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 15 毫秒
1.
The solution structure of reduced Clostridium pasteurianum rubredoxin (MW 6100) is reported here. The protein is highly paramagnetic, with iron(II) being in the S=2 spin state. The Hβ protons of the ligating cysteines are barely observed, and not specifically assigned. Seventy-six percent of the protons have been assigned and 1267 NOESY peaks (of which 1037 are meaningful) have been observed. Nonselective T 1 measurements have been measured by recording four nonselective 180°-τ-NOESY at different τ values, and fitting the intensity recoveries to an exponential recovery. Thirty-six metal-proton upper and lower distance constraints have been obtained from the above measurements. The use of such constraints is assessed with respect to spin delocalization on the sulfur donor atoms. The solution structure obtained with the program DYANA has been refined through restrained energy minimization. A final family of 20 conformers is obtained with no distance violations larger than 0.24?Å, and RMSD values to the mean structure of 0.58 and 1.03?Å for backbone and all heavy atoms, respectively (measured on residues 3–53). The structure is compared to the X-ray structure of the oxidized and of the zinc substituted protein, and to the available structures of other rubredoxins. In particular, the comparison with the crystal structure and the solution structure of the Zn derivative of the highly thermostable Pyrococcus furiosus rubredoxin suggested that the relatively low thermal stability of the clostridial rubredoxin may be tentatively ascribed to the loosening of its secondary structure elements. This research is a further achievement at the frontier of solution structure determinations of paramagnetic proteins.  相似文献   

2.
The traditional NMR‐based method for determining oligomeric protein structure usually involves distinguishing and assigning intra‐ and intersubunit NOEs. This task becomes challenging when determining symmetric homo‐dimer structures because NOE cross‐peaks from a given pair of protons occur at the same position whether intra‐ or intersubunit in origin. While there are isotope‐filtering strategies for distinguishing intra from intermolecular NOE interactions in these cases, they are laborious and often prove ineffectual in cases of weak dimers, where observation of intermolecular NOEs is rare. Here, we present an efficient procedure for weak dimer structure determination based on residual dipolar couplings (RDCs), chemical shift changes upon dilution, and paramagnetic surface perturbations. This procedure is applied to the Northeast Structural Genomics Consortium protein target, SeR13, a negatively charged Staphylococcus epidermidis dimeric protein (Kd 3.4 ± 1.4 mM) composed of 86 amino acids. A structure determination for the monomeric form using traditional NMR methods is presented, followed by a dimer structure determination using docking under orientation constraints from RDCs data, and scoring under residue pair potentials and shape‐based predictions of RDCs. Validation using paramagnetic surface perturbation and chemical shift perturbation data acquired on sample dilution is also presented. The general utility of the dimer structure determination procedure and the possible relevance of SeR13 dimer formation are discussed.  相似文献   

3.
The solution structure of crambin has been refined using a direct nuclear Overhauser effect (NOE) simulation approach (DINOSAUR) following a slow-cooling simulated annealing protocol starting from eight previously obtained nmr and the x-ray structures of crambin. Theoretical NOE intensities calculated with inclusion of local motions were directly compared to the experimental nmr data and forces were derived using a simple first-order approximation for the calculation of the NOE gradient. A dynamic assignment procedure was applied for the peaks involving unassigned diastereotopic proton pairs or equivalent aromatic protons. With this approach, R factors could be minimized in a reasonable simulation time to low values (around 0.26) while deviations from ideal bond lengths and angles are still acceptable. The improvement in R factors is accompanied by an improvement of the precision of the structures, the rms deviations (rmsd; from the average) calculated on the ensemble of nine structures decreasing from 0.65 to 0.55 Å for backbone atoms and from 1.0 to 0.85 Å for all heavy atoms. The solution structure is significantly different from the x-ray structure with rmsd for all atoms of 1.35 Å compared to 0.85 Å between solution structures. The largest differences are found for residues Thr-21 and Pro-22 in the loop region between the two α-helices and for the side chain of Tyr-29. © 1994 John Wiley & Sons, Inc.  相似文献   

