首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Proton magnetic resonance data and conformational calculations of a series of model compounds containing a NH-CαH group substituted as in peptides have been used to generate a proton–proton coupling constant–dihedral angle relation for the peptide unit. For those substances used in which the dihedral angle about the N-Cα bond is not fixed, the angle distribution was calculated from conformational theory. Using eight examples in which the number of theoretical assumptions were least, the best values of the coefficients A, B, and C in the expression J(θ) = Acos2θ + B cosθ + Csin2θ were found by a least-squares procedure to be 7.9, ?1.55, and 1.35, respectively. This relation gives reasonable values for the dihedral angles ? in cyclic oligopeptide structures for which the availability of both NMR data and other structural information allow comparison. When applied to N-acetylamino acid N-methylamides having side chains extending beyond Cβ, however, agreement with the calculated conformational distribution was found for Leu, Met, and Trp, but observed values of J were larger than expected for Val, He, Phe, and Tyr, These disagreements are considered to be the result of interactions not yet taken into account in the usual conformational calculations.  相似文献   

2.
Geminal two‐bond couplings (2J) in proteins were analyzed in terms of correlation with protein secondary structure. NMR coupling constants measured and evaluated for a total six proteins comprise 3999 values of 2JCαN′, 2JC′HN, 2JHNCα, 2JC′Cα, 2JHαC′, 2JHαCα, 2JCβC′, 2JN′Hα, 2JN′Cβ, and 2JN′C′, encompassing an aggregate 969 amino‐acid residues. A seamless chain of pattern comparisons across the spectrum datasets recorded allowed the absolute signs of all 2J coupling constants studied to be retrieved. Grouped by their mediating nucleus, C′, N′ or Cα, 2J couplings related to C′ and N′ depend significantly on ?,ψ torsion‐angle combinations. β turn types I, I′, II and II′, especially, can be distinguished on the basis of relative‐value patterns of 2JCαN′, 2JHNCα, 2JC′HN, and 2JHαC′. These coupling types also depend on planar or tetrahedral bond angles, whereas such dependences seem insignificant for other types. 2JHαCβ appears to depend on amino‐acid type only, showing negligible correlation with torsion‐angle geometry. Owing to its unusual properties, 2JCαN′ can be considered a “one‐bond” rather than two‐bond interaction, the allylic analog of 1JN′Cα, as it were. Of all protein J coupling types, 2JCαN′ exhibits the strongest dependence on molecular conformation, and among the 2J types, 2JHNCα comes second in terms of significance, yet was hitherto barely attended to in protein structure work. Proteins 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

3.
M. T. Cung  M. Marraud 《Biopolymers》1982,21(5):953-967
We use the nmr data concerning the CαH? CβH fragment in eight peptides with rigid side chains to parametrize a Karplus correlation between the vicinal proton Jαβ coupling constant and the dihedral angle θ. When considering molecules containing the fragment CαHα? CβHβHβ′, the three-dimensional structure of the model peptides does not need to be known with accurate precision, since each set of Jαβ and Jαβ′ coupling constants is then related to the coefficients of the Karplus equation. A good correlation is observed with the Karplus equation, which is in substantial agreement with the Jαβ coupling constants reported for rigid as well as rotating Cα? Cβ bonds in peptides.  相似文献   

4.
NMR coupling constants, both direct one‐bond (1J) and geminal two‐bond (2J), are employed to analyze the protein secondary structure of human oxidized ERp18. Coupling constants collected and evaluated for the 18 kDa protein comprise 1268 values of 1JCαHα, 1JCαCβ, 1JCαC′, 1JC′N′, 1JN′Cα, 1JN′HN, 2JCαN′, 2JHNCα, 2JC′HN, and 2JHαC′. Comparison with 1J and 2J data from reference proteins and pattern analysis on a per‐residue basis permitted main‐chain φ,ψ torsion‐angle combinations of many of the 149 amino‐acid residues in ERp18 to be narrowed to particular secondary‐structure motifs. J‐coupling indexing is here being developed on statistical criteria and used to devise a ternary grid for interpreting patterns of relative values of J. To account for the influence of the varying substituent pattern in different amino‐acid sidechains, a table of residue‐type specific threshold values was compiled for discriminating small, medium, and large categories of J. For the 15‐residue insertion that distinguishes the ERp18 fold from that of thioredoxin, the J‐coupling data hint at a succession of five isolated Type‐I β turns at progressively shorter sequence intervals, in agreement with the crystal structure. Proteins 2011. © 2010 Wiley‐Liss, Inc.  相似文献   

