Skip to main content
Protein Science : A Publication of the Protein Society logoLink to Protein Science : A Publication of the Protein Society
. 2012 Feb 14;21(4):562–570. doi: 10.1002/pro.2045

Thermal coefficients of the methyl groups within ubiquitin

T Michael Sabo 1, Davood Bakhtiari 1, Korvin F A Walter 1, Robert L McFeeters 2, Karin Giller 1, Stefan Becker 1, Christian Griesinger 1,*, Donghan Lee 1,*
PMCID: PMC3375756  PMID: 22334336

Abstract

Physiological processes such as protein folding and molecular recognition are intricately linked to their dynamic signature, which is reflected in their thermal coefficient. In addition, the local conformational entropy is directly related to the degrees of freedom, which each residue possesses within its conformational space. Therefore, the temperature dependence of the local conformational entropy may provide insight into understanding how local dynamics may affect the stability of proteins. Here, we analyze the temperature dependence of internal methyl group dynamics derived from the cross-correlated relaxation between dipolar couplings of two CH bonds within ubiquitin. Spanning a temperature range from 275 to 308 K, internal methyl group dynamics tend to increase with increasing temperature, which translates to a general increase in local conformational entropy. With this data measured over multiple temperatures, the thermal coefficient of the methyl group order parameter, the characteristic thermal coefficient, and the local heat capacity were obtained. By analyzing the distribution of methyl group thermal coefficients within ubiquitin, we found that the N-terminal region has relatively high thermostability. These results indicate that methyl groups contribute quite appreciably to the total heat capacity of ubiquitin through the regulation of local conformational entropy.

Keywords: NMR, methyl group dynamics, cross-correlated relaxation, thermal coefficients, protein stability, ubiquitin

Introduction

The intricate relationship between structure, function, and dynamics strongly influences such fundamental physiological processes as protein folding, molecular recognition, and thermal stability. All of these processes occur with concomitant changes in the thermodynamic parameters of the system, specifically the enthalpy (H) and the entropy (S). The linkage between the temperature dependence of H and S is expressed by the heat capacity (Cp).1, 2 Contributions to the Cp of a protein include hydration of solvent exposed surface area, covalent bonds, electrostatic interactions, and hydrogen bonds. In addition, the local conformational entropy (Sconf) of residues within a protein must also be considered. The term Sconf is directly related to the degrees of freedom each residue possesses within the three-dimensional protein structure. Therefore, an understanding of the dependence of Sconf on temperature, encapsulated by Cp, yields insight into the role local dynamics play in controlling the stability of proteins.

Nuclear magnetic resonance (NMR) spectroscopy provides a potent tool for characterizing such local dynamics with atomic resolution on multiple time-scales. Analysis of spin-lattice relaxation (T1), spin-spin relaxation (T2), and steady-state nuclear Overhauser enhancements3 with the “model free” formalism introduced by Lipari and Szabo allows the extraction of a general order parameter (S2) that describes the amplitude of the individual bond vector motions on the ps to ns time-scale.4, 5 The relationship between S2 and Sconf results from the dependence of both parameters on the population distribution of bond vector orientations.68 By measuring relaxation data at multiple temperatures, the temperature dependence of S2 ultimately provides the local Cp for each bond vector.913

An important probe ideally suited for providing fundamental insight into protein folding, stability, and recognition are methyl groups. Located primarily within the hydrophobic core of proteins, methyl groups are typically quite numerous, well dispersed throughout the core, and possess a wide variety of motional amplitudes.9,1416 For these reasons, determining the methyl group order parameters (Inline graphic) provides an avenue for understanding how methyl dynamics on the ps to ns time scale are related to protein stability. To obtain Inline graphic, typically the methyl groups in a protein are deuterated (–CH2D or –CHD2) for performing deuterium relaxation measurements.1719 An alternative to this approach is to utilize cross-correlated relaxation (CCR) between dipolar couplings of two CH bonds (σ) in the methyl group for extracting Inline graphic.20, 21

An inherent advantage to determining Inline graphic from σ versus deuterium relaxation is the savings in measurement time, especially important for measuring Inline graphic at many different temperatures. In the past, Inline graphic calculated at one temperature point with deuterium relaxation studies could take up to 1 week due to the requirement of data being recorded at two fields.17 Recent advances in determining Inline graphic from the measurement of five relaxation rates for –CH2D19 and four relaxation rates for –CHD218 has effectively reduced the amount of time necessary to obtain the same information to ∼1 day. Yet, Inline graphic derived from σ can require as little as 1 h of measurement time and, thus, is ideally suited for studying the temperature dependence of Inline graphic.

Here, we report the temperature dependence of the methyl group order parameters derived from the CCR between dipolar couplings of two CH bonds in ubiquitin. From these measurements, we calculate the thermal coefficients characterizing this temperature dependence, specifically the characteristic thermal coefficient Λ and the local heat capacity Cp. Furthermore, we analyze the distribution of methyl group thermal coefficients within ubiquitin, illustrating the relatively high thermostability of the N-terminal region of this protein.