4.
M J Sutcliffe  C M Dobson 《Proteins》1991,10(2):117-129
The effect of including paramagnetic relaxation data as additional restraints in the determination of protein tertiary structures from NMR data has been explored by a systematic series of model calculations. The system used for testing the method was the 2.0 A resolution tetragonal crystal structure of hen egg white lysozyme (129 amino acid residues) and structures were generated using a version of the hybrid "distance geometry-dynamic simulated annealing" procedure. A limited set of 769 NOEs was used as restraints in all the calculations; the strengths of these were categorized into three classes on the basis of distances observed in the crystal structure. The values of 50 phi angles were also restrained on the basis of amide-alpha coupling constants calculated from the X-ray structure. Five sets of 12 structures were determined using differing sets of paramagnetic relaxation data as restraints additional to those involving the NOE and coupling constant data. The paramagnetic relaxation data were modeled on the basis of the distances of defined protons from the crystallographic binding site of Gd3+ in lysozyme. Analysis of the results showed that the relaxation data significantly improved the correspondence between the set of generated structures and the crystal structure, and that the more well defined the relaxation data, the more significant the improvement in the quality of the structures. The results suggest that the inclusion of paramagnetic relaxation restraints could be of significant value for the experimental determination of protein structures from NMR data.  相似文献   

5.
The three-dimensional structure in aqueous solution of native huwentoxin-I, a neurotoxin from the venom of the spider Selenocosmia huwena, has been determined from two-dimensional 1H NMR data recorded at 500 and 600 MHz. Structural constraints consisting of interproton distances inferred from NOEs and dihedral angles from spin–spin coupling constants were used as input for distance geometry calculation with the program XPLOR 3.1. The best 10 structures have NOE violations <0.3 Å, dihedral violations <2°, and pairwise root-mean-square differences of 1.08 (±0.20) Å over backbone atoms (N, Cα;, C). The molecule adopts a compact structure consisting of a small triple-stranded antiparallel β-sheet and five β-turns. A small hydrophobic patch consisting of Phe 6, Trp 28, and Trp 31 is located on one side of the molecule. All six lysine residues are distributed on the molecular surface. The three disulfidc bridges are buried within the molecule. The structure contains an “inhibitor cystine knot motif” which is adopted by several other small proteins, such as ω-conotoxin, agatoxin IVA, and gurmarin.  相似文献   

6.
Iturin A is a lipopeptide extracted from strains of Bacillus subtilis. Seven peptide residues form a cycle closed by a β-amino acid carrying a hydrophobic tail. This compound is an antifungal and induces the formation of conducting pores in black lipid membranes. Two-dimensional 1H-nmr was used for investigating its conformation in pyridine. A complete set of nuclear Overhauser effects (NOEs) was obtained from which interproton distances were deduced in a rather broad range of 2.2–4.2 Å. A special procedure was then used to optimize simultaneously experimental parameters and intramolecular energy calculated by semiempirical methods. A model of the conformation is proposed for the backbone for which there is an excellent coherence between NOEs, coupling constants, and intramolecular energy. The conformations of Asn, Gln, and Ser side chains appear to be much more flexible because of their interactions with the solvent. From this picture, iturin A seems to have a rather stiff ring surrounded by mobile side chains. Further studies of this lipopeptide and of other members of the family should enable us to approach some structure–activity relationships for this class of antibiotics.  相似文献   

7.
The solvent structure and behavior around a protein were examined by analyzing a trajectory of molecular dynamics simulation of thetrp-holorepressor in a periodic box of water. The calculated selfdiffusion coefficient indicated that the solvent within 10 Å of the protein had lower mobility. Examination of the solvent diffusion around different atoms of different kinds of residues showed no general tendency. Thisfact suggested that the solvent mobility is not influenced significantly bythe kind of the atom or residue they solvated. Distribution analysis aroundthe protein revealed two peaks of water oxygen: a sharp one at 2.8 Å around polar and charged atoms and a broad one at ~3.4 Å aroundapolar atoms. The former was stabilized by water–protein hydrogen bonds, and the latter was stabilized by water-lwater hydrogen bonds, suggesting the existence of a hydrophobic shell. An analysis of protein atom–water radial distribution functions confirmed these shell structures around polar or charged atoms and apolar ones. © 1993 Wiley-Liss, Inc.  相似文献   