5.
Based on a Cambridge Structural Database (CSD) search, a meta‐analysis of 116 structures of alanine H3NCαH(CH3)C′(O)O and its derivatives H3NCαH(CH3)C′(O)O(H/R/M), protonated, esterified, or coordinated at the carboxylic group, shows that in the first step of a chirality chain, the L configuration at Cα induces (M) and (P) conformations with respect to rotation around the central C′─Cα bond. In the second step, the (M) and (P) conformations selectively distort the planar carboxylic group CαC'(Ocis)Otrans to asymmetric flat (R) and (S) tetrahedra. High diastereoselectivities are caused by the two players attraction N…Ocis and repulsion Otrans…CMe, which work together in (L,M,R) configurations but against each other in (L,P,S) configurations.  相似文献   

6.
A coupling constant-dihedral angle correlation for the H? Cα? Cβ? H system of amino acid residues in peptides has been derived from a set of model compounds covering the full range of dihedral angles. The expression obtained, J = 11.0 cos2 θ ?1.4 cos θ + 1.6 sin2θ, is close to those already used in pmr studies of peptide conformation, and provides a firmer foundation for them. A factor limiting the precision of this and other “Karplus relations” is illustrated.  相似文献   

7.
1H and 13C high-resolution nmr spectra of cationic, zwitterionic, and anionic forms of the peptides: H-Trp-(Pro)n-Tyr-OH, n = 0-5, and H-Trp-Pro-OCH3 were obtained in D2O solution. Analysis of Hα(Pro1), Hα(Trp), Cγ(Pro), Hε(Tyr), and Hδ(Trp) resonances provided evidence for the presence of two predominant backbone isomers: the all-trans one and another with the Trp-Pro peptide bond in cis conformation; the latter constituted about 0.8 molar fraction of the total peptide (n > 1) concentration. Relative content of these isomers varied in a characteristic way with the number of Pro residues and the ionization state of the peptides. The highest content of the cis (Trp-Pro) isomer, 0.74, was found in the anionic form of H-Trp-Pro-Tyr-OH; it decreased in the order of: anion ? zwitterion ≈ cation, and with the number of Pro residues to reach the value of 0.42 in the cationic form of H-Trp- (Pro)5-Tyr-OH. Isomerization equilibria about Pro-Pro bond(s) were found to be shifted far (?0.9) in favor of the trans conformation. Interpretation of the measured vicinal coupling constants Jα?β′ and Jα?β″ for CαH-CβH2 proton systems of Trp and Tyr side chains in terms of relative populations of g+, g?, and t staggered rotamers around the χ1 dihedral angle indicated that in all the peptides studied (a) rotation of Trp indole ring in cis (Trp-Pro) isomers is strongly restricted, and (b) rotation of Tyr phenol ring is relatively free. The most preferred χ1 rotamer of Trp (0.8-0.9 molar fraction) was assigned as the t one on the basis of a large value of the vicinal coupling constant between the high-field Hβ and carbonyl carbon atoms of Trp, estimated for the cis (Pro1) form of H-Trp-Pro-Tyr-OH from a 1H, 13C correlated spectroscopy 1H detected multiple quantum experiment. This indicates that cis ? trans equilibrium in the Trp-Pro fragment is governed by nonbonding interactions between the pyrrolidine (Pro) and indole (Trp) rings. A molecular model of the terminal cis Trp-Pro dipeptide fragment is proposed, based on the presented nmr data and the results of our molecular mechanics modeling of low-energy conformers of the peptides, reported elsewhere. © 1993 John Wiley & Sons, Inc.  相似文献   