Results and Discussion

Methyl group CCR

Two dimensional constant time 13C, 1H HSQC measurements without decoupling of 1H during 13C-chemical shift evolution leads to splitting of the methyl group 13C signal into a quartet.22, 23 The peaks in the quartet are separated by the J-coupling constant for C–H bonds in a methyl group (JCH = 125 Hz). The quartet represents the four coherences of Inline graphic, Inline graphic Inline graphic, Inline graphic Inline graphic, and Inline graphicand their intensity ratio (Inline graphic) is 3:1:1:3 in the absence of relaxation.22, 24 With the contribution of the transverse relaxation rate (Inline graphic),25 the intensities can be expressed as:

graphic file with name pro0021-0562-m1.jpg (1)

where Δ is the length of the constant time period.

By considering the dipolar coupling and chemical shift anisotropy (CSA), Inline graphiccan be expressed as:20, 25

graphic file with name pro0021-0562-m2.jpg (2)

where λ is the rate of the autorelaxation and σ and η are the rate of the CCR between dipolar couplings of two CH bonds and between dipolar coupling of CH and CSA of the 13C nucleus, respectively. Thus, the CCR rate between dipolar couplings of two CH bonds can be determined experimentally from the intensities of the quartet,20

graphic file with name pro0021-0562-m3.jpg (3)

Here, we report the temperature dependence of σobs for the methyl groups of uniformly 15N, 13C-labeled human ubiquitin extracted from a series of 2D constant time 13C, 1H HSQC measurements at fourteen temperatures: 275, 278, 281, 283, 286, 288, 291, 293, 296, 298, 301, 303, 305, and 308 K. Ubiquitin possesses 50 methyl groups residing in 30 residues. Immediately clear from the measurement at 275 K (see Fig. 1), the peaks in the highlighted quartets are well resolved, despite the increase in τc accompanied with lowering the temperature. In addition, even at 275 K, a significant amount of motion is evident for the methyl groups of ubiquitin, especially for L8δ1 whose quartet approaches the ideal 3:1:1:3 intensity ratio.

Figure 1.

Figure 1

Constant time 13C, 1H HSQC spectrum of uniformly 15N,13C-labeled human ubiquitin recorded at a proton frequency of 700 MHz and a temperature of 275 K. The concentration of ubiquitin was 3.6 mM in 90%/10% H2O/D2O, containing 50 mM sodium phosphate at pH 6.8, 100 mM NaCl, and 0.1% NaN3. The constant time duration and INEPT delays were set to 27.8 and 2 ms, respectively. The spectrum was recorded with 1024 and 128 complex points in the direct (t2) and indirect (t1) dimensions, respectively, with eight scans per t1 increment. The t1,max and t2,max were 24.3 ms and 113 ms, respectively. Frequency discrimination in the indirectly detected dimension was achieved with the States-TPPI scheme.43 The spectrum was processed with NMRPipe software.44 1D slices of selected quartets are illustrated together with the corresponding coherences for each peak in the multiplet.

For nearly 50% of all methyl groups in ubiquitin, Eq. (3) was used to calculate σobs at each of the 14 temperatures. The results are compiled in Supporting Information Table SI. For the remaining 28 methyl groups, either spectral overlap becomes problematic due to the chemical shift differences between methyl group carbons being similar to JCH, 2JCH, or 3JCH20 and/or strong coupling is present between the δ and γ carbons in leucine as reported for L15δ1, L43δ1, L50δ1, L56δ1, and L69δ2.21 In addition, we only analyzed methyl groups whose quartets were not overlapped with other methyl group quartets over the entire temperature range (275–308 K).

As evident from the range of σobs at each temperature, the methyl groups exist in a wide array of environments. Variations in the mobility of the methyl group contribute to the observed differences in σobs. It should be noted that studying the temperature dependence of methyl group dynamics with deuterium relaxation studies,1719, 26 though more time consuming, enables many more methyl groups to be analyzed due to significantly less spectral overlap. Nevertheless, with the wide availability of residue selective methyl group labeling,2729 which now even includes stereospecific selection of methyl groups,30 we anticipate this approach being applicable to proteins larger than ubiquitin.

Quantification of methyl group dynamics in ubiquitin

The experimental CCR rate can be used to extract the methyl group order parameter (Inline graphic) by comparing σobs with the theoretical value of the CCR (σrigid) in the absence of local motions20, 25, 31:

graphic file with name pro0021-0562-m4.jpg (4)
graphic file with name pro0021-0562-m5.jpg (5)

where μ0 is the permeability of a vacuum, h is the Planck constant, γH and γC are the gyromagnetic ratios of 1H and 13C, respectively, rCH is the CH bond length, ωC is the Larmor frequency of 13C and τc is the rotational correlation time. Inline graphic is a dimensionless quantity utilized for describing the amplitude of motions for the methyl group.25 Values for Inline graphic range from 0 to 1, where 1 represents a rigid methyl group and zero represents unrestricted local motion.