8.
The structure and internal motions of the protein hen egg white lysozyme are studied by analysis of simulation and experimental data. A molecular dynamics simulation and an energy minimization of the protein in vacuum have been made and the results compared with high-resolution structures and temperature factors of hen egg white lysozyme in two different crystal forms and of the homologous protein human lysozyme. The structures obtained from molecular dynamics and energy minimization have root-mean-square deviations for backbone atoms of 2.3 Å and 1.1–1.3 Å, respectively, relative to the crystal structures; the different crystal structures have root-mean-square deviations of 0.73–0.81 Å for the backbone atoms. In comparing the backbone dihedral angles, the difference between the dynamics and the crystal structure on which it is based is the same as that between any two crystal structures. The internal fluctuations of atomic positions calculated from the molecular dynamics trajectory agree well with the temperature factors from the three structures. Simulation and crystal results both show that there are large motions for residues involved in exposed turns of the backbone chain, relatively smaller motions for residues involved in the middle of helices or β-sheet structures, and relatively small motions of residues near disulfide bridges. Also, both the simulation and crystal data show that side-chain atoms have larger fluctuations than main-chain atoms. Moreover, the regions that have large deviations among the x-ray crystal structures, which indicates flexibility, are found to have large fluctuations in the simulation.  相似文献   

9.
Standard procedures for using nuclear Overhauser enhancements (NOE) between protons to generate structures for diamagnetic proteins in solution from NMR data may be supplemented by using dipolar shifts if the protein is paramagnetic. This is advantageous since the electron-nuclear dipolar coupling provides relatively long-range geometric information with respect to the paramagnetic centre which complements the short-range distance constraints from NOEs. Several different strategies have been developed to date, but none of these attempts to combine data from NOEs and dipolar shifts in the initial stages of structure calculation or to determine three dimensional protein structures together with their magnetic properties. This work shows that the magnetic and atomic structures are highly correlated and that it is important to have additional constraints both to provide starting parameters for the magnetic properties and to improve the definition of the best fit. Useful parameters can be obtained for haem proteins from Fermi contact shifts; this approach is compared with a new method based on the analysis of dipolar shifts in haem methyl groups with respect to data from horse and tuna ferricytochromes c. The methods developed for using data from NOEs and dipolar shifts have been incorporated in a new computer program, PARADYANA, which is demonstrated in application to a model data set for the sequence of the haem octapeptide known as microperoxidase-8. Received: 13 October 1997 / Accepted: 19 December 1997  相似文献   

10.
Li Q  Khosla C  Puglisi JD  Liu CW 《Biochemistry》2003,42(16):4648-4657
During polyketide biosynthesis, acyl carrier proteins (ACPs) perform the central role of transferring polyketide intermediates between active sites of polyketide synthase. The 4'-phosphopantetheine prosthetic group of a holo-ACP is a long and flexible arm that can reach into different active sites and provide a terminal sulfhydryl group for the attachment of acyl groups through a thioester linkage. We have determined the solution structure and characterized backbone dynamics of the holo form of the frenolicin acyl carrier protein (fren holo-ACP) by nuclear magnetic resonance (NMR). Unambiguous assignments were made for 433 hydrogen atoms, 333 carbon atoms, and 84 nitrogen atoms, representing a total of 94.6% of the assignable atoms in this protein. From 879 meaningful NOEs and 45 angle constraints, a family of 24 structures has been calculated. The solution structure is composed of three major alpha-helices packed in a bundle with three additional short helices in intervening loops; one of the short helices slowly exchanges between two conformations. Superposition of the major helical regions on the mean structure yields average atomic rmsd values of 0.49 +/- 0.09 and 0.91 +/- 0.08 A for backbone and non-hydrogen atoms, respectively. Although the three-helix bundle fold is conserved among acyl carrier proteins involved in fatty acid synthases and polyketide synthases, a detailed comparison revealed that ACPs from polyketide biosynthetic pathways are more related to each other in tertiary fold than to their homologues from fatty acid biosynthetic pathways. Comparison of the free form of ACPs (NMR structures of fren ACP and the Bacillus subtilis ACP) with the substrate-bound form of ACP (crystal structure of butyryl-ACP from Escherichia coli) suggests that conformational exchange plays a role in substrate binding.  相似文献   