8.
A quantitative J-correlation pulse sequence is described that allows simultaneous determination of one-bond and two-bond nitrogen-carbon coupling constants for protonated or deuterated proteins. Coupling constants are calculated from volume ratios between cross peaks and reference axial peaks observed in a single 3D spectrum. Accurate backbone 1 J NC, 1 J NC, and 2 J NC coupling constants are obtained for the two [15N;13C]-labeled, medium-sized proteins flavodoxin and xylanase and for the [2H;15N;13C]-labeled, large protein DFPase. A dependence of one-bond and two-bond J NC values on protein backbone torsion angles is readily apparent, in agreement with previously found correlations. In addition, the experiment is performed on isotropic as well as aligned protein to measure associated 15N-13C residual dipolar couplings.  相似文献   

9.
The geminal and vicinal 13C-31P coupling constants have been monitored, as a function of pH, for a series of uracil and cytosine 3′- and 5′-nucleotides with a ribose, arabinose, or 2′-deoxyribose sugar. Data were also obtained for two 3′,5′-diphosphates in the ribose and arabinose series. The geminal J(C5′-P5′) and J(C3′-P3′) couplings show only a small dependence on the ionization state of the phosphate, decreasing by < 0.5 Hz in the pH 5–7 range. For the ribose and arabinose 3′-nucleotides, the vicinal J(C4′-P3′) increase (up to 1.5 Hz) on secondary phosphate ionization in the pH 5–7 range, whereas their J(C2′-P3′) couplings decrease (up to 1.5 Hz) over the same pH range. In contrast for the 2′-deoxyribose molecules, both couplings decrease (~0.5 Hz) on phosphate ionization. The titration curves provide information about the influence of the sugar on the conformation about the C3′? O3′ bond. Some conformational trends could be rationalized by consideration of the sugar-puckerdependent contact interactions between the 3′-phosphate and the substituents on the furanose ring.  相似文献   

10.
11.
Quantitative φ-dihedral angle determinations of non-glycine and non-proline residues in Desulfovibrio vulgaris flavodoxin are carried out on the exclusive basis of 3 J coupling constants. In total 124 3 JHNH α , 123 3 JHNC ′i , 118 3 JHNC β , 117 3 JC′ i–1Hα , 109 3 JC′ i–1C′i , and 103 3 JC′ i–1Hβ values form the experimental basis for translating J coupling data into geometry information using various combinations of Karplus parameters given in the literature. In addition, each backbone torsional angle φ is adjusted assuming different models of local geometry, either a rigid torsion, a Gaussian distribution centered at a distinct angle, or a two-site jump model. Numerical optimization is followed by a statistical significance evaluation to assess the results. It is found that experimental coupling constants of most of the residues involved in secondary structure elements agree best with those predicted from rigid local conformations. For dihedral angles in loop regions, mobility effects are not negligible, and a single torsion (Glu 42) is likely to adopt two distinct adjustments. However, α-helix conformations with –60° < φ < –45° give rise to an alternate solution with φ≈+170° with similar statistical significance when using the four traditionally determined proton-involved 3 J couplings. This ambiguity is efficiently avoided only when taking advantage of the complete data set comprising six available experimental 3 J coupling constants and of the degeneracy intrinsic to the Karplus relation. The optimized φ conformations are compared with reference values from the crystal structure of flavodoxin.  相似文献   

