Abstract
Conformational motions play an essential role in enzyme function, often facilitating the formation of enzyme-substrate complexes and/or product release. Although considerable debate remains regarding the role of molecular motions in the conversion of enzymatic substrates to products, numerous examples have found motions to be crucial for optimization of enzyme scaffolds, effective substrate binding, and product dissociation. Conformational fluctuations are often rate-limiting to enzyme catalysis, primarily through product release, with the chemical reaction occurring much more quickly. As a result, the direct involvement of motions at various stages along the enzyme reaction coordinate remains largely unknown and untested. In the following review, we describe the use of solution NMR techniques designed to probe various timescales of molecular motions and detail examples in which motions play a role in propagating catalytic effects from the active site and directly participate in essential aspects of enzyme function.
Keywords: NMR, Relaxation dispersion, CPMG, R1rho, Enzymes
1. NMR methods and theory
Molecular motions are critical in the function of enzymes. Enzymes change their conformation multiple times during the catalytic cycle and these motions must occur with timescales commensurate with the rate constants that define the reaction mechanism. Characterization of these motions is essential to understanding their role in enzyme chemistry. Solution NMR spectroscopy is the experimental method of choice for analyzing molecular motions over an enormous timescale range. Solution NMR can be performed under a wide range of aqueous conditions including those that approximate physiological pH and [salt]. Moreover the incorporation of spin-1/2 nuclei into proteins is minimally perturbative and is relatively straightforward. Further, NMR spectroscopy maintains protein integrity, and the experimental library for the study of protein structure and dynamics is constantly improving. Biomolecular NMR has historically been restricted to proteins smaller than 50 kDa, but novel TROSY [1] methods and 13C-methyl labeling strategies [2–5] now enable the study of much larger proteins through enhanced signal-to-noise (S/N) and resolution.
Here, we briefly review the NMR methods designed to characterize macromolecular motions followed by examining several examples from the authors’ work that illustrate some of the biological insight that can be obtained through the use of this powerful technique.
1.1. Picosecond-to-nanosecond motions (ps–ns)
NMR techniques are powerful for the study of atomic-resolution protein dynamics over an enormous time scale ranging from picoseconds (ps) to seconds (s) (Fig. 1) [6–9]. Motions on the fast end (ps–ns) of this continuum, which are faster than the overall rotational diffusion of the protein under study, reflect stochastic equilibrium fluctuations in the bond vectors of individual atoms. These stochastic motions modulate the chemical shift anisotropy and dipolar interactions between the nuclei. The identity of nuclei and frequency of the motions determine the rate at which Boltzmann equilibrium is established [10–14]. A comprehensive review of the theoretical aspects of NMR spin-relaxation is given by Palmer et al. [15–17] Here, we present a shortened overview of the spin-relaxation formalism. The mathematical expressions below are useful for describing a heteronuclear spin-1/2 pair system such as the amide proton-nitrogen (1H–15N) located in the protein backbone. The longitudinal, transverse magnetization and cross-relaxation of the 15N heteronucleus relax (R1, R2, σIS) or return to their Boltzmann equilibrium state as described by Abragam [18],
| (1) |
| (2) |
| (3) |
where ωI and ωS are the Larmor frequencies of the I (1H) and S (15N) nuclei and c is the chemical shift anisotropy coupling , in which Δσ is the chemical shift anisotropy value of the S nucleus. Rex is the additional contribution to R2 that results from conformational exchange motions that occur with µs–ms frequency and is often equal to zero. The dipolar coupling constant d is described by Eq. (4),
| (4) |
Here, μo is the permeability of free space, h is Planck’s constant, γI and γS are the gyromagnetic ratios of nuclei I and S, and 〈rIS〉 is the average internuclear bond length for the I and S atoms. The spectral density function, J(ω), is the cosine Fourier transform of the autocorrelation function of the I−S bond vector motion [18].
| (5) |
Fig. 1.
Timescale depicting enzyme motions. (Top) A cartoon representation of the types and variety of enzyme motions. (Bottom) A list of the types of solution NMR experiments and the timescale of motions to which they are sensitive.
When internal bond vector fluctuations are uncorrelated with the macromolecular rotational diffusion, the total autocorrelation function is obtained as the product of the autocorrelation functions of the overall (CO) and the internal (CI) motions [19].
| (6) |
For isotropic rotational diffusion, as with a spherical macromolecule, CO decays as a single exponential as shown in Eq. (7);
| (7) |
where τc is the rotational correlation time of the protein. The spectral density function describing this motion is expressed in Eq. (8).
| (8) |
Thus, for bond vector motions due to rotational tumbling of the protein and internal motions, the spectral density function becomes
| (9) |
with . In Eq. (9), τe is the effective correlation time for internal motions and S2 is a generalized order parameter describing the amplitude of internal bond vector motions. S2 values range from 0 to 1, where lower values indicate enhanced flexibility. Thus, measurement of R1, R2, and σIS allow sufficient sampling of the spectral density function to subsequently determine S2 on a per-residue basis [20]. With some caveats, the magnitude of S2 obtained by NMR relaxation experiments is also a measure of the configurational entropy of the individual bond vectors in the protein under consideration [21–23]. In addition to the 15N labeled protein backbone, Cα positions as well as sidechain dynamics on the ps–ns timescale can be characterized by NMR spectroscopy [24–30].
1.2. Microsecond-to-millisecond motions (µs–ms)
Several of the most critical functions of biomolecules occur over the µs–ms time regime (Fig. 1) and many of the useful NMR experiments designed to probe the dynamics of these processes are called relaxation dispersion techniques. Relaxation dispersion NMR has been used to characterize protein motions [31–35], protein folding [36–38], ligand binding [39,40], and enzymatic mechanisms [41–46]. The collection and interpretation of NMR relaxation data depends on conformational fluctuations due to dynamic processes, which are derived and discussed in detail in the following sections.
Molecular motions (µs–ms) involving multiple atoms, amino acids, or elements of protein secondary structure require traversing significant energy barriers. If such motions also involve changes to the magnetic environment of the spin-1/2 nuclei involved in the motional process then these motions can be studied due to the elevated R2 value at these flexible sites. Generically these conformational exchange motions occur between conformations A and B, where with forward (k1) and reverse (k−1) rates constants and fractional populations pA and pB, respectively. The sum of these rate constants is called kex. As noted above, fluctuations in the local magnetic field experienced by each NMR-active nuclei due to motional exchange broaden the resonances and result in an additional contribution (Rex) to the transverse relaxation rate constant, R2, where Eq. (1) can be recast as;
| (10) |
is determined by ps–ns motions at frequencies near the nuclear Larmor frequency, which include internal bond vector fluctuations and overall rotational diffusion of the protein.
Molecular motions exceeding ~5000 s−1, are best characterized by the off-resonance R1ρ experiment, which is capable of providing insight into motional processes up to ~105 s−1 [31]. In the off-resonance R1ρ experiment (Eq. (11)) [47] an effective field (ωe) is used to irradiate the protein where modulation of ωe affects the contribution of molecular motions to the measured relaxation rate [48].
| (11) |
Because the resonance frequencies differ for the nuclei involved in the conformational exchange process, the effective magnetic fields for sites A and B are expressed as and . The frequency offsets from the RF carrier of site A, site B, and population-averaged resonances are ωA, ωB, and ωiso, respectively. The RF field strength is ω1 with a tilt angle θ = arctan(ω1/ωiso). In the case of fast conformational exchange where kex > Δω, characterization of motions from the R1ρ experiment is simplified and uses Eq. (12), in which
| (12) |
where Rex is:
| (13) |
and Φex = papbΔω2/kex. Kinetic processes occurring with rate constants between a few hundred to 5000 s−1 are characterized by the relaxation-compensated Carr–Purcell–Meiboom–Gill (rcCPMG) experiment [32,49]. Relaxation dispersion of the CPMG type utilizes the fact that when 180° radio frequency (RF) refocusing pulses are applied to a system undergoing conformational exchange, the dephasing of the ensemble of spins due to visitation of different magnetic environments is disrupted by these RF pulses. The extent of this modulation depends on the relationship between the pulsing frequency (τcp−1) and kex. This modulation manifests as τcp-dependent differences in the exchange contribution to R2 [50–52]. These CPMG experiments can be performed on single (SQ) [32,49], double [53], zero [53], or multiple quantum (MQ) [54] coherences. Standard indirect-detection probes typically enable shorter 13C pulses than 15N pulses, thus the duty cycle for 13C CPMG experiments can often be greater than those for 15N-based experiments. Below we describe the SQ and MQ experiments as they are more frequently used.
The basis of the single-quantum rcCPMG experiment is given by Eq. (14) where the measured transverse relaxation rate (R2(1/τcp)) varies as the 180° pulse spacing (τcp) is altered [50–52],
| (14) |
in which and are the intrinsic transverse relaxation rates of the two sites in the absence of conformational exchange.
The expression described by Eq. (14) is useful when conformational motion is in slow to intermediate exchange (kex ≤ Δω). Eq. (14) simplifies under fast exchange conditions (kex > Δω). In this exchange regime, conformational exchange processes characterized by the rcCPMG experiment are described by Eq. (15). [55]
| (15) |
where ϕex = pApBΔω2. A full characterization of the exchange parameters for a particular motional process necessitates acquisition of experimental NMR data at two or more static magnetic fields and/or multiple temperatures [56–58].
rcCPMG NMR was first applied to proteins for the study of backbone amide groups [32], but there have been several adaptations to other spin-1/2 pairs [54,59]. Another variant of this experiment for characterization of sidechain methyl groups such as Ile, Leu, Val, and Ala residues is denoted as the multiple quantum (MQ) relaxation dispersion experiment [54,59]. If sufficiently high-quality data are obtained, the MQ dispersion experiment has the added benefit of providing ΔωH and ΔωC, which could yield additional structural information on the lowly populated conformation.
In total the CPMG and R1ρ experiments report on the kinetics of the conformational motion from 102 to 105 s−1, provide the equilibrium populations of the interconverting conformations, and elucidate structural information about ground and higher energy conformations through analysis of Δω values [60–62]. Incorporation of TROSY enhancement into any of these experiments enables studies of very large proteins or macromolecular assemblies (≥ 100 kDa) [1]. Below, applications of these NMR experiments for the study of enzyme function are reviewed. Only three examples are covered here, yet there are many excellent studies in the literature to which the interested reader is referred for additional information [46,63–95].