11.
A model of nine proteins including side-chain atoms have been built from the known Cα coordinates and amino acid sequences using a Monte Carlo Protein Building Annealing method. The Cartesian coordinates for the side-chain atoms were established with bond lengths and angles selected randomly from within previously determined ranges. A simulated annealing technique is used to generate some 300 structures with differing side-chain conformations. The atomic coordinates of the backbone atoms are fixed during the simulated annealing process. The coordinates of the side-chain atoms of 300 low energy conformations are averaged to obtain a mean structure that is minimized with the Cα atoms constrained to their position in the x-ray structure using the OPLS/AMBER force field with the GB/SA water model. The rms deviation of the main-chain atoms (without Cβ) compared with the corresponding crystal structures is in the range 0.20–0.64 Å. The rms deviation of the side-chain atoms is between 1.72 and 2.71 Å and for all atoms is between 1.19 and 1.99 Å. The method is insensitive to random errors in the Cα positions and the computational requirement is modest. © 1997 John Wiley & Sons, Inc.  相似文献   

12.
Abstract

As part of our on-going development of a method, based upon distance geometry calculations, for predicting the structures of proteins from the known structures of their homologues, we have predicted the structure of the 176 residue Flavodoxin from Escherichia coli. This prediction was based upon the crystal structures of the homologous Flavodoxins from Anacystis nidulans, Chondrus crispus, Desulfovibrio vulgaris and Clostridium beijerinckii, whose sequence identities with Escherichia coli were 44%, 33%, 23% and 16%, respectively. A total of 13,043 distance constraints among the alpha-carbons of the Escherichia coli structure were derived from the sequence alignments with the known structures, together with 8,893 distance constraints among backbone and sidechain atoms of adjacent residues, 978 between the alpha-carbons and selected atoms of the flavin mononucleotide cofactor, 116 constraints to enforce conserved hydrogen bonds, and 452 constraints on the torsion angles in conserved residues. An ensemble of ten random Escherichia coli structures was computed from these constraints, with an average root mean square coordinate deviation (RMSD) among the alpha carbons of 0.85 Ångstroms (excluding the first 1 and last 6 residues, which have no corresponding residues in any of the homologues and hence were unconstrained); the corresponding average heavy-atom RMSD was 1.60 Å.

Since the distance geometry calculations were performed without hydrogen atoms, protons were added to the resulting structures and these structures embedded in a 50 × 50 × 40 Å solvent box with periodic boundary conditions. They were then subjected to a 20 picosecond dynamical simulated annealing procedure, starting at 300 K and gradually reduced to 10K, in which all the distance and torsion angle constraints were maintained by means of harmonic restraint functions. This was followed up by 1000 iterations of unrestrained conjugate gradients minimization. The goal of this energy refinement procedure was not to drastically modify the structures in an attempt at a priori prediction, but merely to improve upon the predictions obtained from the geometric constraints, particularly with regard to their local backbone and sidechain conformations and their hydrogen bonds. The resulting structures differed from the respective starting structures by an average of 1.52 Å in their heavy atom RMSD's, while the average RMSD among the heavy atoms of residues 2-170 increased slightly to 1.66 Å. We hope these structures will be good enough to enable the phase problem to be solved for the crystallographic data that is now being collected on this protein.  相似文献   

13.
The structure in solution of crambin, a small protein of 46 residues, has been determined from 2D NMR data using an iterative relaxation matrix approach (IRMA) together with distance geometry, distance bound driven dynamics, molecular dynamics, and energy minimization. A new protocol based on an “ensemble” approach is proposed and compared to the more standard initial rate analysis approach and a “single structure” relaxation matrix approach. The effects of fast local motions are included and R-factor calculations are performed on NOE build-ups to describe the quality of agreement between theory and experiment. A new method for stereospecific assignment of prochiral groups, based on a comparison of theoretical and experimental NOE intensities, has been applied. The solution structure of crambin could be determined with a precision (rmsd from the average structure) of 0.7 Å on backbone atoms and 1.1 Å on all heavy atoms and is largely similar to the crystal structure with a small difference observed in the position of the side chain of Tyr-29 which is determined in solution by both J-coupling and NOE data. Regions of higher structural variability (suggesting higher mobility) are found hi the solution structure, in particular for the loop between the two helices (Gly-20 to Pro-22). © 1993 Wiley-Liss, Inc.  相似文献   