12.
R Tewari  R K Nanda  G Govil 《Biopolymers》1974,13(10):2015-2035
The probability distributions of the torsional angles (Φ′, ω′, ω, Φ, and ψ), which fix the structure of nucleotide backbone, have been calculated using the results of energy calculations based on extended Huckel theory (EHT), complete neglect of differential overlap (CNDO), perturbative configuration interaction using localized orbitals (PCILO), and classical potential functions (CPF) methods. Statistical average values of the vicinal 1H? 1H, 1H? 31P, and 13C? 31P nmr coupling constants 〈J〉 have been calculated from the generalized Karplus relations using the probability distribution in the Φ′, Φ, and ψ space. Experimental 〈J〉 values for polyribouridylic acid (polyU) support the theoretical predictions for these torsional angles. Using Monte Carlo technique, random coils of single-stranded polynucleotides have been simulated and the mean-square end-to-end distance 〈r2〉 has been calculated. Molecular orbital methods (EHT, CNDO, and PCILO) suggest considerable flexibility around O? P bonds, leading to fairly small values for the characteristic ratio (C ~ 4). Observed values of the unperturbed characteristic ratio for polynucleotides are quite large (C ~ 18) suggesting a relatively rigid nucleotide backbone. The results based on molecular orbital calculations can be reconciled with the experimental values by introducing an additional stabilization of ~2 kcal mol?1 for the predicted minimum energy ragion (Φ′ ~ 240°, ω′ ~ 290°, ω 290°, Φ 180°, and ψ 60°). Such a stabilization may arise from the association of water molecules and metal ions with the phosphate group and (or) Coulomb interaction between neighboring phosphate groups. The calculations provide a semiquantitative estimate of torsional rigidity in the nucleotide backbone.  相似文献   

13.
Methionine residues fulfill a broad range of roles in protein function related to conformational plasticity, ligand binding, and sensing/mediating the effects of oxidative stress. A high degree of internal mobility, intrinsic detection sensitivity of the methyl group, and low copy number have made methionine labeling a popular approach for NMR investigation of selectively labeled protein macromolecules. However, selective labeling approaches are subject to more limited information content. In order to optimize the information available from such studies, we have performed DFT calculations on model systems to evaluate the conformational dependence of 3 J CSCC, 3 J CSCH, and the isotropic shielding, σiso. Results have been compared with experimental data reported in the literature, as well as data obtained on [methyl-13C]methionine and on model compounds. These studies indicate that relative to oxygen, the presence of the sulfur atom in the coupling pathway results in a significantly smaller coupling constant, 3 J CSCC/3 J COCC ~ 0.7. It is further demonstrated that the 3 J CSCH coupling constant depends primarily on the subtended CSCH dihedral angle, and secondarily on the CSCC dihedral angle. Comparison of theoretical shielding calculations with the experimental shift range of the methyl group for methionine residues in proteins supports the conclusion that the intra-residue conformationally-dependent shift perturbation is the dominant determinant of δ13Cε. Analysis of calmodulin data based on these calculations indicates that several residues adopt non-standard rotamers characterized by very large ~100° χ3 values. The utility of the δ13Cε as a basis for estimating the gauche/trans ratio for χ3 is evaluated, and physical and technical factors that limit the accuracy of both the NMR and crystallographic analyses are discussed.  相似文献   

14.
Based on a CSD search, a meta‐analysis of 1179 structures of 19 natural amino acids H3NCαH(R)C′(O)O and their derivatives H3NCαH(R)C′(O)O(H/R/M), protonated, esterified, or coordinated at the carboxylic group, shows that the chirality chain with its two steps, established in the preceding paper for alanine, can be extended to natural amino acids. High diastereoselectivities are observed in the induction from the L configuration at Cα to the ?ψ and +ψ conformations, which in turn distort the planar carboxylic group CαC′(Ocis)Otrans to asymmetric flat tetrahedra, showing that the chirality chain is an integral part of natural amino acids.  相似文献   