2. Ribonuclease A (RNase A)
RNase A is a 14 kDa monomeric enzyme that catalyzes the transphosphorylation of single stranded RNA with specificity for pyrimidines on the 3′ side of the bond to be cleaved and purines on the 5′ side of this bond. The slow step in the enzyme-catalyzed reaction has been demonstrated to be product release [96] while biophysical experiments have provided evidence of conformational changes in the enzyme that accompany ligand binding at the active site [97].
2.1. Fast dynamics
Changes in ps–ns motions contribute to the configurational entropy changes in proteins [21–23,98] and there are often compensating changes in S2 values upon ligand binding [99]. To address the role of these fast timescale motions in RNase A function, measurement of the temperature and ligand dependence of S2 was performed [100]. To achieve this, the backbone dynamics of RNase A were characterized in the apo state and bound to a non-cleavable, high-affinity substrate analog, 5′-phosphothymi dine(3′,5′) pyrophosphate adenosine 3′-phosphate (pTppAp). pTppAp binds to the RNase A active site and is a mimic of the pre-catalytic ES complex. With few exceptions, the S2 values in RNase A increase when bound to pTppAp (Fig. 1). Thus, RNase A pays an entropy penalty, in the form of reduced amplitude of backbone motion, upon binding the high-affinity substrate analog. The global effect of restricted motion upon binding pTppAp stands in contrast to the local perturbations of NMR chemical shifts, which primarily occur at the active site. Additional insight comes from the temperature dependence of S2 values [86]. The overall average S2 values for apo and liganded RNase A have the same temperature dependence (dS2/dT = 0.0014 K−1). This indicates no global contribution to the overall heat capacity change due to changes in dynamics. However, on an individual residue basis there are significant differences in the temperature dependence of S2 (Fig. 2). Thus, the differing d(1 − S)/dT values are offset such that dynamical changes to ΔCp = 0. Interestingly, the temperature dependence of S2 for the individual secondary structure units becomes more uniform when RNase A is bound to pTppAp compared to apo RNase A (Fig. 3). This would suggest that binding of pTppAp facilitates a more collective dynamic behavior across the protein with a more uniform shape of the protein energy landscape for RNase A in the bound state.
Fig. 2.
Temperature-dependent S2 values for RNase A for apo (open circles) and pTppAp bound (filled circles). Breaks in the lines indicate residues in which quantitation of dynamical parameters was not possible. Error bars have been omitted to aid in visualization of the data trends. Uncertainties in S2 values for the data shown, in all cases, is ~1–2%. Adapted with permission from [100]. Copyright 2003 American Chemical Society.
Fig. 3.
Temperature dependence of the order parameters for RNase A. The values of 1−S for glutamine 11, lysine 41, and threonine 99 are shown as a function of temperature for apo RNase A (open circles/solid lines) and RNase A/pTppAp (filled circles/dashed lines). The slopes of the fitted lines are Gln11: 7.5 ± 0.4 × 10−4 K−1, 2.9 ± 2.6 × 10−4 K−1; Lys41: 2.0 ± 0.7 × 10−4 K−1, 1.6 ± 0.2 × 10−3 K−1; and Thr99: 0.2 ± 0.1 × 10−3 K−1, 1.5 ± 0.1 × 10−3 K−1 for apo and RNase A/pTppAp, respectively. Reprinted with permission from Ref. [100]. Copyright 2003 American Chemical Society.
2.2. Slow timescale motions (µs–ms)
Early work using CPMG relaxation dispersion studies identified a two-site conformational exchange process at multiple regions throughout the enzyme occurring at a rate of exchange (kex) of 1700 s−1 [41]. These results were intriguing because this conformational exchange rate constant is identical to the RNase A kcat value and to the product release rate constant (koff) determined by NMR lineshape analysis [42]. Product release is known to be rate-limiting in RNase A [101] and thus this initial work suggested that the conformational change measured by NMR might be coupled to the rate-limiting step in enzyme turnover. To examine the role of ms motions in RNase A function, NMR relaxation dispersion experiments were performed on RNase A at various states along the enzymatic reaction coordinate including apo, substrate analog-bound (pTppAp), and product-bound (3′-cytidine monophosphate (3′-CMP)) complexes [43]. The structural changes in the backbone conformation as RNase A proceeds from the unliganded state to the ES complex are shown crystallographically to be localized to active site Loop 4 and Loop 1, which is located ~20 Å from the active site [43,102]. A similar set of structural changes occurs in the product state with the exception of Loops 4 and 1, which appear to be intermediate between the E and ES forms. The timescale of conformational motions in these complexes is largely unchanged compared to the apo state and remains around 1700 s−1 at 20 °C. Moreover, these ms motions are confined to similar locations in Loop 1 and Loop 4 in the apo, ES, and EP states [43]. Whereas the kinetics of motion remain constant, they reflect different processes in each complex. Further, the energetics obtained by temperature-dependent relaxation dispersion methods are similar in each complex, with an activation energy of ~5 kcal/mol. In total, these data suggest that motion in RNase A is a two-site process whereby ligand binding shifts the conformational equilibrium between the two species. Insight into the structure of the exchanging conformations can be obtained from the Δω values determined from fits to the relaxation dispersion data [103]. The Δω values or so-called dynamic chemical shift difference, when compared between different enzyme complexes, can inform on the relation between the interconverting structures and those structures induced by ligand binding. As shown in Fig. 4, Δω values for residues in and near Loop 1 in RNase A when bound to pTppAp are equivalent to the calculated Δω values for apo RNase A. Furthermore, the signs of these Δω as determined from a comparison of HMQC and HSQC experiments [104] are opposite. This suggests that Loop 1 in apo RNase A samples the bound conformation in the absence of pTppAp and when pTppAp is bound to RNase A, this loop samples a minor conformation that resembles the apo enzyme [105]. Loop 1 is approximately 20 Å from the active site but is linked to Thr45 via a series of hydrogen bonds. Thr45 is located at the active site and is one of the residues that plays a role in pyrimidine specificity [106,107]. This observation raised the question of whether motions at Loop 1 could be propagated to the active site to facilitate disruption of the Thr45-substrate interaction and thus enable product dissociation and if so, by what mechanism.
Fig. 4.
Distribution of d(1 − S)/dT across 2° structure. The variation in d(1 − S)/dT within each secondary structure element in RNase A is depicted in the box plots for (A) apo RNase A and (B) RNase A/pTppAp. Elements of 2° structure are arranged in order of occurrence in the RNase A primary structure. The bar inside the box indicates the median value of d(1 − S)/dT, while the ends of the box mark the boundaries with the upper and lower quartile of the data and represent the ±25% variation of the population, the bars extending from the box indicate the minimum and maximum data value. The circles represent outliers in the data set, calculated as points whose value is either: greater than upper quartile +1.5 * IQD or less than lower quartile −1.5 * IQD in which IQD (inter quartile distance) is defined as the upper quartile–lower quartile. The ratio of the number of residues from each 2° structure element used in the box plot relative to the total number of non-proline residues in that segment of 2° structure for apo RNase A(RNase A/pTppAp) is: α-1, 9(9)/9; loop-1, 8(6)/12; α-2, 7(8)/8; loop-2, 8(8)/9; β-1, 5(0)/5; loop-3, 2(2)/3; α-3, 9(9)/9; β-2, 4(4)/4; loop-4, 8(7)/8; β-3, 3(3)/3; loop-5, 4(4)/4; β-4, 7(5)/8; loop-6, 7(7)/9; β-5, 13(13)/15; loop-7, 3(3)/3; β-6, 7(6)/7. The d(1 − S)/dT values presented in this plot are for all residues in which S2 could be quantitated at all three temperatures. Reprinted with permission from Ref. [100]. Copyright 2003 American Chemical Society.
To address this question in more detail, experiments were performed to focus on the catalytic role of molecular motions that occur distant from the active site. Kinetic solvent isotope effects (KSIE) are commonly used to elucidate chemical reaction mechanisms. In the case of RNase A, KSIE were exploited to ascertain the role of hydrogen bonds in the conformational exchange process. It was demonstrated that the rate of conformational motion in RNase A decreased with increasing deuterium solvent content (Fig. 5) [44]. These data indicated that during the conformational motion of RNase A, one or more solvent exchangeable protons were involved in the transition state of this process. Further, the linearity of the isotope exchange studies indicated that there was a single dominant proton transfer step in the conversion between apo-like and bound-like conformations in RNase A [44]. Additional studies narrowed the location of the proton transfer site that is crucial to the motion in RNase A to the Loop 1 region and the surrounding β-sheets (Fig. 6) [108]. In this work it was noted that ms motions of residues in this region were sensitive to deuterium content in the solvent, whereas flexible residues in other parts of RNase A had conformational exchange rate constants that were unaffected by increasing D2O concentrations. This indicated that despite the identical kex values for the solvent sensitive and insensitive regions in RNase A, the motions in each region must be distinct processes and their similarity in H2O was merely coincidental. Moreover, because the solvent isotope effect () on RNA transphosphorylation is the same as the isotope effect on conformational dynamics () the motions in these regions were suggested to be coupled to the rate-limiting step in catalytic turnover [108]. Previous biochemical and biophysical studies suggested that histidine 48, located in the vicinity of Loop 1, was involved in a conformational change in RNase A [109,110]. H48 is located on β-strand 1 and its imidazole group points into the Loop 1 region. To investigate whether protonation of this residue was responsible for the isotope sensitive motion in RNase A, H48 was mutated to alanine [108]. The H48A mutant experiences a 10-fold decrease in catalytic activity despite the imidazole group being 18 Å from the catalytic site. Interestingly, the H48A variant experiences no deuterium solvent isotope effect on kcat, unlike the WT, which has a KSIE value of 2. This suggests that the origin of the isotope effect on enzyme function arises from the H48 residue. NMR relaxation dispersion studies on the H48A variant indicate the loss of ms motions in all previously flexible residues that also had sensitivity to D2O (gold spheres in Figs. 6 and 7). In contrast, residues in WT RNase A that maintained ms motions in H2O and D2O were unaffected by mutation of H48 (Fig. 7) further confirming the motional independence of these two regions of RNase A. These data further indicate that the site of the exchangeable proton that plays a prominent role in enzyme function and conformational motions is located near H48.