For determination of σrigid with Eq. (5), the methyl group C–H bond length (rCH) was taken as 1.095 angstroms, tetrahedral geometry assumed and the τc of ubiquitin at each temperature is reported in Supporting Information Table SII. With σrigid, we calculated Inline graphic with Eq. (4) for all 14 temperatures, presented in the Supporting Information Table SII. Figure 2 illustrates the correlation of calculated Inline graphic at a selected set of temperatures. Readily apparent from the figure is the high degree of correlation for Inline graphic over these temperatures. For every pair-wise combination of Inline graphic, the Pearson correlation coefficient is r ≥ 0.98. Since Inline graphic has a linear dependence on T, the high correlation between Inline graphic at all the temperatures suggests that uncertainties arising from each experimental measurement are consistent over the entire data set. Figure 3(A) presents the correlation plot of Inline graphic values from CCR measurements versus Inline graphic extracted from the rates of multiple spin coherences involving 2H in the methyl group, both at 303 K.26 Figure 3(B) is the correlation plot of the currently reported Inline graphic at 301 K versus Inline graphic extracted from methyl group 13C spin-lattice (T1) relaxation rates and σ modulated by the one bond C–H coupling constant (JCH) measured at 300 K.21 For both comparisons, the Pearson correlation coefficient is high, 0.96 and 0.97, respectively. The agreement between Inline graphic obtained by three independent studies coupled with the high correlation for Inline graphic over the entire temperature range provides strong confirmation that the temperature dependence of Inline graphic can be studied using this method. It is important to point out that deviations from the diagonal most likely indicate small differences in experimental setup, such as sample conditions, slight temperature variations, and/or measurement error. Most importantly, we have significantly reduced the total amount of time required to obtain this information. With methods involving 2H labeling of methyl groups,1719R1 and R2 relaxation rates are typically measured in order to extract Inline graphic at one temperature. As for σ modulated by the one bond C–H coupling constant (JCH), Inline graphic determined at one temperature needs a series of 2D constant time 13C, 1H HSQC measurements with increasing delay times.21 However, with the present method, one temperature point requires only one 2D constant time 13C, 1H HSQC measurement.

Figure 2.

Figure 2

Correlations between methyl group order parameters (Inline graphic) of ubiquitin determined at multiple temperatures. (A) 283 K vs. 275 K, (B) 293 K vs. 275 K, (C) 303 K vs. 275 K, (D) 283 K vs. 281 K, (E) 293 K vs. 281 K, and (F) 303 K vs. 281 K. The solid line represents the best fit. The error bars for Inline graphic in both dimensions are included and in many cases are smaller than the symbol size. The Pearson correlation r is also shown.

Figure 3.

Figure 3

Comparison of Inline graphic in ubiquitin determined from different methods. A: Correlation between Inline graphic obtained from this study using dipolar CCR between CH bonds in the methyl group versus Inline graphic extracted from the rates of multiple spin coherences involving 2H in the methyl group.26 In both cases, Inline graphic was obtained at 303 K. B: Correlation between Inline graphic obtained at 301 K from the present study versus Inline graphic extracted at 300 K from methyl group 13C spin-lattice (T1) relaxation rates and dipolar CCR modulated by the one bond C–H coupling constant (JCH).21 In both figures, the dashed lines represent the best fit to the data. The error bars are included in both dimensions. The Pearson correlation r is also indicated in the figures.

Finally, the time-scale of motion encompassed by Inline graphic is faster than the τc of ubiquitin, describing ps to ns dynamics. An earlier study from our group has determined the methyl group RDC-based order parameters Inline graphic for ubiquitin,32 which reflect time-scales from ps to ms. When comparing the two sets of order parameters at 308 K, the average values of Inline graphic and Inline graphic are 0.59 ± 0.21 and 0.43 ± 0.25, respectively, with a Pearson correlation coefficient of r = 0.84. Clearly, the methyl groups possess additional dynamics on a slower time-scale than encapsulated by Inline graphic.

Analysis of the temperature dependence of Inline graphic in ubiquitin

To quantify the temperature dependencies of Inline graphic, the thermal coefficient κ was taken from the linear fit of Inline graphic versus temperature Inline graphic.15 Table I and Figure 4 present the results of the fitting procedure. An overall decrease is observed in Inline graphic as the temperature increases from 275 to 308 K. The average value of κ is −(2.9 ± 1.5) × 10−3 K−1, which is quite comparable to the value of −(2.6 ± 1.1) × 10−3 K−1 reported previously for ubiquitin over the larger temperature range of 278–328 K.16 Furthermore, similar trends in the deviation of Inline graphic with temperature are also observed for calmodulin bound to a peptide.15 Figure 5(A) illustrates the distribution of κ within ubiquitin (1UBQ).33 The largest κ values are clustered around I61δ1, spatially near the N-terminus of ubiquitin, and progressively decrease toward the C-terminus of the protein.

Table I.