14.
Stereo-specific 1H nuclear magnetic resonance assignments can be obtained following distance geometry structure calculations. The key to this method is to allow stereo-related atoms or methyls to float between pro-R and pro-S configurations, the final configuration being determined by the experimental constraints. Resonances from stereo-related pairs are given initial random assignments (either pro-R or pro-S) for identifying nuclear Overhauser effects (NOEs). A list of distance constraints using these assignments is compiled and a series of structures calculated where the chirality of non-C alpha chiral centers is not constrained; no pseudoatom corrections are required. Calculated structures are both locally and globally well-determined since the assignments rely upon the structure determination rather than the structure quality relying upon stereo-specific assignments. The method represents a global approach to determining stereo-specific assignments versus previously reported methods where only intraresidue NOEs and J-coupling information are used.  相似文献   

15.
Here we present a novel NMR method for the structure determination of calcium-calmodulin (Ca2+-CaM)-peptide complexes from a limited set of experimental restraints. A comparison of solved CaM-peptide structures reveals invariability in CaM’s backbone conformation and a structural plasticity in CaM’s domain orientation enabled by a flexible linker. Knowing this, the collection and analysis of an extensive set of NOESY spectra is redundant. Although RDCs can define CaM domain orientation in the complex, they lack the translational information required to position the domains on the bound peptide and highlight the necessity of intermolecular NOEs. Here we employ a specific isotope labeling strategy in which the role of methionine in CaM-peptide interactions is exploited to collect these critical NOEs. By 1H, 13C-labeling the methyl groups of deuterated methionine against a 2H, 12C background, we can acquire a 13C-edited NOESY characterized by simplified, easily analyzable spectra. Together with measured CaM backbone HN-N RDCs and intrapeptide NOE-based distances, these intermolecular NOEs provide restraints for a low temperature torsion-angle dynamics and simulated annealing protocol used to calculate the complex structure. We have applied our method to a CaM complex previously solved through X-ray crystallography: Ca2+-CaM bound to the CaM kinase I peptide (PDB code: 1MXE). The resulting structure has a backbone RMSD of 1.6 Å to that previously published. We have also used this test complex to investigate the importance of homologous model selection on the calculated outcome. In addition to having application for fast complex structure determination, this method can be used to determine the structures of difficult complexes characterized by chemical shift overlap and broad signals for which the traditional method based on the use of fully 13C, 15N-labeled CaM fails.  相似文献   

16.
Abstract

We have calculated the deoxyribose sugar energy for a wide range of puckering parameters, (q, W), using different force fields. The intra-ring bond lengths, bond angles, and dihedral angles are calculated for every energy minimized structure and compared with 224 sugar ring structures available from DNA single crystal x-ray data. A modified Weiner's force field yields an excellent agreement with x-ray data.

The calculated energy surface shows a variable amplitude repuckering path, with an average distortion of 0.42 Å. Most of the experimental values of (q, W) fall within 1.0 Kcal/mol from the calculated minimum.  相似文献   

17.
Turkey egg-white lysozyme differs from hen egg-white lysozyme in its primary structure in 7 of the 129 residues. We have determined the rotational and translational parameters relating the known co-ordinates of hen egg-white lysozyme molecule to the turkey lysozyme. The rotational parameters were determined using the rotation function, the translational parameters were determined by placing the properly rotated molecule systematically at all positions within the unit cell and searching for those positions producing few intermolecular contacts between the α-carbon atoms of one molecule and all its neighbors. These parameters were refined by minimizing the conventional R factor between observed and calculated structure amplitudes. The final rotational and translational parameters give an R value of 46.7% for reflections with d spacings between 6 Å and 12 Å and have 7 intermolecular contacts closer than 5 Å between the a carbon atoms of one molecule and all its neighbors. An electron density map has been calculated at 5 Å resolution; the packing of the molecules in this form appears to present the entire length of the active cleft in the vicinity of the crystallographic 6-fold axis and does not appear to be blocked by neighboring molecules.  相似文献   

