxation interference: unamiguous determination of rotational diffusion tensors and chemical exchange effects in biological macromolecules. J Am Chem Soc. 1998;120:7905–7915. [Google Scholar]
  • 30.Mandel AM, Akke M, Palmer AG. Backbone dynamics of Escherichia coli ribonuclease HI: Correlations with structure and function in an active enzyme. J Mol Biol. 1995;246:144–163. doi: 10.1006/jmbi.1994.0073. [DOI] [PubMed] [Google Scholar]
  • 31.Lee LK, Rance M, Chazin WJ, Palmer AG. Rotational diffusion anisotropy of proteins from simultaneous analysis of 15N and 13C nuclear spin relaxation. J Biomol NMR. 1997;9 doi: 10.1023/a:1018631009583. [DOI] [PubMed] [Google Scholar]
  • 32.Gill ML, Palmer AG. Local isotropic diffusion approximation for coupled internal and overall molecular motions in NMR spin relaxation. J Phys Chem B. 2014;118:11120–11128. doi: 10.1021/jp506580c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Gill ML, Palmer AG. Correction to “Local isotropic diffusion approximation for coupled internal and overall molecular motions in NMR spin relaxation”. J Phys Chem B. 2015;119:3333. doi: 10.1021/acs.jpcb.5b00764. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Akke M, Brüschweiler R, Palmer AG. NMR order parameters and free energy: An analytic approach and application to cooperative Ca2+ binding by calbindin D9k. J Am Chem Soc. 1993;115:9832–9833. [Google Scholar]
  • 35.Trbovic N, Cho JH, Abel R, Friesner RA, Rance M, Palmer AG. Protein side-chain dynamics and residual conformational entropy. J Am Chem Soc. 2009;131:615–622. doi: 10.1021/ja806475k. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Vugmeyster L, Trott O, McKnight CJ, Raleigh DP, Palmer AG. Temperature-dependent dynamics of the villin headpiece helical subdomain, an unusually small thermostable protein. J Mol Biol. 2002;320:841–854. doi: 10.1016/S0022-2836(02)00537-5. [DOI] [PubMed] [Google Scholar]
  • 37.Johnson E, Palmer AG, Rance M. Temperature dependence of the NMR generalized order parameter. Proteins. 2007;66:796–803. doi: 10.1002/prot.21274. [DOI] [PubMed] [Google Scholar]
  • 38.Mandel AM, Akke M, Palmer AGr. Dynamics of ribonuclease H: temperature dependence of motions on multiple time scales. Biochemistry. 1996;35:16009–16023. doi: 10.1021/bi962089k. [DOI] [PubMed] [Google Scholar]
  • 39.Brüschweiler R, Liao X, W PE. Long-range motional restrictions in a multidomain zinc-finger protein from anisotropic tumbling. Science. 1995;268:886–889. doi: 10.1126/science.7754375. [DOI] [PubMed] [Google Scholar]
  • 40.Wong V, Case DA, Szabo A. Influence of the coupling of interdomain and overall motions on NMR relaxation. Proc Natl Acad Sci U S A. 2009;106:11016–11021. doi: 10.1073/pnas.0809994106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Chang SL, Szabo A, Tjandra N. Temperature dependence of domain motions of calmodulin probed by NMR relaxation at multiple fields. J Am Chem Soc. 2003;125:11379–11384. doi: 10.1021/ja034064w. [DOI] [PubMed] [Google Scholar]
  • 42.Ortega A, Amoros D, Garcia de la Torre J. Prediction of hydrodynamic and other solution properties of rigid proteins from atomic- and residue-level models. Biophys J. 2011;101:892–898. doi: 10.1016/j.bpj.2011.06.046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Loria JP, Rance M, Palmer AG. A relaxation-compensated Carr-Purcell-Meiboom-Gill sequence for characterizing chemical exchange by NMR spectroscopy. J Am Chem Soc. 1999;121:2331–2332. [Google Scholar]
  • 44.Trott O, Palmer AG. R1 relaxation outside of the fast-exchange limit. J Magn Reson. 2002;154:157–160. doi: 10.1006/jmre.2001.2466. [DOI] [PubMed] [Google Scholar]