Thermal Coefficients Extracted from Methyl Group Order Parameters (Inline graphic) Within Ubiquitina

κ (× 10−3 K−1)b Λc Cp (J K−1mol−1)d
I3γ2 −2.9 ± 0.4 7.5 ± 1.4 52.7 ± 11.9e
I3δ1 −2.5 ± 0.4 3.2 ± 0.5 27.3 ± 4.5
T7γ2 −0.8 ± 0.5 1.0 ± 0.8 8.5 ± 7.9
L8δ1 −2.3 ± 0.2 1.4 ± 0.1 13.1 ± 1.1
T9γ2 −3.0 ± 0.3 2.7 ± 0.1 23.4 ± 2.8
T12γ2 −4.4 ± 0.4 4.8 ± 0.4 41.5 ± 4.9
I13δ1 −3.2 ± 0.3 2.5 ± 0.2 22.4 ± 2.3
V17γ2 −3.1 ± 0.4 4.0 ± 0.4 34.7 ± 4.3
T22γ2 −3.3 ± 0.5 8.1 ± 0.4 59.7 ± 14.1e
I23γ2 −2.3 ± 0.4 5.5 ± 1.1 43.9 ± 11.4f
I23δ1 −4.3 ± 0.4 3.1 ± 0.2 27.6 ± 2.6
I36γ2 0.1 ± 0.4 −0.8 ± 0.3 −6.9 ± 3.7
I36δ1 −2.8 ± 0.3 2.2 ± 0.2 19.8 ± 2.4
L43δ2 −3.6 ± 0.4 2.6 ± 0.2 23.4 ± 2.4
I44δ1 −2.6 ± 0.3 1.5 ± 0.2 14.8 ± 1.7
L50δ2 −5.4 ± 0.6 7.5 ± 0.6 64.5 ± 7.7
L56δ2 −5.2 ± 0.5 5.7 ± 0.4 49.3 ± 4.9
I61γ2 −2.5 ± 0.3 4.7 ± 0.6 39.5 ± 6.6
I61δ1 −5.7 ± 0.4 4.6 ± 0.3 40.4 ± 3.0
L67δ2 −2.1 ± 0.2 1.2 ± 0.1 11.8 ± 1.2
V70γ2 −2.3 ± 0.3 1.4 ± 0.1 13.4 ± 1.5
L73δ1 0.0 ± 0.1 0.2 ± 0.1 1.6 ± 0.7
Average −2.9 ± 1.5 3.4 ± 2.4 28.5 ± 18.9
a

All errors were determined from 500 Monte Carlo simulation runs.

b

The temperature dependencies of Inline graphic (Inline graphic) were obtained from the slope of a linear fit of Inline graphic versus temperature (T).15

c

The characteristic thermal coefficient Λ was obtained from the slope of a linear fit of Inline graphicversus lnT.11

d

The heat capacities (Cp) were obtained from the slope of a linear fit of the conformational entropy (Sconf) versus lnT,34 where Inline graphic8,9kB and NA are the Boltzmann constant and Avogadro's number, respectively.

e

For these methyl groups, only Inline graphic from temperatures between 283 and 308 K were used for the fitting procedure due to the requirement of Inline graphic < 0.95 for determining Sconf8.

f

For this methyl group, only Inline graphic from temperatures between 278 and 308 K were used for the fitting procedure due to the requirement of Inline graphic< 0.95 for determining of Sconf8.

Figure 4.

Figure 4

Temperature dependence of Inline graphicin ubiquitin (Inline graphic). All data were fit with a linear regression line in order to obtain κ (solid line).15 For many of the data points, the error bars are smaller than the symbol size. A: Plot of the temperature dependence of Inline graphic for the following leucine δ methyl groups: 8δ1 (○), 43δ2 (□), 50δ2 (⋄), 56δ2 (×), 67δ2 (•), and 73δ1 (▪). B: Plot of the temperature dependence of Inline graphic for the following isoleucine δ1 methyl groups: 3 (○), 13 (⋄), 23 (×), 36 (▵), 44 (•), and 61 (□). C: Plot of the temperature dependence of Inline graphic for the following isoleucine γ2 methyl groups: 3 (○), 23 (⋄), 36 (▵), and 61 (□). D: Plot of the temperature dependence of Inline graphic for the following threonine γ2 methyl groups: 7 (○), 9 (⋄), 12 (□), and 22 (▵). E: Plot of the temperature dependence of Inline graphic for the following valine γ2 methyl groups: 17 (○) and 70 (□).

Figure 5.

Figure 5

Location of the calculated thermal coefficients for methyl groups in ubiquitin (1UBQ).33 A: Distribution of κ. B: Distribution of the characteristic thermal coefficient Λ, which was obtained from a linear fit of Inline graphicversus ln T.11 C: Distribution of the heat capacity (Cp), which was obtained from a linear fit of the conformational entropy (Sconf) versus ln T,34 where Inline graphic8, 9. kB and NA are the Boltzmann constant and Avogadro's number, respectively. Spheres represent the carbon atom of the indicated methyl group. The color gradient is employed for identifying the magnitude of the specified thermal parameter. The figures were created with the program MOLMOL.46

The characteristic thermal coefficient Inline graphic Inline graphicrelates the temperature dependencies of the generalized order parameter, in this case Inline graphic, to the characteristic temperature (T*).10, 11 The term T* describes the density of thermally accessible conformational states.10 Table I presents the results for the determination of Λ, which correlates with κ (Pearson coefficients of r = 0.7). The average value of Λ (3.4 ± 2.4 compared with 2.3 ± 1.0 previously reported for ubiquitin over the larger temperature range of 278–328 K16) indicates that there is significant contributions from rotameric state dynamics.14 In Figure 5(B), Λ is plotted on the structure of ubiquitin. As with κ, the largest values of Λ are located in the N-terminal region of ubiquitin. Toward the C-terminal region a progressive decline of Λ is visible.