18.
Abstract

This paper describes two complexes containing N,N-dimethylproflavine and the dinucleoside monophosphate, 5-iodocytidylyl(3′-5′)guanosine (iodoCpG). The first complex is triclinic, space group PI, with unit cell dimensions a = 11.78 Å, b = 14.55 Å, c = 15.50 Å, a = 89.2°, β = 86.2°, γ = 96.4°. The second complex is monoclinic, space group P21, with a = 14.20 Å, b = 19.00 Å, c = 20.73 Å, β = 103.6°. Both structures have been solved to atomic resolution and refined by Fourier and least squares methods. The first structure has been refined anisotropically to a residual of 0.09 on 5,025 observed reflections using block diagonal least squares, while the second structure has been refined isotropically to a residual of 0.13 on 2,888 reflections with full matrix least squares. The asymmetric unit in both structures contains two dimethylproflavine molecules and two iodoCpG molecules; the first structure has 16 water molecules (a total of 134 non-hydrogen atoms), while the second structure has 18 water molecules (a total of 136 non-hydrogen atoms). Both structures demonstrate intercalation of dimethylproflavine between base-paired iodoCpG dimers. In addition, dimethylproflavine molecules stack on either side of the intercalated duplex, being related by a unit cell translation along b and a axes, respectively.

The basic structural feature of the sugar-phosphate chains accompanying dimethylproflavine intercalation in both structures is the mixed sugar puckering pattern: C3′ endo (3′-5′) C2′ endo. This same structural information is again demonstrated in the accompanying paper, which describes a complex containing dimethylproflavine with deoxyribo-CpG.

Similar information has already appeared for other “simple” intercalators such as ethidium, acridine orange, ellipticine, 9-aminoacridine, N-methyl-tetramethylphenanthrolinium and terpyridine platinum. “Complex” intercalators, however, such as proflavine and daunomycin, have given different structural information in model studies. We discuss the possible reasons for these differences in this paper and in the accompanying paper.  相似文献   

19.
Abstract

This paper describes two complexes containing ethidium and the dinucleoside monophosphate, cytidylyl(3′-5′)guanosine (CpG). Both crystals are monoclinic, space group P21, with unit cell dimensions as follows: modification 1: a = 13.64 Å, b = 32.16 Å, c - 14.93 Å, β = 114.8° and modification 2: a = 13.79 Å, b = 31.94 Å, c = 15.66 Å, β = 117.5°. Each structure has been solved to atomic resolution and refined by Fourier and least squares methods; the first has been refined to a residual of 0.187 on 1,903 reflections, while the second has been refined to a residual of 0.187 on 1,001 reflections. The asymmetric unit in both structures contains two ethidium molecules and two CpG molecules; the first structure has 30 water molecules (a total of 158 non-hydrogen atoms), while the second structure has 19 water molecules (a total of 147 non-hydrogen atoms). Both structures demonstrate intercalation of ethidium between base-paired CpG dimers. In addition, ethidium molecules stack on either side of the intercalated duplex, being related by a unit cell translation along the a axis.

The basic feature of the sugar-phosphate chains accompanying ethidium intercalation in both structures is: C3′ endo (3′-5′) C2′ endo. This mixed sugar-puckering pattern has been observed in all previous studies of ethidium intercalation and is a feature common to other drug-nucleic acid structural studies carried out in our laboratory. We discuss this further in this paper and in the accompanying papers.  相似文献   

20.
The structure of Peptide T was determined by solution NMR spectroscopy, under strong structure‐inducing conditions: 40% hexafluoro‐2‐propanol aqueous solution at 5 °C. Under these conditions it was possible to detect medium‐range NOEs for the first time for this peptide. This allowed a much better‐defined structure to be determined for Peptide T in comparison with earlier NMR and computational studies. Peptide structures consistent with the experimental restraints were generated using a restrained MD simulation with a full empirical force field. Residues 4–8 of Peptide T take on a well‐defined structure with a heavy atom RMSD of 0.78 Å. The structure is stabilized by hydrogen bonding to side‐chain oxygen atoms of Thr 4 and Thr 8, as well as backbone hydrogen bonding between residues 5 and 7 that forms this region into a classic γ‐turn. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  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