15.
P Gupta-Bhaya 《Biopolymers》1975,14(6):1143-1160
The electron-mediated spin–spin coupling constant J between the amide NH and the α-CH protons in the dipeptide fragment Cα? CO(NH? CαH)R? C′ONH? Cα is dependent on the dihedral angle of rotation (Φ) around the N? C bond. Measurement of J in a series of zwitterionic dipeptides H3N+? CHR1? CONH? CHR2? CO2? (which is conformationally similar to the dipeptide fragment) in TFA solution shows that J is independent of R1, but dependent on the steric bulk of R2. The data are interpreted in terms of a model that assumes that what we measure is an average value of J? a thermal average over all the possible rotamers. The groups R1 and R2 are, in most cases, sterically kept apart by the trans and planar amide bonds, and hence the independence of J of R1. This model is consistent with the theoretical calculations done on the dipeptide fragment. The effect of the structural characteristics of the side chains (e.g., the effect of lengthening and branching the side chains) on the J values in dipeptides is discussed in the light of the existing results of theoretical calculations. Study of 〈J〉 values in tripeptides (C6H5CH2OCONH? CHR1? CONH? CHR2? CO2CH3, essentially three linked peptide units) shows that electrostatic interaction between the two amide bonds modifies the potential energy surface and the 〈J〉 value of a dipeptide subunit in the tripeptides. Also in some cases, direct steric interaction between the two side chains in the two adjacent dipeptide subunits in the tripeptide affects the potential energy surfaces of the individual dipeptide subunits and hence the 〈J〉 values. The influence of the structural characteristics of the side chains of individual amino acids on structure formation at or beyond the dipeptide level is discussed at various points. The J(NH? αCH) values of CH3CONH? CHR? CONH2 and CH3CONH? CHR? CO2CH3 with the same R are quite different for R = valine, leucine, phenylalanine, methionine, but equal for R = glycine. This, coupled with the fact that one of the carboxamide NH resonances has a chemical shift different from its counterpart in simple amides like CH3CONH2 and the other carboxamide NH has the same chemical shift as its counterpart in CH3CONH2, suggest the presence of a hydrogen bond in dipeptide CH3CONH? CHR? CONH2 with carboxamide NH as the donor. Theoretical evidence for two seven-membered hydrogen-bonded rings with the carboxamide NH as donor and the acetyl oxygen as acceptor is summarized. Our data cannot suggest the number of such hydrogen-bonded rings, nor can they conclude the relative proportion of these rings in a particular dipeptide. A discussion of the difficulty of interpretation is presented and the data are discussed under certain simplifying assumptions.  相似文献   

16.
A survey has been made, by using computer methods, of the types of helices which polypeptide chains can form, taking into account steric requirements and intramolecular hydrogen-bonding interactions. The influence on these two requirements, of small variations in the bond angles of the peptide residues, or of small changes in the overall dimensions of the helix (pitch and residues per turn), have been assessed for the special case of the α-helix. Criteria for the formation of acceptable hydrogen bonds have also been applied to helices of other types, viz., the 3, γ?, ω?, and π-helices. It was shown that the N? H … O and H … O? C angles in hydrogen bonds are sensitive to changes in either the NCαC′ bond angle or in the rotational angles about the N? Cα and Cα? C′ bonds. However, the variants of the α-helix observed experimentally in myoglobin can all be constructed without distortion of the hydrogen bonds. For α-helices, the steric and hydrogen bonding requirements are more easily fulfilled with an NCαC′ bond angle of 111°, rather than 109.5°. The decreased stability observed for the left-handed α-helix relative to the right-handed one for L -amino acids is due essentially only to interactions of the Cβ atom of the side chains with atoms in adjacent peptide units in the backbone, and interactions with atoms in adjacent turns of the helical backbone are not significantly different in the two helices. Restrictions in the freedom of rotation of bulky side chains may have significant kinetic effects during the formation of the α-helix from the “random coil” state.  相似文献   

17.
Pressure-induced changes in 3h J NC scalar couplings through hydrogen bonds were investigated in the immunoglobulin binding domain of streptococcal protein G. 1H, 15N and 13C triple-resonance NMR spectroscopy coupled with the on-line high pressure cell technique was used to monitor 3h J NC scalar couplings at 30 and 2000 bar in uniformly labeled 15N and 13C protein isotopes. Both increased and decreased 3h J NC scalar couplings were observed at high pressure. No correlation with secondary structure was apparent. The difference in coupling constants as well as pressure-induced chemical shift data suggests a compaction of the helix ends and an increase of the helix pitch at its center in response to pressure. Our data provides the first direct evidence that the electronic orbital overlap in protein backbone hydrogen bonds is altered by pressure.  相似文献   