Determination of the Cp from the temperature dependence of the conformational entropy

An approximate relationship between the conformational entropy (Sconf) and Inline graphic is

graphic file with name pro0021-0562-m6.jpg (6)

where kB is the Boltzmann constant and NA is Avogadro's number.8 The heat capacity (Cp) obtained from the temperature dependence of Sconf is defined as34

graphic file with name pro0021-0562-m7.jpg (7)

Table I presents the values of Cp for the methyl groups in ubiquitin for the temperature interval of 275–308 K. For 22 methyl groups in ubiquitin, we report an average Cp of 28.5 ± 18.9 J K−1 mol−1. For the drkN SH3 domain, similar values for the methyl group Cp were obtained: 17 ± 12 J K−1 mol−1 and 33 ± 23 J K−1 mol−1 for the temperature intervals 287–303 K and 278–287 K, respectively.9 As measured by differential scanning calorimetry, the global Cp for ubiquitin was ∼12.6 kJ K−1 mol−1 at 298 K.35 In this study, the summation of all the individual methyl group Cp equals 626 J K−1 mol−1. Keeping in mind that this value does not include the contributions of 28 additional methyl groups where Cp could not be determined, these results suggest that the total methyl group Cp makes a ∼10% contribution to the total Cp of ubiquitin. It should be noted that a majority Cp can be determined from the primary sequence of proteins.1 Since 16% (200 out of 1231) of all atoms in ubiquitin are from the methyl groups, the ∼10% contribution for the methyl groups to the total ubiquitin Cp seems to be a reasonable estimate. Figure 5(C) presents the distribution of Cp in ubiquitin. The largest values of Cp, as with κ and Λ, are located spatially near the N-terminus of the protein, which is the part of the hydrophobic core of the protein.

Insight into the thermal stability of ubiquitin from the distribution of thermal coefficients

A striking feature regarding the distribution of the thermal coefficients is the concentration of the largest values near the N-terminus of the protein. For these methyl groups to be located within the core of the protein may appear first at to be counterintuitive. Flexible regions tend to populate additional states more readily (especially upon an increase in thermal energy) and thus, would potentially possess larger than average Cp values. In addition, several studies indicate that the N-terminal region of ubiquitin, specifically the α-helix and the turn between β-strands 1 and 2, is quite resilient to temperature fluctuations36, 37 and comprises the core structural elements that form first during ubiquitin folding.38, 39 However, investigations into the temperature dependence of NH order parameters for several thermo-stable proteins reveal some of the larger Cp values to be located within the more rigid regions of secondary structure.1113 Furthermore, the magnitude of Cp depends on the change in Sconf [see Eq. (7)], and, as calculated by three independent groups,7, 8, 11 the steepest change for Sconf occurs as the order parameter, S2, varies between 0.7 and 0.95. Perhaps the high melting temperature of ∼363 K for ubiquitin35 and other thermo-stabile proteins can be attributed in part to the relatively large local Cp of the methyl groups and NH bonds in the folded proteins reducing the magnitude of ΔCp upon unfolding.

One methyl group possesses a slightly negative Cp, I36γ2. Two other studies have reported negative Cp for NH bonds located in secondary structural regions.12, 15 Enhancements in the strength of hydrophobic interactions with increasing temperature were given as a possible explanation for this phenomenon.40 In the case of I36γ2, previous work has demonstrated that the hydrophobic interaction between the side chain of I36 and I30 at the C-terminus of the α-helix is important for the stability of ubiquitin.41 Thus, the negative Cp signifying a decrease in the flexibility and a reduction in the available conformational space for this methyl group may be related to the strengthening of this important interaction with increasing temperature.

Conclusion

In this study, we analyzed the temperature dependence for the methyl group order parameters in ubiquitin. With the requirement of only one 2D constant time 13C, 1H HSQC measurement without decoupling of 1H during 13C-chemical shift evolution, we have significantly reduced the total amount of experimental time required to obtain this information. From these experiments, the extracted thermal coefficients provide a glimpse into the location of ubiquitin thermo-stability near the N-terminus of the protein. Clearly, methyl groups contribute quite appreciably to the total heat capacity of ubiquitin, as well as of other proteins,9, 15 through the regulation of local conformational entropy. Taken together with the temperature dependence of NH order parameters, a per residue gauge of local protein thermodynamics offers a powerful complement to the already well-established methods for determining global thermodynamic parameters in proteins.