Fig. 5.
Chemical shift differences in RNase A complexes. Correlation of Δω values determined by global fitting with Eq. (14) for residues with quantifiable data in both the apo and pTppAp forms of RNase A. The red line represents a linear fit to the data points for Q11, S15, S16, T17, S22, and Q101. Data fro H119 and K66, shown in blue, are excluded from this fit. Reprinted with permission from Ref. [105]. Copyright 2005 American Chemical Society.
Fig. 6.
Relaxation dispersion data for the Cε position of Met29 at 18.8 T for 4.8% (orange), 52% (blue), and 98% (black) D2O. The fitted lines to the data points are the result of a global two-field fit to all residues using Eq. (15). Reprinted with permission from [44]. Copyright 2006 American Chemical Society.
Fig. 7.
Location of flexible residues. Residues involved in chemical exchange are shown as spheres with the amino acid residue number indicated. Gold spheres indicate residues in which a normal 2H solvent isotope effect of 2 is observed. Spheres colored green are atoms that are flexible and do not exhibit a solvent isotope effect. The four histidine residues are depicted as red sticks. Selected secondary structure elements are indicated. Reprinted from Ref. [108] with permission. Copyright 2007 National Academy of Sciences, USA.
Additional studies were undertaken to attempt to address the details of the interactions between H48 and Loop 1 and their role in RNase A function. Primary sequence analysis and X-ray crystallography studies of RNase homologs indicated that they possessed a shorter Loop 1 and substantially lower rates of catalysis. Doucet et al. swapped the shorter loop from the RNase A homolog eosinophil cationic protein (ECP) in place of that loop in RNase A, creating RNaseAECP (Fig. 8) [111]. This chimeric enzyme had 10× lower catalytic activity as well as a 10-fold decrease in the product release rate constant, koff [111]. Importantly, as shown by NMR CPMG relaxation dispersion studies, flexible residues that are involved in product release showed a total loss of ms motions in the chimera. These residues are also those that showed sensitivity to solvent deuterium content. These studies further support the role of the flexible Loop 1 in RNase A activity, as the shorter ECP loop has a limited range of motion, inhibiting product release in RNase A as well as abolishing all motions in the crucial active site β-strands (see Figs. 9 and 10).
Fig. 8.
Solvent isotope and mutation effects. 15N and 13Cε rcCPMG relaxation for 5% (black), 10% (aqua), 52% (blue), and 100% (green) 2H2O. WT data are indicated by circles, and triangles are used to indicate H48A data. NMR CPMG dispersion data were collected at 18.8T (open symbols) and 14.1T (closed symbols) at 298 K. Reprinted from Ref. [108] with permission. Copyright 2007 National Academy of Sciences, USA.
Fig. 9.
Structural comparison of bovine ribonuclease A (RNase A) and human Eosinophil Cationic Protein (ECP). (A) Superposition of RNase A (blue structure, PDB code 1FS3 [130]) and ECP (red structure, PDB code 1DYT [131]). H12 highlights the position of the active site relative to the position of His48 and loop 1 in RNase A. (B) Zoomed view of a few atomic interactions between His48 and loop 1 in RNase A. Note the longer loop 1 in RNase A (D14SSTSAASSSNY25) relative to the shorter loop 1 of ECP (S17LNPPR22, ECP numbering). (C) Structural alignment of RNase A, ECP and EDN. Active site positions Q11, H12, K41, H119 and D121 are marked with stars while cysteine residues forming the strictly conserved four disulfide bridges are represented by arrows. Loop 1 and position 48 are boxed. Alignment was performed with TCoffee Expresso [132] using PDB coordinates 7RSA [133], 1DYT [131] and 1GQV [134]. Reprinted from [111] with permission. Copyright 2009 American Chemical Society.
Fig. 10.
Conformational dynamics in RNase A. (A) Residues in RNase AECP undergoing conformational exchange with a ΔR2 (1/τcp) greater than 1.5 s−1 are plotted on the structure of RNase A (PDB code 7RSA [133]). Red spheres = similar rate constants (kex) in WT and RNase AECP; Blue spheres = Decreased kex values in RNase AECP relative to WT. H12 and H48 are shown as yellow sticks and loop 1 is shown in cyan. (B) CPMG dispersion curves for WT RNase A (black circles), RNase AECP (red squares) and mutant H48A (blue diamonds). 15N relaxation dispersion curves are shown for two residues displaying similar conformational exchange in the three enzymes (Gln69 and Asn71), and three residues displaying significantly affected kex in both RNase AECP and mutant H48A (Thr82, Asp83 and Gln101). Y115 is represented as a motionless control. Fitted lines to data points are from single-field fits at 14.1 T using Eq. (15). Reprinted from [111] with permission. Copyright 2009 American Chemical Society.
In an effort to more accurately define the interactions involved in propagating motion at Loop 1 and H48 to the active site, studies were directed to two threonine residues (T17 and T82) that form hydrogen bonds with H48 [112]. T17 is located in Loop 1 and T82 is located on β-sheet 4, which comprises part of the active site RNA recognition region (Fig. 11). The results of the T17A and T82A mutations highlight the exquisite and finely tuned nature of the interactions and their impact on propagation of these ms motions. The T82A mutation results in a complete loss of ms motions in a manner similar to that observed in RNaseAECP as well as a 10-fold decrease in affinity for the product analog (3′-CMP) of the RNase A reaction (Fig. 12). Additional analysis indicated that the motions in the T82A variant are shifted to a faster timescale (>5000 s−1) but not abolished. The effects of the T17A mutation are subtler than those seen with T82A. Residues, A5, Q11, S15, S18, S22, M30, S32, N34, L35, F46, V47 retain ms motions, albeit at an elevated rate constant of 3200 s−1 compared to 2000 s−1 for WT (Fig. 12). Intriguingly, koff for 3′-CMP in T17A is also slightly elevated to 3000 s−1 compared to its WT value of 2000 s−1. These results suggested that T17 is involved in dampening, for an unknown reason, ms motions that are directly involved in product release. In contrast, removal of the sidechain of T82 results in elevated ms motions with minimal effect on koff, although T82A reduces kon by 10-fold as observed by NMR lineshape analysis. These studies illuminate how residue-specific participation in protein motions can be deciphered and linked to specific aspects of the catalytic cycle.
Fig. 11.
Hydrogen bonding interactions between His48 and its surrounding environment in RNase A. Hydrogen bonds are shown as dashed lines and atoms are displayed according to a standard coloring scheme: green for carbon, red for oxygen, and blue for nitrogen. Reprinted from [112] with permission. Copyright 2011 American Chemical Society.
Fig. 12.
Conserved short-range conformational dynamics in mutant T17A. 14.1 T 15N-CPMG dispersion curves of residues (A) F46, (B) V47, and (C) T82 for WT RNase A (black circles), RNase A-T17A (red squares), RNase A-T82A (orange triangles), RNase A-H48A (green diamonds) [108,111], and RNase AECP (blue inverted triangles) [111]. Insets show zoomed views of the RNase A-T17A data (same legend and axis labels). Reprinted from [112] with permission. Copyright 2011 American Chemical Society.
Specific focus on H48 provided further insight into its role in functional motions of RNase A. The central involvement of H48 was underscored by its mutation to alanine that resulted in complete loss of ms motions in critical regions of the enzyme as well as a significant loss in catalytic activity [111]. To further analyze the role of H48, an unnatural amino acid, 4-thiamethyl imidazole (RNase A-4MI) was incorporated into uniformly 15N-labeled RNase A [113]. This enzyme was studied by kinetic assay, solution NMR, and computational methods. The additional methylene group and sulfur atom in the 4MI increases the sidechain volume by 35 Å3 and slightly alters the imidazole pKa. The 4MI enzyme retained WT catalytic activity and had ms motions similar to those of the WT enzyme; however, the pH dependencies of these motions were not identical to those of WT. This suggested that the altered pKa of the unnatural imidazole was partly responsible. MD simulations indicate that H-bonds at crucial sites are maintained, but were not identical to those of the WT enzyme. Moreover, the pattern of correlated motions observed during the simulation is different for the two enzymes. These data suggest that, in part, protein dynamics could be uncoupled from activity, which has interesting implications for the enzyme design field.
Overall, this work demonstrates the involvement of numerous residues in conformational motions of RNase A and highlights the complex physical chemistry at play in modulating these motions. RNase A possesses two flexible regions that move independently (called cluster 1 and cluster 2) yet at room temperature possess the same kex values (Fig. 13). This complexity was recently studied in a series of temperature and pH dependent CPMG relaxation dispersion experiments [114]. Many residues within RNaseA showed evidence of ms motions at all temperatures (10, 15, 20, 25 °C) and pH values >6.4 (Fig. 14). In several cases, these motions became too fast for the CPMG experiment at pH values ≤5.5. Residues in dynamic cluster 1 were found to be more sensitive to changes in pH than those of cluster 2, and gradual increases in kex were observed with increasing temperature (T = 10 → 25 °C, kex = 980 → 1970 s−1 at pH 7.0) and decreasing pH (pH = 7.0 → 4.5, kex = 1970 → >3000 s−1). At a pH of 4.5, no CPMG dispersion was observed in cluster 1 at any of the temperatures tested. However, 15N R1ρ relaxation dispersion experiments revealed the presence of conformational motions (Fig. 14), and global fits to these data yielded kex = 13,400 s−1 for cluster 1. Interestingly, kex values for cluster 2 residues were insensitive to pH values above 4.5, and low pH kex values for cluster 2 residues were also determined by R1ρ methods. However, temperature-dependent variations in kex on the CPMG timescale were found to be greater in cluster 2 when compared with cluster 1.
Fig. 13.
Crystal structure of apo WT RNase A (PDBID: 7RSA). Active site residues H12, K41, and H119, as well as dynamically important residues H48 and D121 are depicted as sticks and are labeled, along with the elements of secondary structure. Dynamic clusters 1 and 2 are enclosed within the blue boxes. Reprinted from [114] with permission. Copyright 2015 American Chemical Society.
Fig. 14.