  • 45.Cho J-H, O’Connell N, Raleigh DP, Palmer AG. ϕ-value analysis for ultrafast folding proteins by NMR relaxation dispersion. J Am Chem Soc. 2010;132:450–451. doi: 10.1021/ja909052h. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Grey MJ, Wang CY, Palmer AG. Disulfide bond isomerization in basic pancreatic trypsin inhibitor: Multisite chemical exchange quantified by CPMG relaxation dispersion and chemical shift modeling. J Am Chem Soc. 2003;125:14324–14335. doi: 10.1021/ja0367389. [DOI] [PubMed] [Google Scholar]
  • 47.Trott O, Abergel D, Palmer AG. An average-magnetization analysis of R1 relaxation outside of the fast exchange. Mol Phys. 2003;101:753–763. [Google Scholar]
  • 48.Miloushev VZ, Palmer AG. R1 relaxation for two-site chemical exchange: General approximations and some exact solutions. J Magn Reson. 2005;177:221–227. doi: 10.1016/j.jmr.2005.07.023. [DOI] [PubMed] [Google Scholar]
  • 49.Trott O, Palmer AG. Theoretical study of R1 rotating-frame and R2 free-precession relaxation in the presence of n-site chemical exchange. J Magn Reson. 2004;170:104–112. doi: 10.1016/j.jmr.2004.06.005. [DOI] [PubMed] [Google Scholar]
  • 50.Baldwin AJ, Kay LE. An R1 expression for a spin in chemical exchange between two sites with unequal transverse relaxation rates. J Biomol NMR. 2013;55:211–218. doi: 10.1007/s10858-012-9694-6. [DOI] [PubMed] [Google Scholar]
  • 51.Vallurupalli P, Bouvignies G, Kay LE. Studying “invisible” excited protein states in slow exchange with a major state conformation. J Am Chem Soc. 2012;134:8148–8161. doi: 10.1021/ja3001419. [DOI] [PubMed] [Google Scholar]
  • 52.Fawzi NL, Ying J, Torchia DA, Clore GM. Probing exchange kinetics and atomic resolution dynamics in high-molecular-weight complexes using dark-state exchange saturation transfer NMR spectroscopy. Nat Protoc. 2012;7:1523–1533. doi: 10.1038/nprot.2012.077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Bracken C, Carr PA, Cavanagh J, Palmer AG. Temperature dependence of intramolecular dynamics of the basic leucine zipper of GCN4: Implications for the entropy of association with DNA. J Mol Biol. 1999;285:2133–2146. doi: 10.1006/jmbi.1998.2429. [DOI] [PubMed] [Google Scholar]
  • 54.Gill ML, Byrd RA, Palmer AG. Dynamics of GCN4 facilitate DNA interaction: A model-free analysis of an intrinsically disordered region. Phys Chem Chem Phys. 2016 doi: 10.1039/c5cp06197k. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Robustelli P, Trbovic N, Friesner RA, Palmer AG. Conformational dynamics of the partially disordered yeast transcription factor GCN4. J Chem Theory Comput. 2013;9 doi: 10.1021/ct400654r. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Clore GM, Szabo A, Bax A, Kay LE, Driscoll PC, Gronenborn AM. Deviations from the simple two-parameter model-free approach to the interpretation of nitrogen-15 nuclear relaxation of proteins. J Am Chem Soc. 1990;112:4989–4991. [Google Scholar]
  • 57.Massi F, Wang C, Palmer AG. Solution NMR and computer simulation studies of active site loop motion in triosephosphate isomerase. Biochemistry. 2006;45:10787–10794. doi: 10.1021/bi060764c. [DOI] [PubMed] [Google Scholar]
  • 58.Akke M, Brüschweiler R, Palmer AG. NMR order parameters and free energy: An analytic approach and application to cooperative Ca2+ binding by calbindin D9k. J Am Chem Soc. 1993;115:9832–9833. [Google Scholar]
  • 59.Yang DW, Kay LE. Contributions to conformational entropy arising from bond vector fluctuations measured from NMR-derived order parameters: Application to protein folding. J Mol Biol. 1996;263:369–382. doi: 10.1006/jmbi.1996.0581. [DOI] [PubMed] [Google Scholar]