Materials and Methods

NMR spectroscopy and processing

Uniformly 15N,13C-labeled human ubiquitin was expressed and purified as described previously.42 For the NMR measurements, a 3.6 mM sample of the protein was prepared in 350 μL of 50 mM sodium phosphate buffer, 100 mM NaCl, 0.1% NaN3, and pH 6.8 in 10% D2O/90% H2O. A series of 2D constant time 13C, 1H HSQCs without decoupling of 1H during 13C-chemical shift evolution were measured at 14 temperature points: 275, 278, 281, 283, 286, 288, 291, 293, 296, 298, 301, 303, 305, and 308 K. All NMR experiments were performed in succession on a 700 MHz Avance-III Bruker spectrometer equipped with a triple resonance probe head. The constant time duration and INEPT delays were set to 27.8 and 2 ms, respectively. The spectrum was recorded with 1024 and 128 complex points in the direct (t2) and indirect (t1) dimensions, respectively, with eight scans per t1 increment. The t1,max and t2,max were 24.3 ms and 113 ms, respectively. Frequency discrimination in the indirectly detected dimension was achieved with the States-TPPI scheme.43 Each measurement required 59 min.

All time domain data were processed in the same manner with NMRPipe software.44 The data were zero-filled to 8 and 16 k in t1 and t2, respectively. After implementing a time domain solvent correction, a sine-bell window function was applied in the direct dimension, followed by Fourier transformation of t2. For the indirect dimension, a mirror image linear prediction algorithm was used to increase the resolution in t1. Next, a Gaussian window function was employed followed by Fourier transformation of t1. Finally, a polynomial baseline correction in the frequency domain was applied in the direct dimension.

Methods for error determination

The intensities of each peak in the quartet were taken from the program CARA45 and the dipolar-dipolar CCR rates (σobs) were calculated according to Eq. (3). To determine the errors in σobs, the noise levels (q) for each measurement were estimated using NMRPipe and the error (Δσobs) was obtained from the following relation,

graphic file with name pro0021-0562-m8.jpg (8)

where Δ is the length of the constant time period and Ii is the intensity of each peak in the quartet (Inline graphic). For the methyl group order parameters (Inline graphic), the errors (ΔInline graphic) were calculated by propagating the error from Δσobs using the equation,

graphic file with name pro0021-0562-m9.jpg (9)

where σrigid is derived from Eq. (5). Finally, the errors in the conformational entropy (ΔSconf) were determined by propagating the error from ΔInline graphic,

graphic file with name pro0021-0562-m10.jpg (10)

where

graphic file with name pro0021-0562-m11.jpg (11)
graphic file with name pro0021-0562-m12.jpg (12)

kB is the Boltzmann constant and NA is Avogadro's number. To estimate the error in the temperature dependency of Inline graphic(Inline graphic), the characteristic thermal coefficient (Inline graphic), and the heat capacity (Inline graphic), Monte Carlo simulations on 500 randomly generated data sets were performed.

Supplementary material

Additional Supporting Information may be found in the online version of this article.

pro0021-0562-SD1.pdf (741.1KB, pdf)