Representative global fits of relaxation dispersion data for clusters 1 and 2 of RNase A. Data collected at 500 and 600 MHz are represented with ▼ and ▲, respectively. Global 2-field CPMG fits of dynamic cluster 1 at 10 °C, pH 7.0 and dynamic cluster 2 at 25 °C, pH 5.5 and 20 °C, pH 6.4 are shown, as well as lack of CPMG-detected dispersion for Cluster 1 at 600 MHz at 25 °C, pH 4.5. 600 MHz R1ρ data at 25 °C and pH 4.5 of both clusters are shown in the bottom panels. The backbone nitrogen atoms of the residues involved in each global fit are mapped onto the structure in each panel and are color matched to the color of the respective dispersion data. Dynamic cluster 1 residues S15, S16, T17, F46, V47, H48, T82, D83, T100 and Q101 are shown in gray, black, blue, purple, green, dark teal, teal, orange, brown, and dark red, respectively. Dynamic cluster 2 residues A64, C65, K66, T70, and N71 at 25 °C and pH 5.5 are shown in fuchsia, yellow, red, cyan, and green, respectively. Reprinted from [114] with permission. Copyright 2015 American Chemical Society.
To develop a stronger understanding of the thermodynamic principles underlying these temperature and pH-dependent motional changes, the enthalpic and entropic contributions to the energy landscape of RNaseA were analyzed for cluster 1 and cluster 2 under various conditions (Fig. 15). Lowering the pH from 7.0 to 6.4 increases the motional activation energies of cluster 1 (Ea,fwd and Ea,rev) by 3.9 and 1.6 kcal/mol, respectively. Similar changes in were observed; however, a net stabilization of ΔG‡ for the motion of cluster 1 resulted from a favorable overcompensation of the activation entropy. ΔG‡ was found to be >10 kcal/mol for both forward and reverse motions of cluster 1, which is only slightly larger than reported energies of product dissociation () [115], consistent with cluster 1 being linked to product release.
Fig. 15.
Energy landscape surfaces for conformational changes (A ⇌ TS† ⇌ B) in RNase A. pH-dependent changes in enthalpy (left) and entropy (right) are depicted for Cluster 1 (top), 2 (middle), and R33 (bottom). Data were obtained from temperature dependent NMR relaxation experiments. In all cases conformation A is set at a reference value of 0 kcal/mol. Reprinted from [114] with permission. Copyright 2015 American Chemical Society.
Cluster 2, which is pH insensitive over the range of 5.5–7.0, showed little change in its activation parameters over this range. However, at pH 4.5, when cluster 2 dynamics shift to a faster timescale, the transition state free energy is stabilized by ~1 kcal/mol for both components of the motion (). Near neutral pH, entropy was the dominating contributor to the free energy difference (ΔΔG) between the major and minor conformational states of cluster 1. However, as the pH is lowered, the major conformation became more enthalpically favored, as the minor state was destabilized by ~2.3 kcal/mol (ΔΔH). In contrast, the enthalpic and entropic terms governing the conformational states of cluster 2 were essentially identical over the pH range examined, with small changes in these parameters arising at low pH.
To build upon these results, values of Δω, the 15N chemical shift difference between the exchanging conformational states were plotted as a function of pH for clusters 1 and 2 (Fig. 16). All residues in dynamic cluster 1 showed pH dependence to their Δω values, while cluster 2 residues showed no pH dependence. Although there was good correlation of Δω for cluster 1 at various pH values, these data were best approximated with a line of slope 1.3, while Δω for cluster 2 fits to a line of slope 1 at all pH values (Fig. 16B). Both clusters showed similar distributions of Δω, where cluster 1 ranged from 0.99 ≤ Δω ≤ 3.25 ppm and cluster 2 from 0.86 ≤ Δω ≤ 3.08 ppm. While several residues within each cluster showed comparatively high (or low) values of Δω, correlations between a residue’s location and the magnitude of Δω have yet to be identified. These data for cluster 1 suggest that this region interchanges between a highly populated conformation and a pH-dependent averaged conformation that exists in rapid equilibrium between a protonated and deprotonated conformer.
Fig. 16.
The dependence of Δω on pH for clusters 1 and 2 of RNase A. (Top) Correlation plot of Δω at pH 7.0 vs. Δω pH 6.4 of dynamic cluster 1 residues, S16, T17, F46, T82, D83, and Q101. The dashed line of unit slope shows a weak correlation (R2 = 0.46) with the data at the two values of pH, and is depicted for reference. The data is best described by a line of slope 1.29 ± 0.16 (weighted R2 = 0.97), with near-zero y-intercept, suggesting that the global motion may not be adequately described by a simple two-state model. (Bottom) Correlation plot of Δω at pH 5.5 vs. Δω at pH 7.0 (green), 6.4 (red) and 4.5 (blue) of loop 4 residues A64, K66, T70, and N71. Δω values determined at each of four pH values are well-described by a line of unit slope passing through the origin (R2 > 0.98), suggesting that the two chemical environments being sampled by cluster 2 remain unchanged over the pH range. Reprinted from [114] with permission. Copyright 2015 American Chemical Society.
The role of NMR in elucidating dynamic contributions to RNaseA function has been described in detail. CPMG (R1ρ) dispersion experiments have been utilized to examine slow (fast) timescale dynamics of residues near the active site and at distant sites, ultimately identifying two independently functioning dynamic clusters. Further, these NMR results have provided additional insight into the role of H48 dynamics in RNaseA function, and established free energy landscapes describing the temperature-and pH-dependence of RNaseA motions.
The extensive and detailed nature of these studies demonstrate how intricate protein motions can be and furthermore that their relation to enzyme function is not straightforward. NMR was integral in deciphering the role that these motions play in the catalytic cycle. Nonetheless, other biochemical and biophysical experiments were necessary for a full understanding. We anticipate that this multi-faceted approach will be crucial for studies of other enzymes as well.
3. Protein tyrosine phosphatases (PTP1B and YopH)
PTP1B and YopH are structurally similar protein tyrosine phosphatases. PTP1B is responsible for regulating insulin and leptin signaling and has been a target of intense drug discovery efforts [116]. YopH is an essential enzyme for Yersinia virulence and plays a role in disrupting the human phagocytic process [117]. Despite their different origins, PTP1B and YopH are structurally and mechanistically indistinguishable (Fig. 17). PTP1B and YopH each contain essential ten residue active site loops (WPD loop) and conserved eight residue P-loops, which bind a substrate phosphate group and contain nucleophilic residues C215/C403 (PTP1B/YopH). The active site WPD loop contains the acidic residue D181/D356 (PTP1B/YopH) that promotes protonation of the tyrosine leaving group during catalysis, and a 10 Å motion in the position of the WPD loop accompanies the transition from apo (open) to active (closed) enzyme (Fig. 17B and C) Both enzymes cleave the phosphate from their tyrosine phosphate (pY) containing substrates in two steps. First, the WPD loop closes on the pY substrate and catalyzes cleavage (kcleavage), generating a phosphoenzyme intermediate and a tyrosine peptide. Second, the phosphoenzyme is hydrolyzed (khydrolysis), releasing inorganic phosphate (Pi) and regenerating the active enzyme (Fig. 17D) [116].
Fig. 17.
Structural and mechanistic comparison of PTP1B and YopH. (A) Structural overlay of PTP1B (gray, PDB 3I80) and YopH (cyan, PDB 2I42) with a phosphate analog vanadate shown at the active site in red spheres. (B) Open (gray) and closed (blue) WPD loops of PTP1B. Vanadate (spheres) binds in the P-loop (orange) with C215 and R221 shown as sticks. Movement of D181 is indicated by the dashed lines. (C) Superposition of closed loops of PTP1B (gray) and YopH (cyan) with the three catalytic residues shown as sticks. (D) PTP catalytic reaction illustrating cleavage and hydrolysis. Reprinted with permission from Ref. [124]. Copyright 2013 American Association for the Advancement of Science.
PTP1B and YopH have essentially identical mechanisms, and KIE experiments support identical transition states as well [118–120]. However, kcat values for the two enzymes are significantly different, as YopH is over 20-fold more active than PTP1B (i.e. 700–1000 s−1 vs 15–30 s−1 at pH 6.6, respectively) [119,121,122]. More specifically, kcleavage, which is directly influenced by the rate of active site loop closure, has been shown to vary from 1400–2000 s−1 in YopH to 25–80 s−1 in PTP1B [123]. The effects of WPD loop motion on the rates of catalysis in PTP1B and YopH have been investigated using solution NMR spectroscopy and have shown that loop closure dynamics plays a large role in the kinetic differences between the enzymes [124].
3.1. Characterization of the apo enzymes
In a series of 15N CPMG relaxation dispersion experiments, Whittier et al. have shown that apo PTP1B and YopH behave quite differently [124]. Of the assigned residues within the WPD loop of PTP1B, W179, F182, and A189 showed dispersion (Fig. 18B and C) with individual kex values each approximating 900 s−1, suggesting concerted loop motion. A subsequent global fit of these data revealed equilibrium exchange populations of 97.5 ± 1.4% (pa) and 2.5% (pb). Using kex and pa, the rates of loop opening (kopen) and closure (kclose) for apo PTP1B were determined to be 890 ± 190 s−1 and 22 ± 5 s−1, respectively. Based on these data and numerous X-ray crystallographic studies on apo PTP1B, the major conformer was suggested to contain an open WPD loop, while the minor state (pb = 2.5%) contains a closed WPD loop. Interestingly, a single X-ray crystal structure of apo PTP1B is shown with a closed WPD loop [125], suggesting that a closed conformation is accessible in the apo state. This interpretation is consistent with measured differences in the chemical shifts (Δδ) of the open and closed states, which are in good agreement with shifts calculated from X-ray crystal structures. Thus, the WPD loop was found to convert between open and closed states at a rate of 22 ± 5 s−1, directly in the regime of kcleavage (25–80 s−1), and suggested loop closure may occur simultaneously with leaving group protonation during the first step of catalysis.
Fig. 18.
Loop motions in PTP1B and YopH detected by NMR. (A) Open and closed PTP1B and YopH WPD loops. (B and C) 15N CPMG dispersion curves for W179, F182, and A189. (D) TROSY-detected off-resonance 1H-R1ρ dispersion curves for A359 (blue) and S361 (red). Error bars indicate standard errors determined from duplicate measurements. Reprinted with permission from Ref. [124]. Copyright 2013 American Association for the Advancement of Science.