  • 60.Li Z, Raychaudhuri S, Wand AJ. Insights into the local residual entropy of proteins provided by NMR relaxation. Protein Sci. 1996;5:2647–2650. doi: 10.1002/pro.5560051228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Sharp KA, O’Brien E, Kasinath V, Wand AJ. On the relationship between NMR-derived amide order parameters and protein backbone entropy changes. Proteins. 2015;83:922–930. doi: 10.1002/prot.24789. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Robustelli P, Stafford KA, Palmer AG. Interpreting protein structural dynamics from NMR chemical shifts. J Am Chem Soc. 2012;134:6365–6374. doi: 10.1021/ja300265w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Shen Y, Bax A. SPARTA+: a modest improvement in empirical NMR chemical shift prediction by means of an artificial neural network. J Biomol NMR. 2010;48:13–22. doi: 10.1007/s10858-010-9433-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Kiefhaber T, Bachmann A, Jensen KS. Dynamics and mechanisms of coupled protein folding and binding reactions. Curr Opin Struct Biol. 2012;22:21–29. doi: 10.1016/j.sbi.2011.09.010. [DOI] [PubMed] [Google Scholar]
  • 65.Daniels KG, Tonthat NK, McClure DR, Chang YC, Liu X, Schumacher MA, Fierke CA, Schmidler SC, Oas TG. Ligand concentration regulates the pathways of coupled protein folding and binding. J Am Chem Soc. 2014;136:822–825. doi: 10.1021/ja4086726. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Miloushev VZ, Bahna F, Ciatto C, Ahlsen G, Honig B, Shapiro L, Palmer AG. Dynamic properties of a type II cadherin adhesive domain: Implications for the mechanism of strand-swapping of classical cadherins. Structure. 2008;16:1195–1205. doi: 10.1016/j.str.2008.05.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Akke M, Fiala R, Jiang F, Patel D, Palmer AG. Base dynamics in a UUCG tetraloop RNA hairpin characterized by 15N spin relaxation: Correlations with structure and stability. RNA. 1997;3:702–709. [PMC free article] [PubMed] [Google Scholar]
  • 68.Al-Hashimi HM. NMR studies of nucleic acid dynamics. J Magn Reson. 2013;237:191–204. doi: 10.1016/j.jmr.2013.08.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Xue Y, Kellogg D, Kimsey IJ, Sathyamoorthy B, Stein ZW, McBrairty M, Al-Hashimi HM. Characterizing RNA excited states using NMR relaxation dispersion. Methods Enzymol. 2015;558:39–73. doi: 10.1016/bs.mie.2015.02.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Palmer AG, Williams J, McDermott A. Nuclear magnetic resonance studies of biopolymer dynamics. J Phys Chem. 1996;100:13293–13310. [Google Scholar]
  • 71.Lewandowski JR. Advances in solid-state relaxation methodology for probing site-specific protein dynamics. Acc Chem Res. 2013;46:2018–2027. doi: 10.1021/ar300334g. [DOI] [PubMed] [Google Scholar]
  • 72.Lewandowski JR, Halse ME, Blackledge M, Emsley L. Direct observation of hierarchical protein dynamics. Science. 2015;348:578–581. doi: 10.1126/science.aaa6111. [DOI] [PubMed] [Google Scholar]
  • 73.Frishman D, Argos P. Knowledge-based protein secondary structure assignment. Proteins. 1995;23:566–579. doi: 10.1002/prot.340230412. [DOI] [PubMed] [Google Scholar]
  • 74.Camilloni C, De Simone A, Vranken WF, Vendruscolo M. Determination of secondary structure populations in disordered states of proteins using nuclear magnetic resonance chemical shifts. Biochemistry. 2012;51:2224–2231. doi: 10.1021/bi3001825. [DOI] [PubMed] [Google Scholar]

RESOURCES