References

  • 1.Gomez J, Hilser VJ, Xie D, Freire E. The heat-capacity of proteins. Proteins. 1995;22:404–412. doi: 10.1002/prot.340220410. [DOI] [PubMed] [Google Scholar]
  • 2.Prabhu NV, Sharp KA. Heat capacity in proteins. Ann Rev Phys Chem. 2005;56:521–548. doi: 10.1146/annurev.physchem.56.092503.141202. [DOI] [PubMed] [Google Scholar]
  • 3.Kay LE, Torchia DA, Bax A. Backbone dynamics of proteins as studied by 15N inverse detected heteronuclear NMR-spectroscopy-application to Staphylococcal nuclease. Biochemistry. 1989;28:8972–8979. doi: 10.1021/bi00449a003. [DOI] [PubMed] [Google Scholar]
  • 4.Lipari G, Szabo A. Model-free approach to the interpretation of nuclear magnetic-resonance relaxation in macromolecules. 1. Theory and range of validity. J Am Chem Soc. 1982;104:4546–4559. [Google Scholar]
  • 5.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]
  • 6.Akke M, Bruschweiler R, Palmer AG. NMR order parameters and free-energy-an analytical approach and its application to cooperative Ca2+ binding by calbindin-D(9k) J Am Chem Soc. 1993;115:9832–9833. [Google Scholar]
  • 7.Li ZG, 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]
  • 8.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]
  • 9.Yang DW, Mok YK, Forman-Kay JD, Farrow NA, Kay LE. Contributions to protein entropy and heat capacity from bond vector motions measured by NMR spin relaxation. J Mol Biol. 1997;272:790–804. doi: 10.1006/jmbi.1997.1285. [DOI] [PubMed] [Google Scholar]
  • 10.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]
  • 11.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]
  • 12.Led JJ, Vinther JM, Kristensen SM. Enhanced stability of a protein with increasing temperature. J Am Chem Soc. 2011;133:271–278. doi: 10.1021/ja105388k. [DOI] [PubMed] [Google Scholar]
  • 13.Stone MJ, Seewald MJ, Pichumani K, Stowell C, Tibbals BV, Regan L. The role of backbone conformational heat capacity in protein stability: Temperature dependent dynamics of the B1 domain of Streptococcal protein G. Protein Sci. 2000;9:1177–1193. doi: 10.1110/ps.9.6.1177. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Wand AJ, Lee AL. Microscopic origins of entropy, heat capacity and the glass transition in proteins. Nature. 2001;411:501–504. doi: 10.1038/35078119. [DOI] [PubMed] [Google Scholar]
  • 15.Lee AL, Sharp KA, Kranz JK, Song XJ, Wand AJ. Temperature dependence of the internal dynamics of a calmodulin-peptide complex. Biochemistry. 2002;41:13814–13825. doi: 10.1021/bi026380d. [DOI] [PubMed] [Google Scholar]
  • 16.Wand AJ, Song XJ, Flynn PF, Sharp KA. Temperature dependence of fast dynamics in proteins. Biophys J. 2007;92:L43–L45. doi: 10.1529/biophysj.106.102160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Muhandiram DR, Yamazaki T, Sykes BD, Kay LE. Measurement of 2H T1 and T1p relaxation-times in uniformly 13C-labeled and fractionally 2H-labeled proteins in solution. J Am Chem Soc. 1995;117:11536–11544. [Google Scholar]
  • 18.Liao X, Long D, Li DW, Brüschweiler R, Tugarinov V. Probing side-chain dynamics in proteins by the measurement of nine deuterium relaxation rates per methyl group. J Phys Chem B. 2012;116:606–620. doi: 10.1021/jp209304c. [DOI] [PubMed] [Google Scholar]
  • 19.Millet O, Muhandiram DR, Skrynnikov NR, Kay LE. Deuterium spin probes of side-chain dynamics in proteins. 1. Measurement of five relaxation rates per deuteron in 13C-labeled and fractionally 2H-enriched proteins in solution. J Am Chem Soc. 2002;124:6439–6448. doi: 10.1021/ja012497y. [DOI] [PubMed] [Google Scholar]
  • 20.Liu WD, Zheng Y, Cistola DP, Yang DW. Measurement of methyl 13C-1H cross-correlation in uniformly 13C, 15N-labeled proteins. J Biomol NMR. 2003;27:351–364. doi: 10.1023/a:1025884922203. [DOI] [PubMed] [Google Scholar]
  • 21.Zhang X, Sui XG, Yang DW. Probing methyl dynamics from 13C autocorrelated and cross-correlated relaxation. J Am Chem Soc. 2006;128:5073–5081. doi: 10.1021/ja057579r. [DOI] [PubMed] [Google Scholar]
  • 22.Kay LE, Bull TE, Nicholson LK, Griesinger C, Schwalbe H, Bax A, Torchia DA. The measurement of heteronuclear transverse relaxation-times in Ax3 spin systems via polarization-transfer techniques. J Magn Reson. 1992;100:538–558. [Google Scholar]
  • 23.Müller N, Bodenhausen G, Ernst RR. Relaxation-induced violations of coherence transfer selection-rules in nuclear-magnetic-resonance. J Magn Reson. 1987;75:297–334. [Google Scholar]
  • 24.Tugarinov V, Hwang PM, Ollerenshaw JE, Kay LE. Cross-correlated relaxation enhanced 1H–13C NMR spectroscopy of methyl groups in very high molecular weight proteins and protein complexes. J Am Chem Soc. 2003;125:10420–10428. doi: 10.1021/ja030153x. [DOI] [PubMed] [Google Scholar]
  • 25.Kay LE, Torchia DA. The effects of dipolar cross-correlation on 13C methyl-carbon T1, T2, and NOE measurements in macromolecules. J Magn Reson. 1991;95:536–547. [Google Scholar]