Similar 15N CPMG studies of apo YopH showed flat dispersion plots for each loop residue (T358, A359, V360, S361), suggesting that any loop motions occurred outside the µs–ms time regime. Faster motions were then probed with TROSY-enhanced 1H-R1ρ experiments and dispersion was observed for A359 and S361 (Fig. 18D), yielding a global kex = 43,000 ± 6200 s−1. Additional information on the rate of YopH WPD loop motion was obtained from a TROSY Hahn-Echo experiment, where the closure rate (kclose) was found to be 1240 ± 280 s−1. Like PTP1B, kclose for YopH occurs in the same regime as kcleavage (1400–2000 s−1), consistent with direct involvement of loop closure in pY cleavage. These data were consistent with temperature-jump fluorescence experiments that reported similar rates of loop interconversion. However, data from fluorescence anisotropy and molecular dynamics investigations reported faster loop motions [126–128].
3.2. Peptide-bound (substrate mimicked) complexes
Complexation of PTP1B with the representative substrate peptide Ac-DADEXLIP-NH2 [116] (where X = difluoromethylpho sphono-phenylalanine, a non-cleavable substrate mimic) resulted in significant broadening of several WPD loop residues, namely W179 and F182, indicative of a new motional process. Further, CPMG dispersion curves at 600 and 800 MHz for A189 were different from those of the apo enzyme and showed oscillation at lower pulse frequencies stemming from slow conformational exchange (Fig. 19). However, an identical value of Δω in the apo and peptide-bound states (Δω = 3.2 ± 0.5 ppm) of PTP1B suggested that the WPD loop still sampled the open and closed conformations. A single X-ray crystal structure of substrate-bound PTP1B now shows an open WPD loop, again suggesting that both conformations are accessible [129]. From these data, the rate of loop closure was determined to be 30 ± 4 s−1, in line with that of the apo enzyme. Thus, peptide binding had minimal effect on kclose, while loop opening was slowed to 5 ± 1 s−1, further demonstrating the linkage between loop closure and substrate cleavage rates.
Fig. 19.
(A) Overlaid 1H–15N-TROSY HSQC spectra of apo PTP1B (red) in the presence of saturating DADEXILP peptide (X = difluoromethylphosphono-phenylalanine, blue) expanded to show the shift of WPD loop residue A189, indicated by the arrow. (B and C) Shifts of YopH WPD loop residues A359 and S361 compared in the apo enzyme (red) and in the presence of saturating DADEXILP peptide (blue). (D) 15N CPMG dispersion curves for peptide-bound PTP1B (Top) and YopH (Bottom) showing global fits at 600 (red) and 800 (blue) MHz for A189 (PTP1B) and A359 and S361 (YopH). Reprinted with permission from Ref. [124]. Copyright 2013 American Association for the Advancement of Science.
Loop motions in the peptide-bound YopH complex were slowed considerably, falling back into the CPMG exchange regime. Global fits of CPMG dispersion data for loop residues A359 and S361 gave Δω values of 3.5 and 4.5 ppm, respectively, and were similar to those determined for the apo enzyme. An overall kex for the YopH WPD loop was determined to be 1790 ± 240 s−1 with a population of closed conformer above 99%. The resulting kclose value was found to be 1770 ± 240 s−1, similar to the value of kcleavage, again suggesting a close connection between loop motion and the first catalytic step. Interestingly, even with differing substrate cleavage rates, the product (EP complex) mimic tungstate was found to have identical dissociation rates from both PTP1B and YopH and suggests that the rate of product release is similar in both enzymes.
Studies of these related protein tyrosine phosphatases, particularly those of Whittier et al., have shown that loop motions in two members of the tyrosine phosphatase enzyme family are closely correlated with a key step in the catalytic cycle, namely substrate cleavage and protonation of the leaving group. These data have also shown that ligand binding at the active site modulates loop kinetics, especially the rate of opening, and highlighted a dynamic energy profile describing tyrosine phosphatase function.
4. Concluding remarks
Solution NMR spectroscopy remains at the forefront of techniques designed to characterize structural and dynamic properties of biomolecules. Advances in the experimental library of NMR methods have only enhanced the possibility for atomic level information regarding processes such as ligand binding, catalysis, and allostery. The case studies outlined in this review provide examples of the power of NMR in monitoring molecular motions that are directly involved in an enzyme’s catalytic cycle and the propagation of those motions throughout the protein architecture. Coupled with other biophysical and computational methods, solution NMR promises to play a pivotal role in establishing new insight into biomolecular processes for years to come.
Acknowledgements
JPL acknowledges support from NSF grant MCB1121372 and NIH grant GM112781.
Glossary of abbreviations
- NMR
nuclear magnetic resonance
- TROSY
transverse relaxation-optimized spectroscopy
- CPMG
Carr–Purcell–Meiboom–Gill pulse sequence
- RF
radio-frequency
- SQ
single quantum
- MQ
multiple quantum
- RNase A
ribonuclease A
- pTppAp
5′-phosphothymidine(3′,5′) pyrophosphate adenosine 3′-phosphate
- ES
enzyme-substrate complex
- EP
enzyme-product complex
- HMQC
heteronuclear multiple quantum coherence
- HSQC
heteronuclear single quantum coherence
- KSIE
kinetic solvent-isotope effect
- ECP
eosinophil cationic protein
- 3′cMP
3′-cytidine monophosphate
- 4MI
4-thiamethyl imidazole
- PTP1B
protein tyrosine phosphatase 1B
- YopH
protein tyrosine phosphatase YopH from Yersinia
References
- 1.Pervushin K, Riek R, Wider G, Wuthrich K. Attenuated T2 relaxation by mutual cancellation of dipole-dipole coupling and chemical shift anisotropy indicates an avenue to NMR structures of very large biological macromolecules in solution. Proc. Natl. Acad. Sci. USA. 1997;94:12366–12371. doi: 10.1073/pnas.94.23.12366. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Tugarinov V, Kanelis V, Kay LE. Isotope labeling strategies for the study of high-molecular-weight proteins by solution NMR spectroscopy. Nat. Protoc. 2006;1:749–754. doi: 10.1038/nprot.2006.101. [DOI] [PubMed] [Google Scholar]
- 3.Korzhnev DM, Kloiber K, Kanelis V, Tugarinov V, Kay LE. Probing slow dynamics in high molecular weight proteins by methyl-TROSY NMR spectroscopy: application to a 723-residue enzyme. J. Am. Chem. Soc. 2004;126:3964–3973. doi: 10.1021/ja039587i. [DOI] [PubMed] [Google Scholar]
- 4.Shi L, Kay LE. Tracing an allosteric pathway regulating the activity of the HsIV protease. Proc. Natl. Acad. Sci. USA. 2014;111:2140–2145. doi: 10.1073/pnas.1318476111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Grishaev A, Wu J, Trewhella J, Bax A. Refinement of multidomain protein structures by combination of solution small-angle X-ray scattering and NMR data. J. Am. Chem. Soc. 2005;127:16621–16628. doi: 10.1021/ja054342m. [DOI] [PubMed] [Google Scholar]
- 6.Kempf JG, Loria JP. Theory and applications of protein dynamics from solution NMR. Cell Biochem. Biophys. 2002;37:187–211. doi: 10.1385/cbb:37:3:187. [DOI] [PubMed] [Google Scholar]
- 7.Palmer AG, III, Kroenke CD, Loria JP. Nuclear magnetic resonance for quantifying microsecond-to-millisecond motions in biological macromolecules. Methods Enzymol. 2001;339:204–238. doi: 10.1016/s0076-6879(01)39315-1. [DOI] [PubMed] [Google Scholar]
- 8.Kempf JG, Loria JP. Measurement of intermediate exchange phenomena. In: Downing AK, editor. Protein NMR Techniques. Totowa: Humana Press; 2004. pp. 185–231. [DOI] [PubMed] [Google Scholar]
- 9.Lipchock J, Loria JP. Millisecond dynamics in the allosteric enzyme imidazole glycerol phosphate synthase (IGPS) from Thermotoga maritima. J Biomol. NMR. 2009;45:73–84. doi: 10.1007/s10858-009-9337-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.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 magnetic relaxation of proteins. J. Am. Chem. Soc. 1990;112:4989–4991. [Google Scholar]
- 11.Peng JW, Wagner G. Mapping spectral density functions using heteronuclear NMR relaxation measurements. J. Magn. Reson. 1992;98:308–332. [Google Scholar]
- 12.Ishima R, Yamasaki K, Saito M, Nagayama K. Spectral densities of nitrogen nuclei in Escherichia coli ribonuclease HI obtained by 15N NMR relaxation and molecular dynamics. JBNMR. 1995;6:217–220. doi: 10.1007/BF00211786. [DOI] [PubMed] [Google Scholar]
- 13.Mandel A, Akke M, Palmer AG., III 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]
- 14.Farrow NA, Zhang O, Szabo A, Torchia DA, Kay LE. Spectral density function mapping using 15N relaxation data exclusively. J. Biomol. NMR. 1995;6:153–162. doi: 10.1007/BF00211779. [DOI] [PubMed] [Google Scholar]
- 15.Palmer AG., 3rd NMR characterization of the dynamics of biomacromolecules. Chem. Rev. 2004;104:3623–3640. doi: 10.1021/cr030413t. [DOI] [PubMed] [Google Scholar]