18.
Described here is a set of three-dimensional (3D) NMR experiments that rely on CACA-TOCSY magnetization transfer via the weak 3 \textJ\textCa\textCa ^{ 3} {\text{J}}_{{{\text{C}}\alpha {\text{C}}\alpha }} coupling. These pulse sequences, which resemble recently described 13C detected CACA-TOCSY (Takeuchi et al. 2010) experiments, are recorded in 1H2O, and use 1H excitation and detection. These experiments require alternate 13C-12C labeling together with perdeuteration, which allows utilizing the small 3 \textJ\textCa\textCa ^{ 3} {\text{J}}_{{{\text{C}}\alpha {\text{C}}\alpha }} scalar coupling that is otherwise masked by the stronger 1JCC couplings in uniformly 13C labeled samples. These new experiments provide a unique assignment ladder-mark that yields bidirectional supra-sequential information and can readily straddle proline residues. Unlike the conventional HNCA experiment, which contains only sequential information to the 1 3 \textCa ^{ 1 3} {\text{C}}^{\alpha } of the preceding residue, the 3D hnCA-TOCSY-caNH experiment can yield sequential correlations to alpha carbons in positions i1, i + 1 and i2. Furthermore, the 3D hNca-TOCSY-caNH and Hnca-TOCSY-caNH experiments, which share the same magnetization pathway but use a different chemical shift encoding, directly couple the 15N-1H spin pair of residue i to adjacent amide protons and nitrogens at positions i2, i1, i + 1 and i + 2, respectively. These new experimental features make protein backbone assignments more robust by reducing the degeneracy problem associated with the conventional 3D NMR experiments.  相似文献   

19.
Methyl α-cellobioside (methyl β-d-glucopyranosyl-(1→4)-α-d-glucopyranoside) was labeled with 13C at C4′ for use in NMR studies in DMSO-d6 solvent to attempt the detection of a trans-H-bond J-coupling (3hJCCOH) between C4′ and OH3. Analysis of the OH3 signal at 600 MHz revealed only the presence of two homonuclear J-couplings: 3JH3,OH3 and a smaller, longer range JHH. No evidence for 3hJC4′,OH3 was found. The longer range JHH was traced to 4JH4,OH3 based on 2D 1H–1H COSY data and inspection of the H2 and H4 signal lineshapes. A limited set of DFT calculations was performed on a methyl cellobioside mimic to evaluate the structural dependencies of 4JH2,O3H and 4JH4,O3H on the H3–C3–O3–H torsion angle. Computed couplings range from about −0.7 to about +1.1 Hz, with maximal values observed when the C–H and O–H bonds are roughly diaxial.  相似文献   

20.
Nucleic acid duplexes featuring a single alpha-anomeric thymidine inserted into each DNA strand via 3′-3′ and 5′-5′ phosphodiester linkages exhibit local conformational dynamics that are not adequately depicted by conventional restrained molecular dynamics (rMD) methods. We have used molecular dynamics with time-averaged NMR restraints (MDtar) to explore its applicability to describing the conformational dynamics of two α-containing duplexes – d(GCGAAT-3′-3′-αT-5′-5′-CGC)2 and d(ATGG-3′-3′-αT-5′-5′-GCTC)?r(gagcaccau). In contrast to rMD, enforcing NOE-based distance restraints over a period of time in MDtar rather than instantaneously results in better agreement with the experimental NOE and J-data. This conclusion is based on the dramatic decreases in average distance and coupling constant violations (Δd av, J rms, and ΔJ av) and improvements in sixth-root R-factors (R x). In both duplexes, the deoxyribose ring puckering behavior predicted independently by pseudorotation analysis is portrayed remarkably well using this approach compared to rMD. This indicates that the local dynamic behavior is encoded within the NOE data, although this is not obvious from the local R x values. In both systems, the backbone torsion angles comprising the 3′-3′ linkage as well as the (high S-) sugars of the α-nucleotide and preceding residue (α?1) are relatively static, while the conformations of the 5′-5′ linkage and the sugar in the neighboring β-nucleotide (α+1) show enhanced flexibility. To reduce the large ensembles generated by MDtar to more manageable clusters we utilized the PDQPRO program. The resulting PDQPRO clusters (in both cases, 13 structures and associated probabilities extracted from a pool of 300 structures) adequately represent the structural and dynamic characteristics predicted by the experimental data.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号