  • 26.Lee AL, Flynn PF, Wand AJ. Comparison of 2H and 13C NMR relaxation techniques for the study of protein methyl group dynamics in solution. J Am Chem Soc. 1999;121:2891–2902. [Google Scholar]
  • 27.Goto NK, Gardner KH, Mueller GA, Willis RC, Kay LE. A robust and cost-effective method for the production of Val, Leu, Ile (δ1) methyl-protonated 15N-, 13C-, 2H-labeled proteins. J Biomol NMR. 1999;13:369–374. doi: 10.1023/a:1008393201236. [DOI] [PubMed] [Google Scholar]
  • 28.Godoy-Ruiz R, Guo C, Tugarinov V. Alanine methyl groups as NMR probes of molecular structure and dynamics in high-molecular-weight proteins. J Am Chem Soc. 2010;132:18340–18350. doi: 10.1021/ja1083656. [DOI] [PubMed] [Google Scholar]
  • 29.Ruschak AM, Velyvis A, Kay LE. A simple strategy for 13C, 1H labeling at the Ile-γ2 methyl position in highly deuterated proteins. J Biomol NMR. 2010;48:129–135. doi: 10.1007/s10858-010-9449-1. [DOI] [PubMed] [Google Scholar]
  • 30.Gans P, Hamelin O, Sounier R, Ayala I, Dura MA, Amero CD, Noirclerc-Savoye M, Franzetti B, Plevin MJ, Boisbouvier J. Stereospecific isotopic labeling of methyl groups for NMR spectroscopic studies of high-molecular-weight proteins. Angew Chem Int Ed Engl. 2010;49:1958–1962. doi: 10.1002/anie.200905660. [DOI] [PubMed] [Google Scholar]
  • 31.Lee D, Hilty C, Wider G, Wüthrich K. Effective rotational correlation times of proteins from NMR relaxation interference. J Magn Reson. 2006;178:72–76. doi: 10.1016/j.jmr.2005.08.014. [DOI] [PubMed] [Google Scholar]
  • 32.Fares C, Lakomek NA, Walter KFA, Frank BTC, Meiler J, Becker S, Griesinger C. Accessing ns-μs side chain dynamics in ubiquitin with methyl RDCs. J Biomol NMR. 2009;45:23–44. doi: 10.1007/s10858-009-9354-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Vijaykumar S, Bugg CE, Cook WJ. Structure of ubiquitin refined at 1.8 Å resolution. J Mol Biol. 1987;194:531–544. doi: 10.1016/0022-2836(87)90679-6. [DOI] [PubMed] [Google Scholar]
  • 34.Privalov PL, Gill SJ. Stability of protein-structure and hydrophobic interaction. Adv Protein Chem. 1988;39:191–234. doi: 10.1016/s0065-3233(08)60377-0. [DOI] [PubMed] [Google Scholar]
  • 35.Wintrode PL, Makhatadze GI, Privalov PL. Thermodynamics of ubiquitin unfolding. Proteins. 1994;18:246–253. doi: 10.1002/prot.340180305. [DOI] [PubMed] [Google Scholar]
  • 36.Wand AJ, Babu CR, Hilser VJ. Direct access to the cooperative substructure of proteins and the protein ensemble via cold denaturation. Nat Struct Mol Biol. 2004;11:352–357. doi: 10.1038/nsmb739. [DOI] [PubMed] [Google Scholar]
  • 37.Grzesiek S, Cordier F. Temperature-dependence properties as studied by of protein hydrogen bond high-resolution NMR. J Mol Biol. 2002;317:739–752. doi: 10.1006/jmbi.2002.5446. [DOI] [PubMed] [Google Scholar]
  • 38.Jackson SE, Went HM. Ubiquitin folds through a highly polarized transition state. Protein Eng Des Sel. 2005;18:229–237. doi: 10.1093/protein/gzi025. [DOI] [PubMed] [Google Scholar]
  • 39.Brutscher B, Brüschweiler R, Ernst RR. Backbone dynamics and structural characterization of the partially folded A state of ubiquitin by 1H, 13C, and 15N nuclear magnetic resonance spectroscopy. Biochemistry. 1997;36:13043–13053. doi: 10.1021/bi971538t. [DOI] [PubMed] [Google Scholar]
  • 40.Baldwin RL. Temperature-dependence of the hydrophobic interaction in protein folding. Proc Natl Acad Sci USA. 1986;83:8069–8072. doi: 10.1073/pnas.83.21.8069. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Thomas ST, Makhatadze GI. Contribution of the 30/36 hydrophobic contact at the C-terminus of the alpha-helix to the stability of the ubiquitin molecule. Biochemistry. 2000;39:10275–10283. doi: 10.1021/bi0000418. [DOI] [PubMed] [Google Scholar]
  • 42.Handel TM, Johnson EC, Lazar GA, Desjarlais JR. Solution structure and dynamics of a designed hydrophobic core variant of ubiquitin. Structure. 1999;7:967–976. doi: 10.1016/s0969-2126(99)80123-3. [DOI] [PubMed] [Google Scholar]
  • 43.Marion D, Ikura M, Tschudin R, Bax A. Rapid recording of 2D NMR-spectra without phase cycling: Application to the study of hydrogen-exchange in proteins. J Magn Reson. 1989;85:393–399. [Google Scholar]
  • 44.Delaglio F, Grzesiek S, Vuister GW, Zhu G, Pfeifer J, Bax A. Nmrpipe-a multidimensional spectral processing system based on Unix pipes. J Biomol NMR. 1995;6:277–293. doi: 10.1007/BF00197809. [DOI] [PubMed] [Google Scholar]
  • 45.Keller R. Computer Aided Resonance Assignment. 2004. http://cara.nmr.ch/
  • 46.Koradi R, Billeter M, Wüthrich K. MOLMOL: A program for display and analysis of macromolecular structures. J Mol Graph. 1996;14:51–55. doi: 10.1016/0263-7855(96)00009-4. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

pro0021-0562-SD1.pdf (741.1KB, pdf)

Articles from Protein Science : A Publication of the Protein Society are provided here courtesy of The Protein Society

RESOURCES