- 16.Palmer AG, 3rd, Massi F. Characterization of the dynamics of biomacromolecules using rotating-frame spin relaxation NMR spectroscopy. Chem. Rev. 2006;106:1700–1719. doi: 10.1021/cr0404287. [DOI] [PubMed] [Google Scholar]
- 17.Cavanagh J, Fairbrother WJ, Palmer AG, Rance M, Skelton NJ. Protein NMR Spectroscopy: Principles and Practice. second. San Diego: Elsevier Academic Press; 2007. [Google Scholar]
- 18.Abragam A. Principles of Nuclear Magnetism. Oxford: Clarendon Press; 1961. [Google Scholar]
- 19.Wallach D. Effect of internal rotation on angular correlation functions. J. Chem. Phys. 1967;47:5258–5268. [Google Scholar]
- 20.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]
- 21.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]
- 22.Yang D, 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]
- 23.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]
- 24.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]
- 25.Tugarinov V, Sprangers R, Kay LE. Probing side-chain dynamics in the proteasome by relaxation violated coherence transfer NMR spectroscopy. J. Am. Chem. Soc. 2007;129:1743–1750. doi: 10.1021/ja067827z. [DOI] [PubMed] [Google Scholar]
- 26.Tugarinov V, Kay LE. Separating degenerate (1)H transitions in methyl group probes for single-quantum (1)H-CPMG relaxation dispersion NMR spectroscopy. J. Am. Chem. Soc. 2007;129:9514–9521. doi: 10.1021/ja0726456. [DOI] [PubMed] [Google Scholar]
- 27.Tugarinov V, Kay LE. Relaxation rates of degenerate 1H transitions in methyl groups of proteins as reporters of side-chain dynamics. J. Am. Chem. Soc. 2006;128:7299–7308. doi: 10.1021/ja060817d. [DOI] [PubMed] [Google Scholar]
- 28.Liu W, Flynn PF, Fuentes EJ, Kranz JK, McCormick M, Wand AJ. Main chain and side chain dynamics of oxidized flavodoxin from Cyanobacterium anabaena. Biochemistry. 2001;40:14744–14753. doi: 10.1021/bi011073d. [DOI] [PubMed] [Google Scholar]
- 29.Wand AJ, Urbauer JL, McEvoy RP, Bieber RJ. Internal dynamics of human ubiquitin revealed by 13C-relaxation studies of randomly fractionally labeled protein. Biochemistry. 1996;35:6116–6125. doi: 10.1021/bi9530144. [DOI] [PubMed] [Google Scholar]
- 30.LeMaster DM, Kushlan DM. Dynamical mapping of E. coli thioredoxin via 13C NMR relaxation analysis. J. Am. Chem. Soc. 1996;118:9263–9272. [Google Scholar]
- 31.Akke M, Palmer AG. Monitoring macromolecular motions on microsecond-millisecond time scales by R1r–R1 constant-relaxation-time NMR spectroscopy. J. Am. Chem. Soc. 1996;118:911–912. [Google Scholar]
- 32.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]
- 33.Mulder FA, Hon B, Mittermaier A, Dahlquist FW, Kay LE. Slow internal dynamics in proteins: application of NMR relaxation dispersion spectroscopy to methyl groups in a cavity mutant of T4 lysozyme. J. Am. Chem. Soc. 2002;124:1443–1451. doi: 10.1021/ja0119806. [DOI] [PubMed] [Google Scholar]
- 34.Grey MJ, Wang C, Palmer AG., III 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]
- 35.Massi F, Grey MJ, Palmer AG., III Microsecond timescale backbone conformational dynamics in ubiquitin studied by NMR R1rho relaxation experiments. Protein Sci. 2005;14:735–742. doi: 10.1110/ps.041139505. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Hill RB, Bracken C, DeGrado WF, Palmer AG. Molecular motions and protein folding: characterization of the backbone dynamics and folding equilibrium of alpha D-2 using C-13 NMR spin relaxation. J. Am. Chem. Soc. 2000;122:11610–11619. [Google Scholar]
- 37.Korzhnev DM, Salvatella X, Vendruscolo M, Di Nardo AA, Davidson AR, Dobson CM, Kay LE. Low-populated folding intermediates of Fyn SH3 characterized by relaxation dispersion NMR. Nature. 2004;430:586–590. doi: 10.1038/nature02655. [DOI] [PubMed] [Google Scholar]
- 38.Tang Y, Grey MJ, McKnight J, Palmer AG, 3rd, Raleigh DP. Multistate folding of the villin headpiece domain. J. Mol. Biol. 2006;355:1066–1077. doi: 10.1016/j.jmb.2005.10.066. [DOI] [PubMed] [Google Scholar]
- 39.Mittag T, Schaffhausen B, Gunther UL. Direct observation of protein-ligand interaction kinetics. Biochemistry. 2003;42:11128–11136. doi: 10.1021/bi0347499. [DOI] [PubMed] [Google Scholar]
- 40.Tolkatchev D, Xu P, Ni F. Probing the kinetic landscape of transient peptide-protein interactions by use of peptide (15)n NMR relaxation dispersion spectroscopy: binding of an antithrombin peptide to human prothrombin. J. Am. Chem. Soc. 2003;125:12432–12442. doi: 10.1021/ja021238l. [DOI] [PubMed] [Google Scholar]
- 41.Cole R, Loria JP. Evidence for flexibility in the function of ribonuclease A. Biochemistry. 2002;41:6072–6081. doi: 10.1021/bi025655m. [DOI] [PubMed] [Google Scholar]
- 42.Beach H, Cole R, Gill M, Loria JP. Conservation of µs–msenzymemotions in the apo- and substrate-mimicked state. J. Am. Chem. Soc. 2005;127:9167–9176. doi: 10.1021/ja0514949. [DOI] [PubMed] [Google Scholar]
- 43.Kovrigin EL, Loria JP. Enzyme dynamics along the reaction coordinate: critical role of a conserved residue. Biochemistry. 2006;45:2636–2647. doi: 10.1021/bi0525066. [DOI] [PubMed] [Google Scholar]
- 44.Kovrigin EL, Loria JP. Characterization of the transition state of functional enzyme dynamics. J. Am. Chem. Soc. 2006;128:7724–7725. doi: 10.1021/ja061435a. [DOI] [PubMed] [Google Scholar]
- 45.Kempf JG, Loria JP. Protein dynamics from solution NMR: theory and applications. Cell Biochem. Biophys. 2002;39:187–212. doi: 10.1385/cbb:37:3:187. [DOI] [PubMed] [Google Scholar]
- 46.Berlow RB, Igumenova TI, Loria JP. Value of a hydrogen bond in triosephosphate isomerase loop motion. Biochemistry. 2007;46:6001–6010. doi: 10.1021/bi700344v. [DOI] [PubMed] [Google Scholar]
- 47.Trott O, Palmer AG., 3rd R1rho relaxation outside of the fast-exchange limit. J. Magn. Reson. 2002;154:157–160. doi: 10.1006/jmre.2001.2466. [DOI] [PubMed] [Google Scholar]
- 48.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]
- 49.Loria JP, Rance M, Palmer AG. A TROSY CPMG sequence for characterizing chemical exchange in large proteins. J. Biomol. NMR. 1999;15:151–155. doi: 10.1023/a:1008355631073. [DOI] [PubMed] [Google Scholar]
- 50.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. Res. 1972;6:89–105. [Google Scholar]
- 51.Jen J. Chemical exchange and NMR T2 relaxation – the multisite case. J. Magn. Reson. 1978;30:111–128. [Google Scholar]
- 52.Davis DG, Perlman ME, London RE. Direct measurements of the dissociation-rate constant for inhibitor-enzyme complexes via the T1r and T2 (CPMG) methods. J. Magn. Reson., Ser. B. 1994;104:266–275. doi: 10.1006/jmrb.1994.1084. [DOI] [PubMed] [Google Scholar]
- 53.Orekhov VY, Korzhnev DM, Kay LE. Double- and zero-quantum NMR relaxation dispersion experiments sampling millisecond time scale dynamics in proteins. J. Am. Chem. Soc. 2004;126:1886–1891. doi: 10.1021/ja038620y. [DOI] [PubMed] [Google Scholar]
- 54.Korzhnev DM, Kloiber K, Kay LE. Multiple-quantum relaxation dispersion NMR spectroscopy probing millisecond time-scale dynamics in proteins: theory and application. J. Am. Chem. Soc. 2004;126:7320–7329. doi: 10.1021/ja049968b. [DOI] [PubMed] [Google Scholar]
- 55.Luz Z, Meiboom S. Nuclear magnetic resonance study of the protolysis of trimethylammonium ion in aqueous solution—order of the reaction with respect to solvent. J. Chem. Phys. 1963;39:366–370. [Google Scholar]
- 56.Ishima R, Torchia DA. Estimating the time scale of chemical exchange of proteins from measurements of transverse relaxation rates in solution. J. Biomol. NMR. 1999;14:369–372. doi: 10.1023/a:1008324025406. [DOI] [PubMed] [Google Scholar]
- 57.Millet O, Loria JP, Kroenke CD, Pons M, Palmer AG., III The static magnetic field dependence of chemical exchange linebroadening defines the NMR chemical shift time scale. J. Am. Chem. Soc. 2000;122:2867–2877. [Google Scholar]
- 58.Kovrigin EL, Kempf JG, Grey M, Loria JP. Faithful estimation of dynamics parameters from CPMG relaxation dispersion measurements. J. Magn. Reson. 2006;180:93–104. doi: 10.1016/j.jmr.2006.01.010. [DOI] [PubMed] [Google Scholar]
- 59.Isaacson RL, Simpson PJ, Liu M, Cota E, Zhang X, Freemont P, Matthews S. A new labeling method for methyl transverse relaxation-optimized spectroscopy NMR spectra of alanine residues. J. Am. Chem. Soc. 2007;129:15428–15429. doi: 10.1021/ja0761784. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Hansen DF, Vallurupalli P, Lundstrom P, Neudecker P, Kay LE. Probing chemical shifts of invisible states of proteins with relaxation dispersion NMR spectroscopy: how well can we do? J Am. Chem. Soc. 2008;130:2667–2675. doi: 10.1021/ja078337p. [DOI] [PubMed] [Google Scholar]
- 61.Vallurupalli P, Hansen DF, Kay LE. Structures of invisible, excited protein states by relaxation dispersion NMR spectroscopy. Proc. Natl. Acad. Sci. USA. 2008;105:11766–11771. doi: 10.1073/pnas.0804221105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Vallurupalli P, Hansen DF, Kay LE. Probing structure in invisible protein states with anisotropic NMR chemical shifts. J. Am. Chem. Soc. 2008;130:2734–2735. doi: 10.1021/ja710817g. [DOI] [PubMed] [Google Scholar]
- 63.Hwang PM, Bishop RE, Kay LE. The integral membrane enzyme PagP alternates between two dynamically distinct states. Proc. Natl. Acad. Sci. USA. 2004;101:9618–9623. doi: 10.1073/pnas.0402324101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Massi F, Wang C, Palmer AG., 3rd 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]
- 65.Ergel B, Gill ML, Brown L, Yu B, Palmer AG, 3rd, Hunt JF. Protein dynamics control the progression and efficiency of the catalytic reaction cycle of the Escherichia coli DNA-repair enzyme AlkB. J. Biol. Chem. 2014;289:29584–29601. doi: 10.1074/jbc.M114.575647. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Boehr DD, McElheny D, Dyson HJ, Wright PE. The dynamic energy landscape of dihydrofolate reductase catalysis. Science. 2006;313:1638–1642. doi: 10.1126/science.1130258. [DOI] [PubMed] [Google Scholar]
- 67.Zaccardi MJ, O’Rourke KF, Yezdimer EM, Loggia LJ, Woldt S, Boehr DD. Loop-loop interactions govern multiple steps in indole-3-glycerol phosphate synthase catalysis. Protein Sci. 2014;23:302–311. doi: 10.1002/pro.2416. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Wolf-Watz M, Thai V, Henzler-Wildman K, Hadjipavlou G, Eisenmesser EZ, Kern D. Linkage between dynamics and catalysis in a thermophilicmesophilic enzyme pair. Nat. Struct. Mol. Biol. 2004;11:945–949. doi: 10.1038/nsmb821. [DOI] [PubMed] [Google Scholar]
- 69.Eisenmesser EZ, Millet O, Labeikovsky W, Korzhnev DM, Wolf-Watz M, Bosco DA, Skalicky JJ, Kay LE, Kern D. Intrinsic dynamics of an enzyme underlies catalysis. Nature. 2005;438:117–121. doi: 10.1038/nature04105. [DOI] [PubMed] [Google Scholar]
- 70.Henzler-Wildman KA, Lei M, Thai V, Kerns SJ, Karplus M, Kern D. A hierarchy of timescales in protein dynamics is linked to enzyme catalysis. Nature. 2007;450:913–916. doi: 10.1038/nature06407. [DOI] [PubMed] [Google Scholar]
- 71.Kempf JG, Jung JY, Ragain C, Sampson NS, Loria JP. Dynamic requirements for a functional protein hinge. J. Mol. Biol. 2007;368:131–149. doi: 10.1016/j.jmb.2007.01.074. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Berlow RB, Swain M, Dalal S, Sweasy JB, Loria JP. Substrate-dependent millisecond domain motions in DNA polymerase beta. J. Mol. Biol. 2012;419:171–182. doi: 10.1016/j.jmb.2012.03.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Williams JC, McDermott AE. Dynamics of the flexible loop of triosephosphate isomerase: the loop motion is not ligand gated. Biochemistry. 1995;34:8309–8319. doi: 10.1021/bi00026a012. [DOI] [PubMed] [Google Scholar]
- 74.Larion M, Salinas RK, Bruschweiler-Li L, Bruschweiler R, Miller BG. Direct evidence of conformational heterogeneity in human pancreatic glucokinase from high-resolution nuclear magnetic resonance. Biochemistry. 2010;49:7969–7971. doi: 10.1021/bi101098f. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Davulcu O, Skalicky JJ, Chapman MS. Rate-limiting domain and loop motions in arginine kinase. Biochemistry. 2011;50:4011–4018. doi: 10.1021/bi101664u. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Stevens SY, Sanker S, Kent C, Zuiderweg ER. Delineation of the allosteric mechanism of a cytidylyltransferase exhibiting negative cooperativity. Nat. Struct. Biol. 2001;8:947–952. doi: 10.1038/nsb1101-947. [DOI] [PubMed] [Google Scholar]
- 77.Wang L, Pang Y, Holder T, Brender JR, Kurochkin AV, Zuiderweg ER. Functional dynamics in the active site of the ribonuclease binase. Proc. Natl. Acad. Sci. USA. 2001;98:7684–7689. doi: 10.1073/pnas.121069998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Namanja AT, Peng T, Zintsmaster JS, Elson AC, Shakour MG, Peng JW. Substrate recognition reduces side-chain flexibility for conserved hydrophobic residues in human Pin1. Structure. 2007;15:313–327. doi: 10.1016/j.str.2007.01.014. [DOI] [PubMed] [Google Scholar]
- 79.Masterson LR, Cheng C, Yu T, Tonelli M, Kornev A, Taylor SS, Veglia G. Dynamics connect substrate recognition to catalysis in protein kinase A. Nat. Chem. Biol. 2010;6:821–828. doi: 10.1038/nchembio.452. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Masterson LR, Shi L, Metcalfe E, Gao J, Taylor SS, Veglia G. Dynamically committed, uncommitted, and quenched states encoded in protein kinase A revealed by NMR spectroscopy. Proc. Natl. Acad. Sci. USA. 2011;108:6969–6974. doi: 10.1073/pnas.1102701108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Karamanou S, Bariami V, Papanikou E, Kalodimos CG, Economou A. Assembly of the translocase motor onto the preprotein-conducting channel. Mol. Microbiol. 2008;70:311–322. doi: 10.1111/j.1365-2958.2008.06402.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Karamanou S, Gouridis G, Papanikou E, Sianidis G, Gelis I, Keramisanou D, Vrontou E, Kalodimos CG, Economou A. Preprotein-controlled catalysis in the helicase motor of SecA. EMBO J. 2007;26:2904–2914. doi: 10.1038/sj.emboj.7601721. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Gelis I, Bonvin AM, Keramisanou D, Koukaki M, Gouridis G, Karamanou S, Economou A, Kalodimos CG. Structural basis for signal-sequence recognition by the translocase motor SecA as determined by NMR. Cell. 2007;131:756–769. doi: 10.1016/j.cell.2007.09.039. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Sarkar P, Reichman C, Saleh T, Birge RB, Kalodimos CG. Proline cis-trans isomerization controls autoinhibition of a signaling protein. Mol. Cell. 2007;25:413–426. doi: 10.1016/j.molcel.2007.01.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Freiburger LA, Baettig OM, Sprules T, Berghuis AM, Auclair K, Mittermaier AK. Competing allosteric mechanisms modulate substrate binding in a dimeric enzyme. Nat. Struct. Mol. Biol. 2011;18:288–294. doi: 10.1038/nsmb.1978. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Mandel AM, Akke M, Palmer AG. 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]
- 87.Akerud T, Thulin E, Van Etten RL, Akke M. Intramolecular dynamics of low molecular weight protein tyrosine phosphatase in monomer-dimer equilibrium studied by NMR: a model for changes in dynamics upon target binding. J. Mol. Biol. 2002;322:137–152. doi: 10.1016/s0022-2836(02)00714-3. [DOI] [PubMed] [Google Scholar]
- 88.Brath U, Akke M. Differential responses of the backbone and side-chain conformational dynamics in FKBP12 upon binding the transition-state analog FK506: implications for transition-state stabilization and target protein recognition. J. Mol. Biol. 2009;387:233–244. doi: 10.1016/j.jmb.2009.01.047. [DOI] [PubMed] [Google Scholar]
- 89.Stone MJ, Fairbrother WJ, Palmer AG, 3rd, Reizer J, Saier MH, Jr, Wright PE. Backbone dynamics of the Bacillus subtilis glucose permease IIA domain determined from 15N NMR relaxation measurements. Biochemistry. 1992;31:4394–4406. doi: 10.1021/bi00133a003. [DOI] [PubMed] [Google Scholar]
- 90.Gao G, Semenchenko V, Arumugam S, Van Doren SR. Tissue inhibitor of metalloproteinases-1 undergoes microsecond to millisecond motions at sites of matrix metalloproteinase-induced fit. J. Mol. Biol. 2000;301:537–552. doi: 10.1006/jmbi.2000.3976. [DOI] [PubMed] [Google Scholar]
- 91.Ding Z, Lee G, Liang X, Gallazzi F, Arunima A, Van Doren SR. PhosphoThr peptide binding globally rigidifies much of the FHA domain from arabidopsis receptor kinase-associated protein phosphatase. Biochemistry. 2005;44:10119–10134. doi: 10.1021/bi050414a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Ishima R, Freedberg DI, Wang YX, Louis JM, Torchia DA. Flap opening and dimer-interface flexibility in the free and inhibitor-bound HIV protease, and their implications for function. Structure. 1999;7:1047–1055. doi: 10.1016/s0969-2126(99)80172-5. [DOI] [PubMed] [Google Scholar]
- 93.Ishima R, Louis JM. A diverse view of protein dynamics from NMR studies of HIV-1 protease flaps. Proteins. 2008;70:1408–1415. doi: 10.1002/prot.21632. [DOI] [PubMed] [Google Scholar]
- 94.Koveal D, Clarkson MW, Wood TK, Page R, Peti W. Ligand binding reduces conformational flexibility in the active site of tyrosine phosphatase related to biofilm formation A (TpbA) from Pseudomonasaeruginosa. J. Mol. Biol. 2013;425:2219–2231. doi: 10.1016/j.jmb.2013.03.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Peti W, Nairn AC, Page R. Structural basis for protein phosphatase 1 regulation and specificity. FEBS J. 2013;280:596–611. doi: 10.1111/j.1742-4658.2012.08509.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Hammes GG. Multiple conformational changes in enzyme catalysis. Biochemistry. 2002;41:8221–8228. doi: 10.1021/bi0260839. [DOI] [PubMed] [Google Scholar]
- 97.Cathou RE, Hammes GG. Relaxation spectra of ribonuclease. III. Further investigation of the interaction of ribonuclease and cytidine 3’-phosphate. J. Am. Chem. Soc. 1965;87:4674–4680. doi: 10.1021/ja00949a003. [DOI] [PubMed] [Google Scholar]
- 98.Akke M. Molecular Basis for Cooperativity in Calcium Binding by Calbindin D9k as Studied by NMR. Lund University; 1993. [Google Scholar]
- 99.Stivers JT, Abeygunawardana C, Mildvan AS. 15N NMR relaxation studies of free and inhibitor-bound 4-oxalocrotonate tautomerase: backbone dynamics and entropy changes of an enzyme upon inhibitor binding. Biochemistry. 1996;35:16036–16047. doi: 10.1021/bi961834q. [DOI] [PubMed] [Google Scholar]
- 100.Kovrigin EL, Cole R, Loria JP. Temperature dependence of the backbone dynamics of Ribonuclease A in the ground state and bound to the inhibitor 5′-phosphothymidine (3′–5′) pyrophosphate adenosine 3′-phosphate. Biochemistry. 2003;42:5279–5291. doi: 10.1021/bi034027h. [DOI] [PubMed] [Google Scholar]
- 101.Raines RT. Ribonuclease A. Chem. Rev. 1998;98:1045–1066. doi: 10.1021/cr960427h. [DOI] [PubMed] [Google Scholar]
- 102.Zegers I, Maes D, Dao-Thi MH, Poortmans F, Palmer R, Wyns L. The structures of RNase A complexed with 3′-CMP and d(CpA): active site conformation and conserved water molecules. Protein Sci. 1994;3:2322–2339. doi: 10.1002/pro.5560031217. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Grey MJ, Wang C, Palmer AG., 3rd 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]
- 104.Skrynnikov NR, Dahlquist FW, Kay LE. Reconstructing NMR spectra of “invisible” excited protein states using HSQC and HMQC experiments. J. Am. Chem. Soc. 2002;124:12352–12360. doi: 10.1021/ja0207089. [DOI] [PubMed] [Google Scholar]
- 105.Beach H, Cole R, Gill ML, Loria JP. Conservation of µs–ms enzyme motions in the apo- and substrate-mimicked state. J. Am. Chem. Soc. 2005;127:9167–9176. doi: 10.1021/ja0514949. [DOI] [PubMed] [Google Scholar]
- 106.Aguilar CF, Thomas PJ, Moss DS, Mills A, Palmer RA. Novel non-productively bound ribonuclease inhibitor complexes-high-resolution X-ray refinement studies on the binding of Rnase-A to cytidylyl-2′,5′-guanosine (2′,5′CPG) and deoxycytidylyl-3′,5′-guanosine (3′,5′DCPDG) Biochim. Biophys. Acta. 1991;118:6–20. doi: 10.1016/0167-4838(91)90435-3. [DOI] [PubMed] [Google Scholar]
- 107.McPherson A, Brayer G, Cascio D, Williams R. The mechanism of binding of a polynucleotide chain to pancreatic ribonuclease. Science. 1986;232:765–768. doi: 10.1126/science.3961503. [DOI] [PubMed] [Google Scholar]
- 108.Watt ED, Shimada H, Kovrigin EL, Loria JP. The mechanism of rate-limiting motions in enzyme function. Proc. Natl. Acad. Sci. USA. 2007;104:11981–11986. doi: 10.1073/pnas.0702551104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Markley JL. Correlation proton magnetic resonance studies at 250 MHz of bovine pancreatic ribonuclease. I. Reinvestigation of the histidine peak assignments. Biochemistry. 1975;14:3546–3554. doi: 10.1021/bi00687a006. [DOI] [PubMed] [Google Scholar]
- 110.Markley JL. Correlation proton magnetic resonance studies at 250 MHz of bovine pancreatic ribonuclease. II. pH and inhibitor-induced conformational transitions affecting histidine-48 and one tyrosine residue of ribonuclease A. Biochemistry. 1975;14:3554–3561. [PubMed] [Google Scholar]
- 111.Doucet N, Watt ED, Loria JP. The flexibility of a distant loop modulates active site motion and product release in ribonuclease A. Biochemistry. 2009;48:7160–7168. doi: 10.1021/bi900830g. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Doucet N, Khirich G, Kovrigin EL, Loria JP. The alteration of hydrogen bonding in the vicinity of histidine 48 disrupts millisecond motions in RNase A. Biochemistry. 2011;50:1723–1730. doi: 10.1021/bi1018539. [DOI] [PubMed] [Google Scholar]
- 113.Watt ED, Rivalta I, Whittier SK, Batista VS, Loria JP. Reengineering rate-limiting, millisecond enzyme motions by introduction of an unnatural amino acid. Biophys. J. 2011;101:411–420. doi: 10.1016/j.bpj.2011.05.039. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Khirich G, Loria JP. Complexity of protein energy landscapes studied by solution NMR relaxation dispersion experiments. J. Phys. Chem. B. 2015;119:3743–3754. doi: 10.1021/acs.jpcb.5b00212. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Lee GC, Chan SI. A 31P NMR study of the association of uridine-3′-monophosphate to ribonuclease A. Biochem. Biophys. Res. Commun. 1971;43:142–148. doi: 10.1016/s0006-291x(71)80098-0. [DOI] [PubMed] [Google Scholar]
- 116.Zhang ZY. Protein tyrosine phosphatases: prospects for therapeutics. Curr. Opin. Chem. Biol. 2001;5:416–423. doi: 10.1016/s1367-5931(00)00223-4. [DOI] [PubMed] [Google Scholar]
- 117.Brubaker RR. Factors promoting acute and chronic diseases caused by yersiniae. Clin. Microbiol. Rev. 1991;4:309–324. doi: 10.1128/cmr.4.3.309. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Guan KL, Dixon JE. Evidence for protein-tyrosine-phosphatase catalysis proceeding via a cysteine-phosphate intermediate. J. Biol. Chem. 1991;266:17026–17030. [PubMed] [Google Scholar]
- 119.Hengge AC, Sowa GA, Wu L, Zhang ZY. Nature of the transition state of the protein-tyrosine phosphatase-catalyzed reaction. Biochemistry. 1995;34:13982–13987. doi: 10.1021/bi00043a003. [DOI] [PubMed] [Google Scholar]
- 120.Denu JM, Lohse DL, Vijayalakshmi J, Saper MA, Dixon JE. Visualization of intermediate and transition-state structures in protein-tyrosine phosphatase catalysis. Proc. Natl. Acad. Sci. USA. 1996;93:2493–2498. doi: 10.1073/pnas.93.6.2493. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Brandao TA, Hengge AC, Johnson SJ. Insights into the reaction of protein-tyrosine phosphatase 1B: crystal structures for transition state analogs of both catalytic steps. J. Biol. Chem. 2010;285:15874–15883. doi: 10.1074/jbc.M109.066951. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Zhang ZY, Thieme-Sefler AM, Maclean D, McNamara DJ, Dobrusin EM, Sawyer TK, Dixon JE. Substrate specificity of the protein tyrosine phosphatases. Proc. Natl. Acad. Sci. USA. 1993;90:4446–4450. doi: 10.1073/pnas.90.10.4446. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Zhang ZY. Kinetic and mechanistic characterization of a mammalian protein-tyrosine phosphatase, PTP1. J. Biol. Chem. 1995;270:11199–11204. doi: 10.1074/jbc.270.19.11199. [DOI] [PubMed] [Google Scholar]
- 124.Whittier SK, Hengge AC, Loria JP. Conformational motions regulate phosphoryl transfer in related protein tyrosine phosphatases. Science. 2013;341:899–903. doi: 10.1126/science.1241735. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Pedersen AK, Peters GG, Moller KB, Iversen LF, Kastrup JS. Water-molecule network and active-site flexibility of apo protein tyrosine phosphatase 1B. Acta Crystallogr D Biol. Crystallogr. 2004;60:1527–1534. doi: 10.1107/S0907444904015094. [DOI] [PubMed] [Google Scholar]
- 126.Hu X, Stebbins CE. Dynamics of the WPD loop of the Yersinia protein tyrosine phosphatase. Biophys. J. 2006;91:948–956. doi: 10.1529/biophysj.105.080259. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Juszczak LJ, Zhang ZY, Wu L, Gottfried DS, Eads DD. Rapid loop dynamics of Yersinia protein tyrosine phosphatases. Biochemistry. 1997;36:2227–2236. doi: 10.1021/bi9622130. [DOI] [PubMed] [Google Scholar]
- 128.Khajehpour M, Wu L, Liu S, Zhadin N, Zhang ZY, Callender R. Loop dynamics and ligand binding kinetics in the reaction catalyzed by the Yersinia protein tyrosine phosphatase. Biochemistry. 2007;46:4370–4378. doi: 10.1021/bi602335x. [DOI] [PubMed] [Google Scholar]
- 129.Burke TR, Jr, Kole HK, Roller PP. Potent inhibition of insulin receptor dephosphorylation by a hexamer peptide containing the phosphotyrosyl mimetic F2Pmp. Biochem. Biophys. Res. Commun. 1994;204:129–134. doi: 10.1006/bbrc.1994.2435. [DOI] [PubMed] [Google Scholar]
- 130.Chatani E, Hayashi R, Moriyama H, Ueki T. Conformational strictness required for maximum activity and stability of bovine pancreatic ribonuclease A as revealed by crystallographic study of three Phe120 mutants at 1.4 A resolution. Protein Sci. 2002;11:72–81. doi: 10.1110/ps.31102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Mallorqui-Fernandez G, Pous J, Peracaula R, Aymami J, Maeda T, Tada H, Yamada H, Seno M, de Llorens R, Gomis-Ruth FX, Coll M. Three-dimensional crystal structure of human eosinophil cationic protein (RNase 3) at 1.75 angstrom resolution. J. Mol. Biol. 2000;300:1297–1307. doi: 10.1006/jmbi.2000.3939. [DOI] [PubMed] [Google Scholar]
- 132.Armougom F, Moretti S, Poirot O, Audic S, Dumas P, Schaeli B, Keduas V, Notredame C. Expresso: automatic incorporation of structural information in multiple sequence alignments using 3D-Coffee. Nucleic Acids Res. 2006;34:W604–W608. doi: 10.1093/nar/gkl092. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Wlodawer A, Svensson LA, Sjolin L, Gilliland GL. Structure of phosphate-free ribonuclease A refined at 1.26 A. Biochemistry. 1988;27:2705–2717. doi: 10.1021/bi00408a010. [DOI] [PubMed] [Google Scholar]
- 134.Swaminathan GJ, Holloway DE, Veluraja K, Acharya KR. Atomic resolution (0.98 A) structure of eosinophil-derived neurotoxin. Biochemistry. 2002;41:3341–3352. doi: 10.1021/bi015911f. [DOI] [PubMed] [Google Scholar]



















